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 active outward translocation of phospholipid analogues from the inner to the outer membrane leaflet of human erythrocytes by the multi-drug resistance protein MRP1 (ABCC1) depends on intracellular reduced glutathione (GSH). Entrapment of ATP and increasing amounts of GSH inside resealed ghosts prepared from erythrocytes resulted in an up to six-fold increase of the translocation rate. Entrapped oxidized glutathione (GSSG) acted inhibitory but produced stimulation after addition of the disulphide-reducing reagent dithioerythritol. Modification of GSH by esterification of the C-terminal carboxylate of Gly, removal of the N-terminal Glu or substitution of the SH group by an anionic S-dicarboxyethyl or sulphonate group abolished stimulation. The effect of S-alkylation of GSH depended on the length of the alkyl group. S-methyl GSH was somewhat more effective than GSH, but maximal stimulation was similar. S-butyl GSH acted poorly stimulatory while S-hexyl GSH was essentially ineffective. Analyses of the kinetic data of translocation revealed Km values for GSH and methyl-GSH of respectively 7.4±2.4 and 4.9±1.1 mmol l?1. At high GSH levels and defined constant ATP levels using an ATP-regenerating system, the Km for ATP of the outward translocation was 0.16±0.02 mmol l?1. In the same system lacking GSH, the Km for ATP of the inward translocation by the aminophospholipid flippase was 0.53±0.23 mmol l?1.  相似文献   

2.
3.
Human erythrocytes are essentially impermeable to glutamate and yet there is a continual requirement for the amino acid for glutathione synthesis. In addition, the intracellular glutamate concentration is approximately five times that of plasma. We present evidence that glutamate enters the red cell as small peptides which are rapidly hydrolysed by cytoplasmic peptidase(s) and that with the estimated physiological levels of plasma glutamyl-peptides the rate of inward flux would be adequate to maintain the glutamate pool at its observed level. Experimentally, we used 1H spin-echo n.m.r. spectroscopy to follow peptide hydrolysis, since peptide spectra are different from those of the free amino acids and the spin-echo sequence enables the monitoring of reactions in concentrated lysates and whole cell suspensions. Thus, the system was studied under near-physiological conditions. Weighted non-linear regression analysis of progress curves using the integrated Michaelis-Menten equation was used to obtain estimates of Km and Vmax. for the hydrolysis of alpha-L-glutamyl-L-alanine and L-alanyl-alpha-L-glutamate in lysates and whole cell suspensions; the values for lysates were Km = 3.60 +/- 0.29 and 5.4 +/- 0.4 mmol/l and Vmax. = 120 +/- 4 and 46.7 +/- 1.7 mmol/h per 1 of packed cells respectively. In whole cell suspensions the rate of peptide hydrolysis was much slower and dominated by the transmembrane flux-rate. The estimates of the steady-state kinetic parameters for the transport were Kt = 2.35 +/- 0.41 and 11.2 +/- 1.0 mmol/l and Vmax. = 3.26 +/- 0.13 and 19.7 +/- 0.7 mmol/h per 1 of packed cells respectively for the previously mentioned peptides. Using the n.m.r. procedure we failed to detect any glutaminase activity in whole cells or lysates; thus, we exclude the possibility that glutamate gains entry to the cell as glutamine which is subsequently hydrolysed by glutaminase.  相似文献   

4.
The effect of enalapril and captopril on total glutathione content (GSSG + GSH) and selenium-dependent glutathione peroxidase (Se-GPx) and glutathione reductase (GSSG-Rd) activities was investigated in mouse tissues. CF-1 mice (4-mo-old females) received water containing enalapril (20 mg/l) or captopril (50 mg/l) for 11 wk. Enalapril increased GSSG + GSH content (P < 0.05) in erythrocytes (147%), brain (112%), and lung (67%), and captopril increased GSSG + GSH content in erythrocytes (190%) and brain (132%). Enalapril enhanced Se-GPx activity in kidney cortex (42%) and kidney medulla (23%) and captopril in kidney cortex (30%). GSSG-Rd activity was enhanced by enalapril in erythrocytes (21%), brain (21%), liver (18%), and kidney cortex (53%) and by captopril in erythrocytes (25%), brain (19%), and liver (34%). In vitro erythrocyte oxidant stress was evaluated by thiobarbituric acid-reactive substances (TBARS) production (control 365 +/- 11, enalapril 221 +/- 26, captopril 206 +/- 17 nmol TBARS x g Hb(-1) x h(-1); both P < 0.05 vs. control) and phenylhydrazine-induced methemoglobin (MetHb) formation (control 66.5 +/- 3.5, enalapril 52.9 +/- 0.4, captopril: 56.4 +/- 2.9 micromol MetHb/g Hb; both P < 0.05 vs. control). Both angiotensin-converting enzyme inhibitor treatments were associated with increased nitric oxide production, as assessed by plasma NO-(3) + NO-(2) level determination (control 9.22 +/- 0.64, enalapril 13.7 +/- 1.9, captopril 17.3 +/- 3.0 micromol NO-(3) + NO-(2)/l plasma; both P < 0.05 vs. control). These findings support our previous reports on the enalapril- and captopril-induced enhancement of endogenous antioxidant defenses and include new data on glutathione-dependent defenses, thus furthering current knowledge on the association of ACE inhibition and antioxidants.  相似文献   

5.
The assimilation of tri- and tetrapeptides by human erythrocytes   总被引:1,自引:0,他引:1  
Evidence is presented that tripeptides enter human erythrocytes via saturable transport system(s) at rates similar to those previously described for dipeptides (King, G.F. and Kuchel, P.W. (1985) Biochem. J. 227, 833-842) but that the transmembrane flux rates for tetrapeptides are considerably less. 1H spin-echo NMR spectroscopy was used to monitor the coupled uptake and hydrolysis of peptides by red cells, since it enabled the simultaneous measurement of the levels of substrates and products of peptidase-catalysed reactions in suspensions with haematocrits similar to those found in vivo. Weighted non-linear least-squares regression of the integrated Michaelis-Menten equation onto progress curves obtained from the hydrolysis of Tyr-Gly-Gly and Gly-Gly-Gly in RBC lysates gave Km = 2.11 +/- 0.08 and 23.4 +/- 0.9 mmol/l and Vmax = 307 +/- 3 and 905 +/- 22 mmol/h per 1 packed cells, respectively. In whole cell suspensions, the rate of hydrolysis was considerably less and was dominated by the transmembrane flux of tripeptide. Progress curve analysis thus yielded the steady-state kinetic parameters for peptide transport; the values were Km = 11.6 +/- 1.1 and 56 +/- 18 mmol/l and Vmax = 12.9 +/- 3.0 and 36.4 +/- 3.2 mmol/h per 1 packed cells, respectively, for the previously mentioned peptides. The rate of transport of the tetrapeptide Gly-Gly-Gly-Gly was considerably less than either of the tripeptides. The above mentioned steady-state kinetic parameters were used in computer simulations of the coupled uptake and hydrolysis of tripeptides by human erythrocytes under physiological conditions; these simulations revealed certain similarities between the rates of peptide uptake by erythrocytes and the intestine in vivo.  相似文献   

6.
Nitric oxide (NO) in many cells inactivates aconitase and mitochondrial respiratory chain, and influenced glyceraldehyde 3-phosphate dehydrogenase activity. The aim of this study was to evaluate role of nitroglycerin (NTG), a widely used NO donor, on energy metabolism of rat reticulocytes. Rat reticulocyte rich red blood cell suspensions containing 70-100% of reticulocytes, were aerobically incubated without (control) or in the presence of different concentrations of (a) NTG (0.1, 0.25, 0.5, 1.0, 1.5 mmol/l), (b) 8-Br-cGMP (0.1, 0.5, 1.0 mmol/l) and (c) NaNO2 and NaNO3 (1 mmol/l). NTG in dose- and time-dependent manner decreased total (p>0.05; EC50 = 0.78+/-0.05 mmol/l) and coupled (p<0.05; EC50 = 0.50+/-0.04 mmol/l) and increased uncoupled oxygen consumption (p<0.05: EC50 = 0.36+/-0.01 mmol/l). They were accompanied by stimulation of glycolysis, as measured by increased glucose consumption and lactate accumulation (p<0.001 EC50 = 0.53 and 0.53 mmol/l, respectively). Levels of all glycolytic intermediates in the presence of NTG indicate stimulation of HK-PFK, GA3PDH and PK activity. NTG significantly decreased ATP level, which accompanied by increased ADP and AMP levels. However, level of total adenine nucleotides (TAN) was significantly lower, which was consequence of increased catabolism of adenine nucleotides (increased hypoxanthine level; p<0.05). Stimulation of glycolysis accompanied with inhibition of the OxP, activation of HK-PFK, decrease of ATP and simultaneous rise of ADP and AMP levels, all together represent an example of Pasteur effect occurring in NTG-treated reticulocytes. In rat reticulocytes under steady state conditions 93% of overall energy was produced by OxP, but only 7% by glycolysis. Due to decrease of coupled oxygen consumption in the presence of NTG, ATP production via OxP was significantly diminished. Simultaneous increase of glycolytic ATP production is not enough to provide constant either ATP production or concentration. Calculated mean ATP-turnover time was prolonged even for 45% in the presence of 1.5 mmol/l NTG. Metabolic effects of NTG were not mimic by exogenous 8-Br-cGMP, NaNO2 or NaNO3, which indicate that NTG induced a) inhibition of coupled respiration and b) stimulation of glycolysis in rat reticulocytes are mediated by NO as an effector molecule.  相似文献   

7.
Erythrocytes are both an important source and target of reactive oxygen species in sickle cell disease. Levels of glutathione, a major antioxidant, have been shown to be decreased in sickle erythrocytes and the mechanism leading to this deficiency is not known yet. Detoxification of reactive oxygen species involves the oxidation of reduced glutathione (GSH) into glutathione-disulfide (GSSG) which is actively transported out of erythrocyte. We questioned whether under oxidative conditions, GSSG efflux is increased in sickle erythrocytes. Erythrocytes of 18 homozygous sickle cell patients and 9 race-matched healthy controls were treated with 2,3-dimethoxy-l,4-naphthoquinone, which induces intracellular reactive oxygen species generation, to stimulate GSSG production. Intra- and extracellular concentrations of GSH and GSSG were measured at baseline and during 210-minute 2,3-dimethoxy-l,4-naphthoquinone stimulation. While comparable at baseline, intracellular and extracellular GSSG concentrations were significantly higher in sickle erythrocytes than in healthy erythrocyte after 210-minute 2,3-dimethoxy-l,4-naphthoquinone stimulation (69.9 ± 3.7 μmol/l vs. 40.6 ± 6.9 μmol/l and 25.8 ± 2.7 μmol/l vs. 13.6 ± 1.7 μmol/l respectively, P<0.002). In contrast to control erythrocytes, where GSH concentrations remained unchanged (176 ± 8.4 μmol/l vs. 163 ± 13.6 μmol/l, NS), GSH in sickle erythrocytes decreased significantly (from 167 ± 8.8 μmol/l to 111 ± 11.8 μmol/l, P<0.01) after 210-minute 2,3-dimethoxy-l,4-naphthoquinone stimulation. Adding multidrug resistance-associated protein-1 inhibitor (MK571) to erythrocytes blocked GSSG efflux in both sickle and normal erythrocytes. GSSG efflux, mediated by multidrug resistance-associated protein-1, is increased in sickle erythrocytes, resulting in net loss of intracellular glutathione and possibly higher susceptibility to oxidative stress.  相似文献   

8.
Glutathione transferases exhibit both isomerase and transferase activity. The acceptance of steroids as substrates for or inhibitors of these activities was studied using a 350-fold enriched preparation of the enzyme from human placenta. As an isomerase, the enzyme preparation catalyzed the conversion of pregn-5-ene-3,20-dione (Km 0.03 mmol/l) and androst-5-ene-3,17-dione (Km 0.05 mmol/l) to the respective 4-ene-3-oxosteroids (specific activity 0.8 U/mg protein). This isomerase activity strictly depended on the presence of glutathione (Km 0.04 mmol/l). As a transferase, the enzyme preparation catalyzed the conjugation of glutathione (Km 0.5 mmol/l) with 1-chloro-2,4-dinitrobenzene (Km 1.0 mmol/l) (specific activity 100 U/mg protein). This transferase activity was inhibited by all phenolic (KI values 0.2-1.5 mmol/l) and some of the neutral steroids (KI values 1.4-3.5 mmol/l) tested. Phenolic steroids inhibited the enzyme activity competitively to 1-chloro-2,4-dinitrobenzene and non-competitively to both substrates. The results indicate that steroids can interact with the placental glutathione transferase in vitro both as substrates and as inhibitors. Since, however, the observed Km and KI values of the steroids are far above the values of their concentrations in the placenta, these interactions are of only minor physiological relevance.  相似文献   

9.
1. Urea synthesis was studied in isolated perfused rat liver during cell volume regulatory ion fluxes following exposure of the liver to anisotonic perfusion media. Lowering of the osmolarity in influent perfusate from 305 mOsm/l to 225 mOsm/l (by decreasing influent [NaCl] by 40 mmol/l) led to an inhibition of urea synthesis from NH4Cl (0.5 mmol/l) by about 60% and a decrease of hepatic oxygen uptake by 0.43 +/- 0.03 mumol g-1 min-1 [from 3.09 +/- 0.13 mumol g-1 min-1 to 2.66 +/- 0.12 mumol g-1 min-1 (n = 9)]. The effects on urea synthesis and oxygen uptake were observed throughout hypotonic exposure (225 mOsm/l). They persisted although volume regulatory K+ efflux from the liver was complete within 8 min and were fully reversible upon reexposure to normotonic perfusion media (305 mOsm/l). A 42% inhibition of urea synthesis from NH4Cl (0.5 mmol/l) during hypotonicity was also observed when the perfusion medium was supplemented with glucose (5 mmol/l). Urea synthesis was inhibited by only 10-20% in livers from fed rats, and was even stimulated in those from starved rats when an amino acid mixture (twice the physiological concentration) plus NH4Cl (0.2 mmol/l) was infused. 2. The inhibition of urea synthesis from NH4Cl (0.5 mmol/l) during hypotonicity was accompanied by a threefold increase of citrulline tissue levels, a 50-70% decrease of the tissue contents of glutamate, aspartate, citrate and malate, whereas 2-oxoglutarate, ATP and ornithine tissue levels, and the [3H]inulin extracellular space remained almost unaltered. Further, hypotonic exposure stimulated hepatic glutathione (GSH) release with a time course roughly paralleling volume regulatory K+ efflux. NH4Cl stimulated lactate release from the liver during hypotonic but not during normotonic perfusion. In the absence of NH4Cl, hypotonicity did not significantly affect the lactate/pyruvate ratio in effluent perfusate. With NH4Cl (0.5 mmol/l) present, the lactate/pyruvate ratio increased from 4.3 to 8.2 in hypotonicity, whereas simultaneously the 3-hydroxybutyrate/acetoacetate ratio slightly, but significantly decreased. 3. Addition of lactate (2.1 mmol/l) and pyruvate (0.3 mmol/l) to influent perfusate did not affect urea synthesis in normotonic perfusions, but completely prevented the inhibition of urea synthesis from NH4Cl (0.5 mmol/l) induced by hypotonicity. Restoration of urea production in hypotonic perfusions by addition of lactate and pyruvate was largely abolished in the presence of 2-cyanocinnamate (0.5 mmol/l). Addition of 3-hydroxybutyrate (0.5 mmol/l), but not of acetoacetate (0.5 mmol/l) largely reversed the hypotonicity-induced inhibition of urea synthesis from NH4Cl.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

10.
A direct method for determination of Fe, Cu, Zn, Mg and Se in erythrocytes was developed. The aim of the present study was to establish a method for examining perioperative levels of the above mentioned elements simultaneously in erythrocytes and plasma by atomic absorption spectrophotometry in 11 patients undergoing neurosurgery for acute spinal nerve compressions because of intervertebral disk prolapses. Reference values for erythrocytes were 11.49 +/- 3.48 mmol/mmol Hb; 0.82 +/- 0.087 mmol/mmol Hb; 9.01 +/- 2.20 mmol/mmol Hb; 0.104 +/- 0.032 mmol/mmol Hb; 0.07 +/- 0.050 mmol/mmol Hb for iron, copper, zinc, magnesium, and selenium, respectively. Postoperative erythrocyte concentrations did not differ significantly compared to those obtained preoperatively and remained within the reference ranges perioperatively. For plasma the following reference values were used: 19.0 +/- 8.0 mmol/l (Fe); 20.1 +/- 8.2 mmol/l (Cu); 15.4 +/- 4.6 mmol/l (Zn); 0.9 +/- 0.15 mmol/l (Mg); 1.02 +/- 0.3 mmol/l (Se). There was a significant decrease in the concentration of copper in plasma (13.41 +/- 3.46 mmol/l, p < 0.1) and zinc (10.73 +/- 2.73 mmol/l, p < 0.1) immediately postoperative, iron (10.56 +/- 3.91 mmol/l, p < 0.1) and zinc on day 1 (11.28 +/- 1.88 mmol/l, p < 0.10), and a significant postoperative increase of copper on day 5 (18.81 +/- 3.97 mmol/l, p < 0.1), postoperatively. The mean plasma concentrations of iron, copper, zinc magnesium and selenium remained within the reference ranges during the entire period.  相似文献   

11.
PURPOSE: To investigate whether caffeic acid phenethyl ester (CAPE) and cortisone prevent proliferative vitreoretinopathy (PVR). METHODS: Twenty pigmented rabbits were used in this study. All rabbits except controls received an intravitreal injection of 0.15 ml (75,000 U) of platelet-rich plasma into their left eye. The animals were divided into four groups: group I was treated with intraperitoneal injection of 0.5 ml (15 micromol/kg) of CAPE for 3 days, group II received 0.15 ml (4 mg/kg) of intravitreal cortisone, group III received nothing (blank group), and group IV (control group) received only 1 ml of 1% ethanol intraperitoneally daily for 3 days. Proliferative changes were graded in a masked fashion by indirect ophthalmoscopy for a 15-day follow-up period. The malondialdehyde (MDA), reduced glutathione (GSH) and total nitrite (NO) levels were measured in the vitreous humor. RESULTS: The grades of PVR were B-C in group I, and C-D in group II. The PVR grade in the control group was C-D. The mean MDA level in group I (4.0+/-0.8 micromol/l) was significantly lower than in the blank group (6.0 micromol/l) (p < 0.05). The mean GSH level in group I (71.0+/-11.2 micromol/l) was significantly different than in the blank group (p < 0.05). The MDA and GSH levels in group II were 4.7+/-0.6 micromol/l and 53.8+/-7.8 micromol/l, respectively. Both these levels were not significantly different from the blank group (p > 0.05). The NO levels in both treatment groups were significantly lower than in the blank group (p < 0.001). CONCLUSION: These findings suggest an inhibitory effect of CAPE on PVR. The inhibitory effect was supported by lower MDA and NO with higher GSH levels in treatment groups than in the blank group. There was no detected significant effect of cortisone for preventing PVR experimentally.  相似文献   

12.
The objective of this study was to investigate if erythrocytes play a role in the maintenance of redox homeostasis of the plasma. Thus, we studied L-cysteine efflux and influx in vitro in human erythrocytes. In the present study, we exposed the erythrocytes to different concentrations of L-cysteine and then measured the intracellular free -SH concentrations. Erythrocytes treated in the same manner were later utilized for the cysteine efflux studies. The effect of temperature on the influx and the efflux processes were also evaluated. Change in the free -SH content of the buffer was evaluated as a measure for the presence of an efflux process. The effects of free -SH depletion on L-cysteine transport is also investigated. We also determined the rate of L-cysteine efflux in the presence and absence of buthionine sulfoximine (BSO) in erythrocytes that are pretreated with 1-chloro-2,4-dinitro benzene, a glutathione (GSH) depletory. Our L-cysteine influx studies demonstrated that erythrocytes can respond to increases in L-cysteine concentration in the extracellular media and influx L-cysteine in a concentration-dependent manner. Free -SH concentrations in erythrocytes treated with 1 mM L-cysteine reached to 1.64 +/- 0.06 mM in 1 h whereas this concentration reached to 4.30 +/- 0.01 mM in 10 mM L-cysteine treated erythrocytes. The L-cysteine efflux is also determined to be time-and concentration-dependent. Erythrocytes that are pretreated with higher L-cysteine concentrations displayed a higher efflux process. Outside concentration of free -SH in 1 mM L-cysteine pretreated erythrocytes reached to 0.200 +/- 0.005 mM in 1 h whereas this concentration reached to 1.014 +/- 0.002 with 10 mM L-cysteine pretreated erythrocytes. Our results also indicate that the rate of inward and outward transport of L-cysteine is affected by the oxidative status of the erythrocytes. When GSH is depleted and GSH synthesis is blocked, the L-cysteine uptake and the efflux processes are significantly decreased. Depending on our results, it could be concluded that erythrocytes play a role in the regulation of the plasma redox status and intracellular level of GSH determines the rate of the L-cysteine efflux.  相似文献   

13.
The effect of a transmembrane pH gradient on the ouabain, bumetanide, and phloretin resistant H+ efflux was studied in rabbit erythrocytes. Proton equilibration was reduced by the use of DIDS (125 microM) and acetazolamide (1 mM). H+ efflux from acid loaded erythrocytes (pHi = 6.1) was measured in a K+ (145 mM) medium, pH0 = 8.0, in the presence and absence of 60 microM 5,N,N-dimethyl-amiloride (DMA). The H+ efflux rate in a K+-containing medium was 116.38 +/- 4.5 mmol/l cell X hr. Substitution of Nao+ for Ko+ strongly stimulated H+ efflux to 177.89 +/- 7.9 mmol/l cell X hr. The transtimulation of H+ efflux by Nao+ was completely abolished by DMA falling to values not different from controls with an ID50 of about 8.6 X 10(-7) M. The sequence of substrate selectivities for the external transport site were Na greater than greater than greater than Li greater than choline, Cs, K, and Glucamine. The transport system has no specific anion requirement, but is inhibited by NO3-. The DMA sensitive H+ efflux was a saturable function of [Na+]o, with an apparent Km and Vmax of about 14.75 +/- 1.99 mM and 85.37 +/- 7.68 mmol/l cell X hr, respectively. However, the Nao+-dependent and DMA-sensitive H+ efflux was sigmoidally activated by [H+]i, suggesting that Hi+ interacts at both transport and modifier sites. An outwardly directed H+ gradient (pHi 6.1, pH = 8.0) also promoted DMA sensitive Na+ entry (61.2 +/- 3.0 mmol/l cell X hr) which was abolished when pHo was reduced to 6.0. The data is therefore consistent with the presence of a Na+/H+ exchange system in rabbit erythrocytes.  相似文献   

14.
Hepatic glutathione (GSH) plays an important role in the detoxification of reactive molecular intermediates. Because of evidence that the intrahepatic turnover of glutathione in the rat may be largely accounted for by efflux from hepatocytes into the general circulation, the quantitation of plasma GSH turnover in vivo could provide a noninvasive index of hepatic glutathione metabolism. We developed a method to estimate plasma glutathione turnover and clearance in the intact, anesthetized rat using a 30-min unprimed, continuous infusion of 35S-labelled GSH. A steady state of free plasma glutathione specific radioactivity was achieved within 10 min, as determined by high-pressure liquid chromatography with fluorometric detection after precolumn derivatization of the plasma samples with monobromobimane. The method was tested after two treatments known to alter hepatic GSH metabolism: 90 min after intraperitoneal injection of 4 mmol/kg buthionine sulfoximine (BSO), an inhibitor of glutathione synthesis, and after a 48-h fast. Liver glutathione concentration (mean +/- SEM) was 5.00 +/- 0.53 mumol/g wet weight in control rats. It decreased to 3.10 +/- 0.35 mumol/g wet weight after BSO injection and to 3.36 +/- 0.14 mumol/g wet weight after fasting (both p less than 0.05). Plasma glutathione turnover was 63.0 +/- 7.46 nmol.min-1.100 g-1 body weight in control rats, 35.0 +/- 2.92 nmol.min-1.g-1 body weight in BSO-treated rats, and 41.7 +/- 2.28 nmol.min-1.g-1 body weight after fasting (both p less than 0.05), thus reflecting the hepatic alterations. This approach might prove useful in the noninvasive assessment of liver glutathione status.  相似文献   

15.
The oxidative action of 1 mmol l(-1) phenylhydrazine hydrochloride (PH) was studied on human erythrocytes treated with the antioxidants vitamin C (vit. C) and vitamin E (vit. E). The erythrocytes were resuspended in PBS to obtain 35% cell packed volume, and then submitted to the oxidative action of PH for 20 min, with or without previous incubation for 60 min with vit. C or vit. E. Heinz bodies and methemoglobin formation by PH were inhibited in the presence of vit. C. At the concentration of 90 mmol l(-1), vit. C, not only seemed to lose its antioxidant effect, but it also promoted an increase in methemoglobin formation. Vit. C (0.5-80 mmol l(-1)) did not protect against GSH depletion by PH. Vit. C alone produced insignificant hemolysis, but, in the presence of PH, the hemolysis indices were more accentuated. Heinz body formation by PH was inhibited in the presence of vit. E. Formation of methemoglobin induced by PH was decreased by vit. E (0.1-2 mmol l(-1)), although vit. E (3-80 mmol l(-1)) did not lower the concentration of methemoglobin and did not lead to the recovery of the GSH depleted by PH. The results obtained suggest that vit. C and vit. E contribute to the decrease in oxidative stress caused by PH.  相似文献   

16.
It was previously shown that 1,25-dihydroxyvitamin D(3) (1, 25(OH)(2)D(3)) enhances the cytotoxic activity of tumor necrosis factor alpha (TNFalpha), doxorubicin and menadione. A feature shared by these anticancer agents is the involvement of reactive oxygen species (ROS) in their action. In this work we found that 1, 25(OH)(2)D(3) acted synergistically with interleukin 1 beta (IL-1beta) or interleukin 6 (IL-6) to inhibit the proliferation of MCF-7 breast cancer cells. The extent of the synergism was maximal at 1 nM, a concentration at which 1,25(OH)(2)D(3), acting singly, only marginally reduced the cell number. The thiol antioxidant, N-acetylcysteine (NAC) abolished the synergism between IL-1beta or IL-6 and 1,25(OH)(2)D(3), but had only a small protective effect when the cytokines acted alone. NAC and reduced glutathione (GSH) protected MCF-7 cells from cytotoxicity induced both by TNFalpha alone and by TNFalpha and 1,25(OH)(2)D(3). A two-day exposure to TNFalpha caused a 27.7+/-3.1% (mean +/- SEM) reduction in GSH content. This effect increased to 46.4+/-5.5% by co-treatment with 1, 25(OH)(2)D(3) which did not affect GSH levels on it own. We conclude that 1,25(OH)(2)D(3) can act synergistically with anticancer cytokines present in the tumor milieu and that ROS plays a mediatory role in this interaction.  相似文献   

17.
1. The (Na+/K+)-ATPase activity of brown fat membranes is increased by norepinephrine, the physiological mediator of thermogenesis in this tissue. 2. This increased ATPase activity was inhibited approximately 50% by either propranolol (a beta-adrenergic blocker) or phentolamine (an alpha-blocker). 3. The alpha-agonist, phenylephrine and the beta-agonist, isoproterenol, also stimulated the ATPase activity. 4. That these latter effects were receptor-specific is supported by the finding that: (a) l(-)isoproterenol stimulation was inhibited by propranolol but not by phentolamine; (b) d(+)isoproterenol had no stimulatory effect on the ATPase activity; and (c) the l(-)phenylephrine-induced increase was inhibited by phentolamine but not by propranolol. 5. (-)norepinephrine, l(-)isoproterenol and l(-)phenylephrine all decreased the apparent Km for K+ of the (Na+/K+)-ATPase but did not alter the apparent Km for ATP or the Vmax of the reaction.  相似文献   

18.
The intracellular "Na+ activity" was measured in erythrocytes of normotensive subjects (46), in essential hypertensive patients (18), in their children (20) and in patients with secondary hypertension (8). In normotensive subjects without a genetic trait of hypertension intracellular "Na+ activity" was 7.3 +/- 0.8 mmol/l, in secondary hypertensive patients was 7.5 +/- 0.6 mmol/l, in essential hypertensive patients was 10.9 +/- 1.1 mmol/l and in their children was 8.6 +/- 2.1 mmol/l. In this group (children) it was possible to differentiate between 2 population, the 1 degree with height intracellular "Na+ activity" (8); the 2 degrees with normal intracellular "Na+ activity".  相似文献   

19.
生物合成谷胱甘肽种间耦合ATP再生系统的构建   总被引:6,自引:0,他引:6  
利用重组大肠杆菌Ⅱ 1中的谷胱甘肽合成酶系和面包酵母WSH J7中的ATP生物合成酶系 ,构建了一个以葡萄糖为能源的种间耦合ATP再生系统。经过通透性处理的酵母细胞几乎不能消耗葡萄糖。在反应体系中添加 1mmol/LAMP和 0 0 5mmol/LNADH ,即可启动酵母的酵解途径。提高耦合系统中的葡萄糖浓度 ,可促进GSH的合成。当葡萄糖浓度为 40 0mmol/L时 ,系统内GSH浓度达到 1 0 4mmol/L(3 2 g/L)。Mg2 +缺乏时 ,耦合系统和外加ATP的非耦合系统均不能合成谷胱甘肽。耦合系统中Mg2 +与ATP形成螯合物 ,可能是导致耦合系统中GSH产量较低的原因。在耦合系统中补加Mg2 +,反应 6h时GSH浓度达到 1 4 3mmol/L(4 4g/L)。  相似文献   

20.
The molecular basis of the toxic properties of phenoxy herbicides in humans and animals has been insufficiently studied. In this study, damage parameters [levels of reduced glutathione (GSH) and total glutathione; activity of glutathione reductase (GR); activities of catalase (CAT) and superoxide dismutase (SOD); levels of adenine nucleotides and adenine energy charge (AEC)] were measured in human erythrocytes exposed in vitro to 2,4,5-trichlorophenoxyacetic acid (2,4,5-T) and its metabolite 2,4,5-trichlorophenol (2,4,5-TCP). Both 2,4,5-T and 2,4,5-TCP decreased the level of reduced glutathione (GSH) in erythrocytes in comparison to the control, but did not significantly change the total glutathione (2GSH + GSSG). This suggests that GSH concentration decreases concomitantly with an increase in oxidized glutathione (GSSG). 2,4,5-TCP at 100 ppm significantly decreased catalase and SOD activities. 2,4,5-T and 2,4,5-TCP did not significantly change the activity of glutathione reductase. 2,4,5-TCP decreased the level of ATP and increased the content of ADP and AMP, indicating a fall in AEC. 2,4,5-T and 2,4,5-TCP significantly changed the erythrocyte morphology. All these data are evidence of oxidative stress in erythrocytes incubated with 2,4,5-T and 2,4,5-TCP; the stress appears to be more intense in the case of 2,4,5-TCP.  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