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 93 毫秒
1.
Monochlorobimane (MCB) is often used to visualize glutathione (GSH) levels in cultured cells, since it is quickly converted to a fluorescent GSH conjugate (GS–MCB). To test for consequences of MCB application on the GSH metabolism of astrocytes, we have studied rat astrocyte-rich primary cultures as model system. MCB caused a concentration dependent rapid decrease in the cellular GSH content. Simultaneously, a transient accumulation of GS–MCB in the cells was observed with a maximal content 5 min after MCB application. The cellular accumulation was followed by a rapid release of GS–MCB into the medium with a maximal initial export rate of 27.9 ± 6.5 nmol h−1 mg protein−1. Transporters of the family of multidrug resistance proteins (Mrps) are likely to be involved in this export, since the Mrp inhibitor MK571 lowered the export rate by 60%. These data demonstrate that, due to its rapid export from astrocytes, GS–MCB is only under well-defined conditions a reliable indicator of the cellular GSH concentration and that MK571 can be used to maintain maximal GS–MCB levels in astrocytes.  相似文献   

2.
The mechanism of cytotoxicity on liver carcinoma Bel-7402 cells induced by copper-1,10-phenanthroline, Cu(OP)2, has been studied. Cell viability and apoptotic rate were examined in cells treated with Cu(OP)2 or Cu2+ alone. It was found that the apoptosis induced by Cu(OP)2 could not be induced by Cu2+ or OP alone in our experimental conditions. Total copper content in cells was measured by atomic absorption spectrophotometry, and the abnormal elevation of intracellular copper transported by lipophilic OP ligand may play the role of initial factor in the apoptosis, which caused subsequent redox state changes in cells. Intracellular levels of reactive oxygen species (ROS) were detected by fluorescent probe 2′,7′-dichlorofluorescein diacetate (DCFH-DA). Reduced (GSH) and total glutathione (GSSG + GSH) were determined by High-performance liquid chromatography (HPLC) after derivatization, and the ratios of GSH/GSSG were subsequently calculated. The overproduction of ROS and the decreased GSH/GSSG ratio were observed in cells which represented the occurrence of oxidative stress in the apoptosis. Oxidative DNA damage was also found in cells treated with Cu(OP)2 in the early stage of the apoptosis, and it suggests that the activation of DNA repair system may be involved in the pathway of the apoptosis induced by Cu(OP)2.  相似文献   

3.
The ontogeny of two gonadotropin-releasing-hormone (GnRH) systems, salmon GnRH (sGnRH) and chicken GnRH-II (cGnRH-II), was investigated in zebrafish (Danio rerio). In situ hybridization (ISH) first detected sGnRH mRNA-expressing cells at 1 day post-fertilization (pf) anterior to the developing olfactory organs. Subsequently, cells were seen along the ventral olfactory organs and the olfactory bulbs, reaching the terminal nerve (TN) ganglion at 5–6 days pf. Some cells were detected passing posteriorly through the ventral telencephalon (10–25 days pf), and by 25–30 days pf, sGnRH cells were found in the hypothalamic/preoptic area. Continuous documentation in live zebrafish was achieved by a promoter-reporter expression system. The expression of enhanced green fluorescent protein (EGFP) driven by the sGnRH promoter allowed the earlier detection of cells and projections and the migration of sGnRH neurons. This expression system revealed that long leading processes, presumably axons, preceded the migration of the sGnRH neuron somata. cGnRH-II mRNA expressing cells were initially detected (1 day pf) by ISH analysis at lateral aspects of the midbrain and later on (starting at 5 days pf) at the midline of the midbrain tegmentum. Detection of red fluorescent protein (DsRed) driven by the cGnRH-II promoter confirmed the midbrain expression domain and identified specific hindbrain and forebrain cGnRH-II-cells that were not identified by ISH. The forebrain DsRed-expressing cells seemed to emerge from the same site as the sGnRH-EGFP-expressing cells, as revealed by co-injection of both constructs. These studies indicate that zebrafish TN and hypothalamic sGnRH cell populations share a common embryonic origin and migratory path, and that midbrain cGnRH-II cells originate within the midbrain. This study was supported by the US-Israel Bi-national Agricultural Research and Development (BARD) Foundation (grant 3428-03).  相似文献   

4.
The green alga Pseudokirchneriella subcapitata has been extensively used for the assessment of adverse impacts of pollutants. Glutathione is involved in antioxidant defence and drug detoxification. Intracellular reduced glutathione (GSH) concentration can be used as an indicator of the health of cells. This work describes a simple and fast fluorescent cell-based assay for the evaluation of intracellular GSH in the alga P. subcapitata, using monochlorobimane (mBCl). Metabolically active algal cells incubated with 50?μmol?L?1 mBCl form fluorescent bimane–glutathione (B-SG) adducts that can be measured fluorometrically. The distribution of GSH (B-SG adducts) in whole cells can be observed by epifluorescence microscopy, in the form of blue fluorescent spots. Depletion of cellular GSH with iodoacetamide, inhibition of glutathione S-transferase with ethacrynic acid or heat-induced death of the cells inhibited the formation of fluorescent adducts in the presence of mBCl. The fluorometric assay, using the 96-well microplate format, was able to detect GSH depletion in algal cells. This cell-based assay can be used to evaluate decreases in GSH content due to exposure to toxicants. This assay is amenable to automation and may be useful in high-throughput toxicity screening using the alga P. subcapitata.  相似文献   

5.
&#110 -Glutamyltransferase (GGT) has a central role in glutathione homeostasis by initiating the breakdown of extracellular GSH. We investigated in the present study whether nitric oxide exposure of CC531 rat colon carcinoma cells modulates GGT and how the activity of the enzyme affects the level of intracellular GSH. The data show that GGT activity was induced in a dose-related manner by two NO-donors (spermineNONOate and nitrosoglutathione) and that antioxidants partly inhibited the induction. SpermineNONOate lowered intracellular GSH and induced apoptosis. Cultivating the cells in cystine-depleted medium also resulted in a 50% lowering of GSH, but this was avoided when GSH was added to the medium. This effect was mediated by the activity of GGT and shown after inhibiting GGT activity with acivicin and cyst(e)ine transporters with alanine and homocysteic acid. This shows that the cells benefit from GGT in maintaining the intracellular GSH level. Cells with induced GGT activity obtained after NO incubation showed a higher uptake rate of cysteine (2-fold), measured by incubating the cells with 35 S-radiolabeled GSH. The enzyme was also induced by interferon- &#110 and tumor necrosis factor- &#102, but this induction was not connected to activation of the endogenous nitric oxide synthase, as the addition of aminoguanidine, a NO-synthase inhibitor, did not affect the induction. The present study shows that the activity of GGT is upregulated by NO-donors and that the colon carcinoma cells, when cultivated in cystine-depleted medium, benefit from the enzyme in maintaining the intracellular level of GSH. Thus, the enzyme will add to the protective measures of the tumor cells during nitrosative stress.  相似文献   

6.
Age-related changes were examined in the distribution and severity of spontaneous lesions in the neuroepithelium and Bowman’s glands in mouse olfactory mucosa. The olfactory mucosa of female ICR mice at postnatal ages from 10 days to 16 months were investigated histologically by hematoxylin and eosin staining, high-iron diamine-Alcian blue (HID-AB) staining, and immunohistochemistry for olfactory marker protein (OMP), βIII tubulin (βIIIT), and Ki67. The lesions in the neuroepithelium and Bowman’s glands were quantitatively assessed by morphometric analyses of sections stained with anti-OMP antibody or HID-AB. The first appearance of neuroepithelial abnormality was observed in the dorsomedial portion of the olfactory mucosa in 5-month-old mice. The distribution and severity of lesions progressed with increasing age. In mildly affected epithelium in which OMP-positive olfactory receptor neurons (ORNs) were present but in smaller amounts, the numbers of βIIIT-positive and Ki67-positive neuroepithelial cells tended to be increased, indicating that neurogenesis was upregulated in these areas. In contrast, severely affected epithelium in which OMP-positive ORNs were virtually absent showed high variability in the numbers of βIIIT- and Ki67-positive cells among the areas examined, probably reflecting differences in the capacity of the basal cells remaining in the affected area to generate new neuronal cells. Histological analysis with HID-AB revealed that spontaneous lesions in Bowman’s glands also occurred in aged mouse olfactory mucosa. Lesions in the neuroepithelium and underlying Bowman’s glands tended to be spatially co-localized, suggesting a close association between pathogeneses in these two structures. Moreover, lesions in Bowman’s glands were associated with changes in the biochemical composition of mucus on the olfactory mucosa. This information should prove useful in improving the understanding of the pathogenetic mechanisms underlying age-related changes in the peripheral olfactory system. This work was supported by grants from the Japanese Ministry of Education, Culture, Sports, Science, and Technology (nos. 14770886, 16790987 and 18799002; K. Kondo) and a grant from the Japanese Ministry of Health, Labour, and Welfare (Comprehensive Research on Aging and Welfare, no. H13-choju-012; K. Nibu).  相似文献   

7.
The mechanism of glutathione (GSH) depletion by isoniazid (INH) was studied inM. smegmatis. INH increased the activity of γ-glutamyl transferase (GGT) whether added before medium inoculation or to actively growing cells. The activity of GGT in cells grown from the beginning in INH-containing medium increased significantly on growth days 2–6. Three-day oldM. smegmatis cells treated with INH exhibited a 30–65 % increase in the activity of GGT. The activities of γ-glutamyl-cysteine synthase (GGCS) and GSH synthase (GS) were lowered by 50 and 56 % respectively on the second day of growth whenM. smegmatis was grown in a medium supplemented with 1.5 mg INH per L. In 3-day oldM. smegmatis, INH significantly inhibited the activities of GSH biosynthetic enzymes. The results demonstrate that the increased activity of GGT and decreased activities of GSH biosynthetic enzymes are responsible for GSH depletion by INH inM. smegmatis.  相似文献   

8.
Rapid fluorescence techniques were evaluated for the detection of bacterial contaminants in papermaking chemicals including starch and the resin-based sizes and starch slurries used in the paper industry. Viable and non-viable bacterial cells were visualised by fluorescent probes and detected by epifluorescence microscopy and flow cytometry. The best discrimination ability was obtained with the fluorescent probes LIVE/DEAD and SYBR Green, based on the staining of cellular nucleic acid, and ChemChrome V3, which demonstrated cellular enzymatic activity. The process samples had to be diluted and filtered before fluorescence staining and analysis because they were viscous and contained solid particles. Fluorescence microscopic counts of bacteria in highly contaminated process samples were similar to plate counts, but flow cytometric enumeration of bacterial cells in process samples yielded 2- to 10-fold lower counts compared with plate counts, depending on the consistency of the sample. The detection limits in flow cytometric analysis and in epifluorescence microscopy were 103–106 cells ml−1 and 105–106 cells ml−1, respectively. Intrinsic bacterial contamination was detectable with fluorescence techniques and highly contaminated process samples could be analysed with fluorescence methods. Electronic Publication  相似文献   

9.
The chemo-electrical transduction process in olfactory neurons is accompanied by a rapid and transient increase in intracellular calcium concentrations. The notion that Na+/Ca2+ exchanger activities may play a major role in extruding calcium ions out of the cell and maintaining Ca2+ homeostasis in olfactory receptor cells was assessed by means of laser scanning confocal microscopy in combination with the fluorescent indicators Fluo-3 and Fura-Red. The data indicate that high exchanger acitivity, which was inhibited by amiloride derivatives, is located in the dendritic knob and probably in the olfactory cilia. This result was supported by experiments using specific antiserum raised against retinal Na+/Ca2+ exchanger protein which labelled an immunoreactive protein of 230 kDa in Western blots from olfactory tissue and strongly stained the ciliary layer of the olfactory epithelium.  相似文献   

10.
Olfactory sensory neurons are wrapped by ensheathing glial cells in the olfactory nerve layer (ONL). Neither functional roles nor electrical properties of ensheathing glial cells have been, as yet, fully clarified. Four subunits (SK1–4) of small conductance Ca2+-activated K+ (SK) channels have been cloned. In the present study, immunohistochemical analyses showed that SK3 channels are expressed in ensheathing glial cells in the rat olfactory bulb, in addition to neuronal cells in other regions. Western blotting analysis demonstrated that SK3 was predominantly expressed in the olfactory bulb, thalamus, moderately in the hippocampus and cerebellum and modestly in the cerebral cortex of the rat brain. SK3 immunoreactivity was detected in the ONL of the olfactory bulb, neural cell body and fibers of the substantia nigra and hypothalamus. SK3 immunoreactivity was quite intense in the outer (superficial) part of the ONL. SK3-immunoreactive structures were overlapped with glial fibrillary acidic protein (GFAP), but not with vimentin, markers for glial cells and olfactory sensory axons, respectively. Immunoelectron microscopy showed that SK3 immunoreactivity was localized in thin processes that enfolded fascicles of immunonegative olfactory nerve axons. These results indicate that SK3 is expressed specifically in the olfactory ensheathing glial cells in olfactory regions.This work was supported in part by a Grant-in-Aid to A.F. for Scientific Research from the Ministry of Education, Science, Sports and Culture of Japan, and by scholarship from Ono Pharmaceutical Company, and by Narishige Neuroscience Research Foundation.  相似文献   

11.
Propyl gallate (PG) as a synthetic antioxidant exerts a variety of effects on tissue and cell functions. Here, we investigated the effects of MAPK (MEK, JNK and p38) inhibitors on PG-treated HeLa cells in relation to cell death, ROS and GSH levels. PG induced cell growth inhibition and apoptosis in HeLa cells, which was accompanied by the loss of mitochondrial membrane potential (MMP; ΔΨm). ROS levels were increased or decreased in PG-treated HeLa cells depending on the incubation times. PG also increased GSH depleted cell numbers in HeLa cells. All the MAPK inhibitors slightly enhanced cell growth inhibition, death and MMP (ΔΨm) loss, and increased ROS levels in PG-treated HeLa cells. However, MAPK inhibitors did not significantly affect GSH depletion in PG-treated cells. In conclusion, the enhanced effect of MAPK inhibitors on PG-induced HeLa cell death was accompanied by increasing ROS levels but the effect was not related to changes of GSH level.  相似文献   

12.
Ganoderma lucidum powder using hot water and methanol extraction methods indicated a twofold more active cytotoxic activity with IC50 of 44 ± 3.8 μg/ml in the latter method. The representative dose-response curves of the G. lucidum crude extracts on J558 cell-lines revealed that there were great similarities between the curves which reflected rapid killing activities. The percentage viability of the J558 cell exposed to these crude extracts was dose dependent only up to 150 μg/ml. After which, there was no significant reduction when the dose was increased to 200 or 400 μg/ml. The morphological alterations induced by the crude extract were examined under the phase contrast, fluorescent and electron microscopy. When J558 cells were treated with doses higher than 50 μg/ml of the crude extract, obvious morphological changes and apoptosis occurred after 72 h. At 400 μg/ml, most of the cells showed necrosis characterized as small fragments with uniformly stained red nuclei. The apoptotic and necrotic cells increased by 16.5 and 29.1%, respectively whereas the viable cells decreased by as much as 45.6. The mode of cell death via apoptosis was 3.6% higher than necrosis. However, these morphological changes were not observed in the case of 3T3 cells. Results obtained from scanning electron microscopy and transmission electron microscopy further confirmed the occurrence of various apoptotic and necrotic features.  相似文献   

13.
It is well known that antioxidants containing sulfhydryl (−SH) groups are protective against the toxic effects of mercury. The current study was designed to elucidate the mechanism(s) of the cytoprotective effects of glutathione (GSH) and N-acetylcysteine (NAC) against the toxicity of inorganic mercury (HgCl2) in neuroblastoma cells (N-2A). The obtained results demonstrated the protective effects of these compounds in a dose dependant manner up to 95 and 74% cell viability, respectively as compared to the control of HgCl2 of 10%. The administration of buthionine sulfoximine (BSO), an inhibitor of GSH synthesis, increased the toxicity of HgCl2 in a dose dependent manner. Moreover, BSO treatment attenuated the levels of the cellular free −SH concentrations at low concentrations (1–100 μM) of HgCl2. The data also show that cellular thiol concentrations were augmented in the presence of GSH and NAC and these compounds were cytoprotective against HgCl2 and this is due to up regulating of GSH synthesis. A reduction in intracellular levels of GSH was observed with treatment of HgCl2. In addition, the ratio of GSH/GSSG increased from 16:1 to 50:1 from 1 to 10 μM concentration of HgCl2. The ratio of GSH/GSSG then decreased from 4:1 to 0.5:1 with the increase of concentration of HgCl2 between 100 μM and 1 mM due to the collapse of the N-2A cells. It was of interest to note that the synthesis of GSH was stimulated in cells exposed to low concentration of HgCl2 when extra GSH is available. These data support the idea that the loss of GSH plays a contributing role to the toxic effects of HgCl2 and that inorganic mercury adversely affects viability, through altering intracellular −SH concentrations. The data further indicate that the availability of GSH to the cells may not be sufficient to provide protection against mercury toxicity and the de novo synthesis of intracellular GSH is required to prevent the damaging effects of mercury.  相似文献   

14.
Efflux of glutathione (GSH) from astrocytes has been suggested as a key factor for neuroprotection by astrocytes. Here we evaluated if the Nrf2 activator curcumin affects basal and stimulated (Ca2+ omission) GSH efflux from cultures of astroglial cells. Stimulated efflux of GSH was observed at medium concentration of 0, 0.1 mM Ca2+, but not at 0.2 or 0.3 mM Ca2+. Astroglia treated with 30 μM curcumin increased the cellular content of GSH in parallel with elevated basal and stimulated efflux. Conversely treatment with buthionine sulfoximine lowered efflux of GSH. The efflux stimulated by Ca2+- omission was not affected by the P2X7-receptor antagonist Blue Brilliant G (100 nM) or the pannexin mimetic/blocking peptide 10Panx1 but inhibited by the gap junction blocker carbenoxolone (100 μM) and a hemichannel blocker Gap26 (300 μM). RNAi directed against Nrf2 partly inhibited the effect of curcumin. The results show that elevated cellular GSH by curcumin treatment enhance efflux from astroglial cells, a process which appear to be a prerequisite for astroglial mediated neuroprotection.  相似文献   

15.
This study investigated whether contact with the olfactory bulb was necessary for developing and renewing olfactory receptor neurons (ORNs) to attain normal odorant responsiveness, and whether the anatomical and functional recoveries of the olfactory epithelium were similar in both bulbectomized (BE) and bilaterally axotomized (AX) preparations. In vivo electrophysiological recordings were obtained in response to amino acids, a bile acid [taurolithocholic acid sulfate(TLCS)] and a pheromonal odorant [17α, 20β,-dihydroxy-4-pregnen-3-one (17,20P)] from sexually immature goldfish. Both transmission and scanning electron microscopy indicated that the olfactory epithelium degenerated in BE and AX goldfish. Within 1–2 weeks subsequent to the respective surgeries, responses to high concentrations (>0.1 mmol · l−1) of the more stimulatory amino acids remained, whereas responses were no longer obtainable to TLCS and 17,20P. At 4 weeks, responses to amino acid stimuli recovered to control levels, while responses to TLCS and 17,20P were minimal. By 7 weeks post bilateral axotomy, the olfactory epithelium recovered to a condition similar to control sensory epithelium; however, the rate of degeneration and proliferation of receptor neurons in BE preparations appeared to remain in balance, thus blocking further recovery of the olfactory epithelium. At 7 weeks post surgery, odorant responses of AX and BE goldfish to TLCS and 17,20P were still recovering. Accepted: 14 June 1997  相似文献   

16.
The role of glutathione (GSH) in the adaptation of wild type Arabidopsis thaliana plants to Cd stress was investigated. The nutrient solution (control or containing 50 or 100 μM Cd) was supplemented with buthionine sulfoximine (BSO; 50, 100, 500 μM, to decrease the GSH content in plants) or GSH (50, 100, 500 μM, to increase its content in plants) in order to find how GSH content could regulate Cd stress responses. BSO application did not influence plant biomass, while exogenous GSH (especially 500 μM) reduced root biomass. BSO (500μM) in combination with Cd (100 μM) increased Cd toxicity on root growth (by over 50 %), most probably due to reduced GSH content and phytochelatin (PC) accumulation (by over 96 %). On the other hand, combination of exogenous GSH (500 μM) with Cd (100 μM) was also more toxic to plants than Cd alone despite a significant increase in GSH and PC accumulation (up to 2.7 fold in the roots). This fact could indicate that the natural content of endogenous GSH in wild type A. thaliana plants is sufficient for Cd-tolerance. A decrease in this GSH content led to decreased Cd-tolerance of the plants but an increase in GSH content did not enhance Cd-tolerance, and it showed even toxic effect on the plants.  相似文献   

17.
This study was conducted to evaluate the protective effects of GABA (gamma-amino butyric acid)-enriched sea tangle juice (STJ) by Lactobacillus brevis BJ-20 fermentation against alcohol hepatotoxicity. The protective effects were determined by assessing glutathione (GSH) content levels and gamma-glutamyl transpeptidase (GGT) activity against ethanol-induced cytotoxicity in HepG2 cells. In ethanol-treated cells, GSH content decreased to 44.35% of control (ethanol-untreated cell) values; however, treatment with fermented sea tangle juice (FSTJ) at a concentration of 25 μg/mL increased GSH levels to 67.08%.These results suggest that FSTJ may prevent intracellular GSH depletion caused by ethanol consumption. Treatment with FSTJ against alcohol-injured HepG2 cells resulted in a dose-dependent decrease in GGT activity. The expression of cytochrome P450 2E1 (CYP2E1) enzyme, a major contributor to ethanol-induced oxidative stress, was also completely inhibited in FSTJ-treated cells at a concentration of 25 μg/mL. Thus, this study demonstrated that ethanolinduced cytotoxicity could be attenuated by inhibition of GSH depletion, GGT activity, and CYP2E1 expression.  相似文献   

18.
A new, non-destructive assay is described to quantify cytoplasmic glutathione (GSH) levels in vivo in single cells or populations of cells from Arabidopsis suspension cultures. Cytoplasmic GSH was labelled with monochlorobimane (MCB) in situ to give a fluorescent GSH-bimane (GSB) conjugate. At low (10-100 microM) concentrations of MCB, labelling was mediated by a glutathione S-transferase, which confers specificity for GSH. HPLC analysis of MCB-labelled low molecular-weight thiols showed that the assay measures the total GSH pool, including the oxidized glutathione. The progress curve for the labelling could be described using Michaelis-Menten kinetics with an apparent KM of 40 microM and Vmax of 470 micromol lcyt -1 min-1. There was no evidence for de novo synthesis of GSH during the labelling period of 2 h, suggesting that control of GSH synthesis is not mediated by feedback control of gamma-glutamylcysteine synthetase in this system. The total cellular level of GSH was calculated from the plateau value of the progress curve, after appropriate calibration, as 830-942 nmol g-1 FW. The volume fraction of cytoplasm was measured from serial optical sections of bimane-labelled cells collected by confocal laser scanning microscopy (CLSM) with excitation 442 nm, or two-photon laser scanning microscopy (TPLSM) with excitation 770 nm. A value of 42 +/- 3% cytoplasm was determined by manual segmentation, and a value of 37 +/- 2% using stereological techniques. Using these figures, values for cytoplasmic [GSH] were estimated to be between 2.7 +/- 0.3 and 3.2 +/- 0.3 mM for cell populations. In addition, measurement of GSH levels in individual cells using CLSM and TPLSM gave values of 3.0 +/- 0.5 and 3.5 +/- 0.7 mM, respectively.  相似文献   

19.
The objective of this study was to assess the effects of nickel chloride on human and rainbow trout erythrocytes in vitro. The cells were incubated with 0, 0.5 and 1 mM nickel chloride for 1 h at pH 7.40 and 25°C, then K+ efflux, SO42− uptake and GSH and GSSG concentrations were measured. In both kind of cells, “high concentration” nickel treatment increased KCl efflux with respect to the control. The SO42− uptake was not significantly different at “low nickel concentration” but was lower in erythrocytes treated with 1 mM nickel chloride; the rate constant of SO42− uptake decreased by 35% in human erythrocytes and by 44% in fish erythrocytes. Nickel chloride also acts on cellular metabolism and in particular on erythrocyte glutathione peroxidase with consequent increase in oxidative stress; the data show a significant decrease in intracellular GSH in both human (25%) and fish erythrocytes (18%) after treatment with nickel chloride, with concomitantly high GSSG concentrations and lower GSH/GSSG ratios.  相似文献   

20.
The presence of high levels, as well as tissue-specific forms, of cytochrome P450 enzymes in mammalian olfactory mucosa (OM) has important implications in the bioactivation and toxicity of xenobiotics entering the tissue. Previous studies have shown that coumarin, a known olfactory toxicant in rats, is bioactivated by OM microsomal P450s to a number of products, presumably via coumarin-3,4-epoxide and other epoxide intermediates. The aim of the current study was to obtain direct evidence for the formation of such reactive intermediates in rat OM through the detection of protein covalent binding and glutathione (GSH) adduct formation. Protein covalent binding experiments with [14C]coumarin (10 μM) displayed a 7–9-fold higher NADPH-dependent radioactivity binding in rat OM microsomes (2.5 nmol/mg/30 min) compared to those in rat and human liver microsomes; the binding value in rat OM microsomes was substantially but not completely reduced by the addition of GSH (5 mM). LC/MS analyses detected a number of GSH adducts in GSH-supplemented coumarin metabolism reaction in rat OM microsomes; 3-glutathionyl coumarin was found to be the major one, indicating 3,4-epoxidation as the main bioactivation pathway. Additional GSH adducts were identified, presumably forming via the same pathway or epoxidation on the benzene moiety. Our findings provide direct evidence for the formation of multiple coumarin reactive intermediates in rat OM, leading to protein covalent binding and GSH conjugation.  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