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.
The effects of exposure of cultured P388D1 cells to H2O2 on intracellular free calcium ([Ca++]i) was investigated utilizing the intracellular fluorescent calcium chelator "Quin 2." [Ca++]i rose from approximately 150 nM to greater than 2 microM over a time course that was strongly dependent on the concentration of H2O2 used (5 X 10(-5) to 5 X 10(-3) M). After exposure of P388D1 cells to 5 X 10(-3) M H2O2, Quin 2 was fully saturated between 15 and 30 min exposure. During this time, no apparent change in the rate of equilibration of 45Ca++ from the extracellular medium could be detected, whereas in cells preloaded with 45Ca, net 45Ca was lost from the cells at a greater rate than controls. Measurements of total cellular calcium by atomic absorption spectroscopy confirmed that there was a net loss of calcium from the cells during the first 30 min. At time points greater than 45 min after exposure to H2O2 the influx of extracellular 45Ca and net intracellular Ca++, Na+ and K+ rapidly increased. Half times for H2O2 catabolism by the cells varied from about 8 min at 5.0 X 10(-4) M H2O2 to 14.0 min at 5.0 X 10(-3) M. When the total [Ca++]i-buffering capacity of the Quin 2 pool was varied by increasing the loading of intracellular Quin 2 by 68-fold (1.1 X 10(2) - 7.6 X 10(3) amol per cell), the rate of rise of [Ca++]i was depressed by only 1.6-fold following exposure to 5 mM H2O2. During the rise of intracellular [Ca++]i, cell morphology was observed by both light and scanning electron microscopy and revealed that "surface blebs" appeared during this phase of injury. Both the rise in [Ca++]i and "blebbing" were observable before any loss in cell viability was detected by either loss of Trypan blue exclusion or loss of preloaded 51Cr from the cells. From these results we conclude the following, H2O2 exposure induces a dose-dependent disturbance of intracellular calcium homeostatis; the rise in [Ca++]i is mediated by exposure to H2O2 in the early phase of the injury, and is not dependent on the continuing presence of the oxidant; the rate of rise of [Ca++]i is largely independent of the quantity of calcium mobilized to the Quin 2 pool; during the early phase (less than 30 min) of rise of [Ca++]i, only intracellular calcium is involved in the response; these events occur concomitantly with gross morphological changes to the plasma membrane.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

2.
3.
Prior heavy exercise markedly alters the O2 uptake (VO2) response to subsequent heavy exercise. However, the time required for VO2 to return to its normal profile following prior heavy exercise is not known. Therefore, we examined the VO2 responses to repeated bouts of heavy exercise separated by five different recovery durations. On separate occasions, nine male subjects completed two 6-min bouts of heavy cycle exercise separated by 10, 20, 30, 45, or 60 min of passive recovery. The second-by-second VO2 responses were modeled using nonlinear regression. Prior heavy exercise had no effect on the primary VO2 time constant (from 25.9 +/- 4.7 s to 23.9 +/- 8.8 s after 10 min of recovery; P = 0.338), but it increased the primary VO2 amplitude (from 2.42 +/- 0.39 to 2.53 +/- 0.41 l/min after 10 min of recovery; P = 0.001) and reduced the VO2 slow component (from 0.44 +/- 0.13 to 0.21 +/- 0.12 l/min after 10 min of recovery; P < 0.001). The increased primary amplitude was also evident after 20-45 min, but not after 60 min, of recovery. The increase in the primary VO2 amplitude was accompanied by an increased baseline blood lactate concentration (to 5.1 +/- 1.0 mM after 10 min of recovery; P < 0.001). Baseline blood lactate concentration was still elevated after 20-60 min of recovery. The priming effect of prior heavy exercise on the VO2 response persists for at least 45 min, although the mechanism underpinning the effect remains obscure.  相似文献   

4.
Mouse peritoneal macrophages respond to environmental stimuli in different ways depending on their state of differentiation. Macrophages from mice with bacillus Calmette--Guerin (BCG) infection produced large amounts of H2O2 in response to phorbol diesters (PDEs), while those from noninfected mice produced little or no H2O2. The effects of PDEs on cells are mediated by specific cellular receptors for these ligands. The purpose of this study was to determine if the varying responses of macrophages from different groups of mice were caused by differences in their receptors for the PDE ligands. By all parameters studied, the binding of [20-3H]phorbol 12,13-dibutyrate ( [3H]PDBu) was similar in all macrophages irrespective of their ability to produce H2O2 in response to PDEs. Binding of [3H]PDBu was rapid at 23 degrees C reaching a maximum at 10-20 min with a subsequent decline to 50-60% of maximum by 30-60 min. Binding was slower at 0 degrees C reaching a maximum at 90-120 min. The binding was reversible, with dissociation kinetics paralleling association kinetics. The binding was saturable; the Kd's (45 to 91 nM) and number of binding sites (about 7-14 X 10(5)/cell or 11-12 pmol/mg protein) were essentially the same for the different classes of macrophages. The binding was specific, and analogs of PDBu inhibited [3H]PDBu binding to macrophages with potencies comparable to their potencies in causing in vivo tumor promotion and elicitation of other cellular responses in vitro. The ligands [3H]PDBu and [3H]PMA were degraded to comparable degrees by macrophages from normal or BCG-infected mice. Macrophages from C3H/HeJ and C3H/HeN mice, although known to differ in their abilities to respond to stimuli such as lymphokines and LPS, did not differ in their ability to produce H2O2 in response to PDEs or in their receptors for PDEs. Results of this study suggest that in vivo "activation" of macrophages in mice infected with BCG is not associated with a change in the cells' receptors for PDEs, but may be associated with "postreceptor" changes such as linkage of the PDE receptor with NAD(P)H oxidase, a change in NAD(P)H oxidase, or induction of synthesis of NAD(P)H oxidase.  相似文献   

5.
The mitochondria-rich epithelial cells of the renal medullary thick ascending limb (mTAL) reabsorb nearly 25% of filtered sodium (Na(+)) and are a major source of cellular reactive oxygen species. Although we have shown that delivery of Na(+) to the mTAL of rats increases superoxide (O(2)(·-)) production in mTAL, little is known about H(2)O(2) production, given the lack of robust and selective fluorescent indicators for determining changes within the whole cell, specifically in the mitochondria. The present study determined the effect of increased tubular flow and Na(+) delivery to mTAL on the production of mitochondrial H(2)O(2) in mTAL. H(2)O(2) responses were determined in isolated, perfused mTAL of Sprague-Dawley rats using a novel mitochondrial selective fluorescent H(2)O(2) indicator, mitochondria peroxy yellow 1, and a novel, highly sensitive and stable cytosolic-localized H(2)O(2) indicator, peroxyfluor-6 acetoxymethyl ester. The results showed that mitochondrial H(2)O(2) and cellular fluorescent signals increased progressively over a period of 30 min following increased tubular perfusion (5-20 nl/min), reaching levels of statistical significance at ~10-12 min. Responses were inhibited with rotenone or antimycin A (inhibitors of the electron-transport chain), polyethylene glycol-catalase and by reducing Na(+) transport with furosemide or ouabain. Inhibition of membrane NADPH-oxidase with apocynin had no effect on mitochondrial H(2)O(2) production. Cytoplasmic H(2)O(2) (peroxyfluor-6 acetoxymethyl ester) increased in parallel with mitochondrial H(2)O(2) (mitochondria peroxy yellow 1) and was partially attenuated (~65%) by rotenone and completely inhibited by apocynin. The present data provide clear evidence that H(2)O(2) is produced in the mitochondria in response to increased flow and delivery of Na(+) to the mTAL, and that whole cell H(2)O(2) levels are triggered by the mitochondrial reactive oxygen species production. The mitochondrial production of H(2)O(2) may represent an important target for development of more effective antioxidant therapies.  相似文献   

6.
We investigated the effect of IL-2 in the isolated guinea pig lung perfused with phosphate-buffered Ringer's solution (containing 0.5 g/100 ml albumin and 5.5 mM dextrose) to determine the mechanism of IL-2-induced pulmonary edema. IL-2 (0 to 10,000 U/ml) was added to the perfusate following a 10 min baseline steady-state period. Pulmonary arterial pressure (Ppa), pulmonary capillary pressure (Ppc), and change in lung weight (as a measure of developing pulmonary edema) were recorded at 0, 10, 30, 40, and 60 min. The capillary filtration coefficient (Kf.c), an index of vascular permeability to water, was measured at 30 and 60 min. Infusion of IL-2 increased Ppc (from 3.9 +/- 0.1 cm H2O at baseline to 8.8 +/- 1.1 cm H2O at 60 min for IL-2 at 2000 U/ml, p less than 0.01; and from 3.8 +/- 0.1 cm H2O at baseline to 8.9 +/- 0.6 cm H2O at 60 min for IL-2 at 10,000 U/ml, p less than 0.01. The lung weight also increased (32% at IL-2 concentration of 2000 U/ml, and 26% at IL-2 concentration of 10,000 U/ml) The capillary filtration coefficient did not change with IL-2 infusion. The IL-2 response was prevented using the pulmonary vasodilator, papaverine. The infusion of IL-2 was associated with the generation of thromboxane A2(TxA2) in the effluent perfusate. Inhibition of TxA2 synthetase using Dazoxiben prevented the pulmonary vasoconstriction and edema response to IL-2. In addition, IL-2 had no effect on the transendothelial clearance of 125I-albumin. The results indicate that IL-2 causes pulmonary edema secondary to an increase in Ppc. The response is mediated by IL-2 stimulation of TxA2 generation from the lung.  相似文献   

7.
8.
Survival after H2O2 exposure or heat shock of asynchronous Chinese hamster ovary cells (HA-1) was assayed following pretreatment with mildly toxic doses of either H2O2 or hyperthermia. H2O2 cytotoxicity at 37 degrees C, expressed as a function of mM H2O2 was found to be dependent on cell density at the time of treatment. The density dependence reflected the ability of cells to reduce the effectiveness of H2O2 as a cytotoxic agent. When the survival data were plotted as a function of mumoles H2O2/cell at the beginning of the treatment, survival was independent of cell density. Cells pretreated with 0.1 mM (3-5 mumoles/cell X 10(-7)) H2O2 for 1 hr at 37 degrees C (30-50% survival) became resistant to a subsequent H2O2 treatment 16-36 hr after pretreatment [dose modifying factor (DMF) at 1% isosurvival = 4-6]. Their resistance to 43 degrees C heating, however, was only slightly increased over controls 16-36 hr following pretreatment (DMF at 1% isosurvival = 1.2). During this same interval, the synthesis of protein migrating in the 70 kD region of a one-dimensional SDS-polyacrylamide gel was enhanced twofold in the H2O2-pretreated cells. When the cells were heated for 15 min at 45 degrees C (40-60% survival), the survivors became extremely resistant to 43 degrees C heating and somewhat resistant to H2O2 (DMF at 1% isosurvival = 2). The heat-induced resistance to heat developed much more rapidly (reached a maximum between 6 and 13 hr) following pretreatment than the heat-induced resistance to H2O2 (16-36 hr). The enhanced synthesis of 70 kD protein after heat shock was greater in magnitude and occurred more rapidly following preheating than following H2O2 pretreatment. The cells that became resistant to H2O2 by either pretreatment (H2O2 or heat shock) also increased their ability to reduce the H2O2 cytotoxicity from the treatment medium beyond that of the untreated HA-1 cells. This may be one of the mechanisms involved in the increased resistance and a common adaptive mechanism induced by both stresses. These data indicate that mammalian cells develop resistance to H2O2 following mild pretreatment with H2O2 or heat shock. The cross-resistance induced by H2O2 and heat shock reinforce the hypothesis that some overlap in mechanisms exist between the cellular responses to these two stresses. However, the failure of H2O2 pretreatment to induce much resistance to heat indicates that there are also differences in the actions of the two agents.  相似文献   

9.
Prolonged exhaustive submaximal exercise in humans induces marked metabolic changes, but little is known about effects on muscle Na+-K+-ATPase activity and sarcoplasmic reticulum Ca2+ regulation. We therefore investigated whether these processes were impaired during cycling exercise at 74.3 +/- 1.2% maximal O2 uptake (mean +/- SE) continued until fatigue in eight healthy subjects (maximal O2 uptake of 3.93 +/- 0.69 l/min). A vastus lateralis muscle biopsy was taken at rest, at 10 and 45 min of exercise, and at fatigue. Muscle was analyzed for in vitro Na+-K+-ATPase activity [maximal K+-stimulated 3-O-methylfluorescein phosphatase (3-O-MFPase) activity], Na+-K+-ATPase content ([3H]ouabain binding sites), sarcoplasmic reticulum Ca2+ release rate induced by 4 chloro-m-cresol, and Ca2+ uptake rate. Cycling time to fatigue was 72.18 +/- 6.46 min. Muscle 3-O-MFPase activity (nmol.min(-1).g protein(-1)) fell from rest by 6.6 +/- 2.1% at 10 min (P <0.05), by 10.7 +/- 2.3% at 45 min (P <0.01), and by 12.6 +/- 1.6% at fatigue (P <0.01), whereas 3[H]ouabain binding site content was unchanged. Ca2+ release (mmol.min(-1).g protein(-1)) declined from rest by 10.0 +/- 3.8% at 45 min (P <0.05) and by 17.9 +/- 4.1% at fatigue (P < 0.01), whereas Ca2+ uptake rate fell from rest by 23.8 +/- 12.2% at fatigue (P=0.05). However, the decline in muscle 3-O-MFPase activity, Ca2+ uptake, and Ca2+ release were variable and not significantly correlated with time to fatigue. Thus prolonged exhaustive exercise impaired each of the maximal in vitro Na+-K+-ATPase activity, Ca2+ release, and Ca2+ uptake rates. This suggests that acutely downregulated muscle Na+, K+, and Ca2+ transport processes may be important factors in fatigue during prolonged exercise in humans.  相似文献   

10.
1,25-dihydroxyvitamin D3 (1,25-(OH)2D3) increases synthesis of heat shock proteins in monocytes and U937 cells and protects these cells from thermal injury. We examined whether 1,25-(OH)2D3 would also modulate the susceptibility of U937 cells to H2O2-induced oxidative stress. Cell viability was assessed by trypan blue exclusion and [3H]thymidine incorporation into DNA. Prior incubation for 24 h with 1,25-(OH)2D3 (25 pM or higher) unexpectedly increased H2O2 toxicity. Since cellular Ca2+ may be a mediator of cell injury we investigated effects of altering extracellular Ca2+ ([Ca2+]e) on 1,25-(OH)2D3-enhanced H2O2 toxicity as well as effects of 1,25-(OH)2D3 and H2O2 on cytosolic free Ca2+ concentration ([Ca2+]f). Basal [Ca2+]f in medium containing 1.5 mM Ca as determined by fura-2 fluorescence was higher in 1,25-(OH)2D3-pretreated cells than control cells (137 versus 112 nM, P less than 0.005). H2O2 induced a rapid increase in [Ca2+]f (to greater than 300 nM) in both 1,25-(OH)2D3-treated and control cells, which was prevented by a reduction in [Ca2+]e to less than basal [Ca2+]f. The 1,25(OH)2D3-induced increase in H2O2 toxicity was also prevented by preincubation with 1,25-(OH)2D3 in Ca2+-free medium or by exposing the cells to H2O2 in the presence of EGTA. Preexposure of cells to 45 degrees C for 20 min, 4 h earlier, partially prevented the toxic effects of H2O2 particularly in 1,25-(OH)2D3-treated cells, even in the presence of physiological levels of [Ca2+]e. Thus 1,25-(OH)2D3 potentiates H2O2-induced injury probably by increasing cellular Ca2+ stores. The 1,25-(OH)2D3-induced amplification of the heat shock response likely represents a mechanism for counteracting the Ca2+-associated enhanced susceptibility to oxidative injury due to 1,25-(OH)2D3.  相似文献   

11.
Maximal O2 delivery and O2 uptake (VO2) per 100 g of active muscle mass are far greater during knee extensor (KE) than during cycle exercise: 73 and 60 ml. min-1. 100 g-1 (2.4 kg of muscle) (R. S. Richardson, D. R. Knight, D. C. Poole, S. S. Kurdak, M. C. Hogan, B. Grassi, and P. D. Wagner. Am. J. Physiol. 268 (Heart Circ. Physiol. 37): H1453-H1461, 1995) and 28 and 25 ml. min-1. 100 g-1 (7.5 kg of muscle) (D. R. Knight, W. Schaffartzik, H. J. Guy, R. Predilleto, M. C. Hogan, and P. D. Wagner. J. Appl. Physiol. 75: 2586-2593, 1993), respectively. Although this is evidence of muscle O2 supply dependence in itself, it raises the following question: With such high O2 delivery in KE, are the quadriceps still O2 supply dependent at maximal exercise? To answer this question, seven trained subjects performed maximum KE exercise in hypoxia [0.12 inspired O2 fraction (FIO2)], normoxia (0.21 FIO2), and hyperoxia (1.0 FIO2) in a balanced order. The protocol (after warm-up) was a square wave to a previously determined maximum work rate followed by incremental stages to ensure that a true maximum was achieved under each condition. Direct measures of arterial and venous blood O2 concentration in combination with a thermodilution blood flow technique allowed the determination of O2 delivery and muscle VO2. Maximal O2 delivery increased with inspired O2: 1.3 +/- 0.1, 1.6 +/- 0.2, and 1.9 +/- 0.2 l/min at 0.12, 0.21, and 1.0 FIO2, respectively (P < 0.05). Maximal work rate was affected by variations in inspired O2 (-25 and +14% at 0.12 and 1.0 FIO2, respectively, compared with normoxia, P < 0.05) as was maximal VO2 (VO2 max): 1.04 +/- 0.13, 1. 24 +/- 0.16, and 1.45 +/- 0.19 l/min at 0.12, 0.21, and 1.0 FIO2, respectively (P < 0.05). Calculated mean capillary PO2 also varied with FIO2 (28.3 +/- 1.0, 34.8 +/- 2.0, and 40.7 +/- 1.9 Torr at 0.12, 0.21, and 1.0 FIO2, respectively, P < 0.05) and was proportionally related to changes in VO2 max, supporting our previous finding that a decrease in O2 supply will proportionately decrease muscle VO2 max. As even in the isolated quadriceps (where normoxic O2 delivery is the highest recorded in humans) an increase in O2 supply by hyperoxia allows the achievement of a greater VO2 max, we conclude that, in normoxic conditions of isolated KE exercise, KE VO2 max in trained subjects is not limited by mitochondrial metabolic rate but, rather, by O2 supply.  相似文献   

12.
Two kit preparations of the organometallic precursor [(188)Re(H(2)O)(3)(CO)(3)](+) in aqueous media are presented. Method A uses gaseous carbon monoxide and amine borane (BH(3).NH(3)) as the reducing agent. In method B CO(g) is replaced by K(2)[H(3)BCO(2)] that releases carbon monoxide during hydrolysis. Both procedures afford the desired precursor in yields >85% after 10 min at 60 degrees C. HPLC and TLC analyses revealed 7 +/- 3% of unreacted (188)ReO(4)(-) and <5% of colloidal (188)ReO(2). Solutions of up to 14 GBq/mL Re-188 have been successfully carbonylated with these two methods. The syntheses of two tailor-made bifunctional ligand systems for the precursor [(188)Re(H(2)O)(3)(CO)(3)](+) are presented. The tridentate chelates consist of a bis[imidazol-2-yl]methylamine or an iminodiacetic acid moiety, respectively. Both types of ligand systems have been prepared with alkyl spacers of different length and a pendent primary amino or carboxylic acid functionality, enabling the amidic linkage to biomolecules. The tridentate coordination of the ligands to the rhenium-tricarbonyl core could be elucidated on the macroscopic level by X-ray structure analyses and 1D and 2D NMR experiments of two representative model complexes. On the nca level, the ligands allow labeling yields >95% with [(188)Re(H(2)O)(3)(CO)(3)](+) under mild reaction conditions (PBS buffer, 60 degrees C, 60 min) at ligand concentrations between 5 x 10(-4) M and 5 x 10(-5) M. Thus, specific activities of 22-220 GBq pe micromol of ligand could be achieved. Incubation of the corresponding Re-188 complexes in human serum at 37 degrees C revealed stabilities between 80 +/- 4% and 45 +/- 10% at 24 h, respectively, and 63 +/- 3% and 34 +/- 3% at 48 h postincubation in human serum depending on the chelating system. Decomposition product was mainly (188)ReO(4)(-). The routine kit-preparation of the precursor [(188)Re(H(2)O)(3)(CO)(3)](+) in combination with tailor-made ligand systems enables the organometallic labeling of biomolecules with unprecedented high specific activities.  相似文献   

13.
3H-labelled metabolites were determined in the perchloric acid-soluble fraction of blood plasma and liver of adult male Wistar rats, following the application of [5 - 3H]uridine. Ten minutes after the injection of uridine, only 20% of the total 3H activity of the plasma could be attributed to [3H]uridine. The remaining radioactivity was found chiefly in [3H]uracil (40%) and 3H2O (20%). In the liver, at 10 min, [3H]-uridine and [3H]uracil together accounted for less than 0.5% of the total radioactivity; about 70% of the radioactivity was due to [3H]beta-alanine, and 15% to 3H2O. 45 min after the injection, 70% of the radioactivity in the plasma was due to 3H2O, whereas uridine and uracil represented about 4% and 6%, respectively. At this time, about 55% of the radioactivity in the liver was due to [3H]beta-alanine, about 40% to 3H2O, and about 5% to unidentified metabolites; [3H]uridine and [3H]uracil were not observed. A comparison of the rate of catabolism of [5-3H]-uridine, [5-3H]cytidine and [6-3H]thymidine showed that cytidine is degraded in the organism 25 times more slowly than uridine or thymidine. The biological half lives for the total degradation of the [3H]nucleosides to 3H2O, based on the values in the plasma, were: uridine 1.1 h; thymidine 1.3 h; cytidine 25 h. Furthermore, the turnover time of exogenous uridine in the plasma was found to be 9 min, which gives a half life of 6 min for the metabolism of exogenous uridine to uracil.  相似文献   

14.
Western blotting with anti-phosphotyrosine antibodies was employed in order to study insulin-dependent protein tyrosine phosphorylation in intact Fao cells. In insulin-treated cells, a prominent 180-kDa protein underwent tyrosine phosphorylation, which peaked at 45 s and then rapidly declined. Pretreatment of the cells with 1 mM Bt2cAMP or 0.16 microM 12-O-tetradecanoylphorbol-13-acetate inhibited the insulin-dependent phosphorylation of pp 180, while 1 mM vanadate or 3 mM H2O2 markedly potentiated it. These results indicate that phosphorylation of pp 180 is respectively regulated by agents that are known to synergize with or antagonize the action of the insulin receptor kinase. pp 180 is therefore likely to mediate physiological functions of this receptor kinase. Incubation of Fao cells with 3 mM H2O2 for 30 min prior to their treatment with insulin for 45 s allowed the detection of additional, previously undescribed, proteins pp 150, 114, 100, 85, 68, and 56 kDa that underwent insulin-dependent tyrosine phosphorylation. The potentiating effects of H2O2 were time- and dose-dependent and could be reversed by 2 mM dithiothreitol. Proteins phosphorylated in response to H2O2 plus insulin maintained their fully phosphorylated state for at least 20 min. We suggest that these phosphoproteins are potential physiological substrates for the insulin receptor kinase.  相似文献   

15.
In this study, the underlying mechanisms of stimulation by cyclocommunin, a natural pyranoflavonoid, of respiratory burst in rat neutrophils was investigated. Cyclocommunin evoked a concentration-dependent stimulation of superoxide anion (O2*-) generation with a slow onset and long lasting profile. The maximum response (16.4+/-2.3 nmol O2*-/10 min per 10(6) cells) was observed at 3-10 microM cyclocommunin. Cyclocommunin did not activate NADPH oxidase in a cell-free system. Cells pretreated with pertussis toxin or n-butanol did not affect the cyclocommunin-induced O2*- generation. However, a protein kinase inhibitor staurosporine and EGTA greatly reduced the O2*-generation caused by cyclocommunin. Treatment of neutrophils with phorbol 12-myristate 13-acetate (PMA), but not with formylmethionyl-leucyl-phenylalanine (fMLP), for 20 min significantly reduced the O2*- generation following the subsequent stimulation of cells with cyclocommunin. Cyclocommunin did not affect the cellular mass of phosphatidic acid (PA). Neither the tyrosine kinase inhibitor, genistein, nor the p38 mitogen-activated protein kinase (MAPK) inhibitor, SB203580, affected cyclocommunin-induced O2*- generation. The enzyme activities of neutrophil cytosolic and membrane-associated protein kinase C (PKC) were both increased significantly with 100 microM cyclocommunin. The membrane-associated PKC-theta and PKC-beta were increased following the stimulation of neutrophils with 30 and 100 microM cyclocommunin, respectively. Cyclocommunin reduced the [3H]phorbol 12,13-dibutyrate ([3H]PDB) binding to cytosolic PKC in a concentration-dependent manner. Cyclocommunin (> or =3 microM) significantly evoked a slow and long lasting [Ca2+]i elevation in neutrophils, and a phospholipase C (PLC) inhibitor U73122 greatly inhibited these Ca2+ responses. Moreover, the increase in cellular inositol bis- and trisphosphate (IP2 and IP3) levels were observed in neutrophils stimulated with 30 microM cyclocommunin for 3 min. Collectively, these results indicate that the stimulation of respiratory burst by cyclocommunin is probably mediated by the synergism of PKC activation and [Ca2+]i elevation in rat neutrophils.  相似文献   

16.
Early determinants of H2O2-induced endothelial dysfunction   总被引:4,自引:0,他引:4  
Reactive oxygen species (ROS) can stimulate nitric oxide (NO(*)) production from the endothelium by transient activation of endothelial nitric oxide synthase (eNOS). With continued or repeated exposure, NO(*) production is reduced, however. We investigated the early determinants of this decrease in NO(*) production. Following an initial H(2)O(2) exposure, endothelial cells responded by increasing NO(*) production measured electrochemically. NO(*) concentrations peaked by 10 min with a slow reduction over 30 min. The decrease in NO(*) at 30 min was associated with a 2.7-fold increase in O(2)(*-) production (p < 0.05) and a 14-fold reduction of the eNOS cofactor, tetrahydrobiopterin (BH(4), p < 0.05). Used as a probe for endothelial dysfunction, the integrated NO(*) production over 30 min upon repeated H(2)O(2) exposure was attenuated by 2.1-fold (p = 0.03). Endothelial dysfunction could be prevented by BH(4) cofactor supplementation, by scavenging O(2)(*-) or peroxynitrite (ONOO(-)), or by inhibiting the NADPH oxidase. Hydroxyl radical (()OH) scavenging did not have an effect. In summary, early H(2)O(2)-induced endothelial dysfunction was associated with a decreased BH(4) level and increased O(2)(*-) production. Dysfunction required O(2)(*-), ONOO(-), or a functional NADPH oxidase. Repeated activation of the NADPH oxidase by ROS may act as a feed forward system to promote endothelial dysfunction.  相似文献   

17.
We have investigated the effect of oxidants on ligand recognition and internalization by the macrophage mannose receptor. Rat bone marrow macrophages were treated with increasing concentrations of H2O2 for 30 min at 37 degrees C. Fifty percent inhibition of ligand uptake was observed at 250 microM, with only 10% of control uptake remaining following exposure to 1 mM H2O2 for 30 min. Electron micrographic analysis of macrophages following H2O2 treatment showed no morphological alterations compared to untreated cells. Ligand uptake was also inhibited by the following H2O2 generating systems: menadione, xanthine/xanthine oxidase, glucose/glucose oxidase, and phorbol 12-myristate 13-acetate-stimulated polymorphonuclear leukocytes. Inhibition could be blocked by catalase plus or minus superoxide dismutase. Treatment of macrophages at 4 degrees C with H2O2 had no effect on ligand binding, whereas treatment with H2O2 at 37 degrees C reduced binding to 15% of control levels and decreased the number of surface receptors to one-third of control cells. H2O2 treatment inhibited ligand degradation by macrophages, but did not prevent ligand movement from the surface to the interior of the cell. In addition, ligand delivery to lysosomes was blocked by oxidant treatment. These results suggest that treatment of macrophages with reagent H2O2 or H2O2-generating systems inhibits the normal ligand delivery and receptor recycling process involving the mannose receptor. Potential mechanisms might include receptor oxidation, alterations in ATP levels, or membrane lipid peroxidation.  相似文献   

18.
Previously it has been demonstrated that the human epidermis synthesises and degrades acetylcholine and expresses both muscarinic and nicotinic receptors. These cholinergic systems have been implicated in the development of the epidermal calcium gradient and differentiation in normal healthy skin. In vitiligo severe oxidative stress occurs in the epidermis of these patients with accumulation of H2O2 in the 10(-3)M range together with a decrease in catalase expression/activity due to deactivation of the enzyme active site. It was also shown that the entire recycling of the essential cofactor (6R)-l-erythro-5,6,7,8-tetrahydrobiopterin via pterin-4a-carbinolamine dehydratase (PCD) and dihydropteridine reductase (DHPR) is affected by H2O2 oxidation of Trp/Met residues in the enzyme structure leading to deactivation of these proteins. Using fluorescence immunohistochemistry we now show that epidermal H2O2 in vitiligo patients yields also almost absent epidermal acetylcholinesterase (AchE). A kinetic analysis using pure recombinant human AchE revealed that low concentrations of H2O2 (10(-6)M) activate this enzyme by increasing the Vmax>2-fold, meanwhile high concentrations of H2O2 (10(-3)M) inhibit the enzyme with a significant decrease in Vmax. This result was confirmed by fluorescence excitation spectroscopy following the Trp fluorescence at lambdamax 280nm. Molecular modelling based on the established 3D structure of human AchE supported that H2O2-mediated oxidation of Trp(432), Trp(435), and Met(436) moves and disorients the active site His(440) of the enzyme, leading to deactivation of the protein. To our knowledge these results identified for the first time H2O2 regulation of AchE. Moreover, it was shown that H2O2-mediated oxidation of AchE contributes significantly to the well-established oxidative stress in vitiligo.  相似文献   

19.
Hydrogen-peroxide-resistant Chinese hamster fibroblasts, derived from the HA-1 cell line, were isolated following continuous culturing in the presence of progressively increasing concentrations of hydrogen peroxide. The hydrogen-peroxide-resistant phenotype has been stable for over 360 days following removal from H2O2 stress. These H2O2-resistant cell lines demonstrate increased resistance to hyperthermic cell killing mediated by continuous heating at 43 degrees C but not 45 degrees C. The relationship between mammalian cellular adaptation to oxidative stress mediated by H2O2 and resistance to 43 degrees C hyperthermia is discussed.  相似文献   

20.
M W Washabaugh  W P Jencks 《Biochemistry》1988,27(14):5044-5053
Rate constants for C(2)-proton exchange from thiamin, N(1')-methylthiamin, and several 3-substituted-4-methylthiazolium ions catalyzed by D2O and deuterioxide ion were determined by 1H NMR at 30 degrees C and ionic strength 2.0 M. Values of pKa for the thiazolium ions, including thiamin itself, were found to be in the range pKa = 17-19; the pKa values for N(1')-protonated thiamin and free thiamin C(2)-H in H2O are 17.7 and 18.0, respectively. The pKa value for N(1')-protonated thiamin was calculated from the observed rate constant for the pD-independent reaction with D2O after correction for a secondary solvent deuterium isotope effect of kH2O/kD2O = 2.6. The pKa value for free thiamin was calculated from the rate constant for catalysis by OD- after correction by a factor of 3.3 = 8/2.4 for an 8-fold negative deviation of kOD from the Br?nsted plot of slope 1.0 for general base catalysis and a secondary solvent isotope effect of kOD/kOH = 2.4. Values of k-a = 2 X 10(10) and 3 X 10(9) M-1 s-1 were assumed for diffusion-controlled protonation of the C(2) ylide in the reverse direction by H3O+ and H2O, respectively. The Hammett rho I value for the exchange reaction catalyzed by deuterioxide ion or D2O is 8.4 +/- 0.2. There is no positive deviation of the rate constants for free or N(1')-substituted thiamin analogues in either Hammett correlation. This shows that the aminopyrimidinyl group does not provide significant intramolecular catalysis of nonenzymic C(2)-proton removal in the coenzyme.  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