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 ability of the soluble glutathione S-transferases to bind the membrane (liposome) bound substrates 1-chloro-2,4-dinitrobenzene and sulfobromophthalein was determined. The transferases were found to have access only to substrates in the aqueous phase. They could not bind membrane-bound substrates and, thus, enzymatic activites were reduced by the membrane partitioning of the substrates. The reduction in enzymatic activity was directly proportional to the lipid solubility of the substrate. The liposomes had no direct effect on the enzyme per se. [35S]Sulfobromophthalein and [14C]chlorodinitrobenzene bound to liposomes were found to have rapid rates of release into the aqueous phase. Rates of hydration of chlorodinitrobenzene from liposomes were rapid enough such that rates of catalysis (measured in a stopped-flow spectrophotometer) were affected only by the partition coefficient of substrate between lipid phase and water, and not by the rate of transfer of substrate from lipid to water phase.  相似文献   

2.
Acrylamide, an α,β unsaturated electrophile and a cumulative neurotoxin, was found to react with glutathione giving rise to an S-conjugate of acrylamide. Glutathione S-transferase of rat liver and brain cytosols (active on both acrylamide and 1-chloro 2,4 dinitrobenzene) emerged as a single major peak on elution from Sephadex G-75. The enzymic conjugation of acrylamide with glutathione increased with protein and was dependent on incubation time and pH of medium. Acrylamide inhibited glutathione-S-transferase activity towards 1-chloro 2,4-dinitro-benzene of both liver and brain cytosol, in a concentration and time dependent manner. Enzyme catalyzed conjugation of acrylamide with glutathione was induced significantly by phenobarbital and t-SO (tans-stilbene oxide). The enzymic conjugation of acrylamide increased two fold from neonatal to adult and then showed a decreasing pattern at subsequent ages.  相似文献   

3.
Evidence is presented for an enzyme-catalyzed conjugation of acrylamide (ACR) in rat erythrocytes. Daily exposure of rats to ACR for a period of 7, 14 and 21 days resulted in a time-dependent decrease in glutathione content. In vitro incubation of ACR with rat erythrocytes suspension caused a concentration-dependent decrease in glutathione levels. Red blood cell (RBC) enzyme-catalyzed conjugation of ACR with glutathione increased with protein concentration and was dependent on pH and time of incubation. Glutathione-S-transferase (GST) activity using acrylamide and 1-chloro 2,4-dinitrobenzene (CDNB) as substrates followed the order: liver > kidney > brain > erythrocytes. Glutathione peroxidase activity of RBCs was inhibited by the in vitro addition of ACR to erythrocytes. These results suggest that rat erythrocytes are equipped with the mechanism which can inactivate toxic electrophilic chemicals, such as acrylamide.  相似文献   

4.
The amount of glutathione present in hepatic protein mixed disulfides was determined to be 20–30 nmole/g liver. This was established using two specific enzymatic methods: (a) the coupled assay with DTNB and glutathione (GSSG) reductase and (b) a newly developed test using GSH transferase and 1-chloro-2,4-dinitrobenzene for the estimation of GSH released from proteins after borohydride treatment; further, these results were confirmed by HPLC analysis. Thus, authentic glutathione makes up only 2–6% of the value for total protein mixed disulfides. The latter were determined with the generally employed o-phthalaldehyde assay, which is not necessarily specific for GSH. The amount of glutathione mixed disulfides depends linearly on the content of glutathione disulfide in the liver cell in the range studied. By increasing the GSSG levels from 20 to about 60 nmole/g liver with paraquat, nitrofurantoin or t-butyl hydroperoxide, glutathione protein mixed disulfides are increased by a similar amount.  相似文献   

5.
A member of the Theta class of human glutathione transferases (GST T1-1) was found to display the greatest catalytic activity towards the cytostatic drug 1,3-bis(2-chloroethyl)-1-nitrosourea (BCNU) of the GSTs studied. In this investigation (the most extensive to date), enzymes from four classes of the soluble human GSTs were heterologously expressed, purified, and kinetically characterized. From the 12 enzymes examined, only GST M2-2, GST M3-3 and GST T1-1 had significant activities with BCNU. This establishes that the activity is not a characteristic of a particular class of GSTs. Although GST M3-3 was previously reported to have the greatest activity with BCNU, the current investigation demonstrates that GST M2-2 is equally active and that GST T1-1 has an approximately 20-fold higher specific activity than either of the Mu class enzymes. A more rigorous kinetic analysis of GST T1-1 gave the following parameters with BCNU: a k(cat) of 0.035 +/-0.003s(-1) and a K(M) of 1.0 +/- 0.1mM. The finding that GST T1-1 has the highest activity towards BCNU is significant since GST T1-1 is expressed in the brain, a common target for BCNU treatment. Furthermore, the existence of a GST T1-1 null allele in up to 60% in some populations, may influence both the sensitivity of tumors to chemotherapy and the severity of adverse side-effects in patients treated with this agent.  相似文献   

6.
Induction of cytosolic long-chain acyl-CoA hydrolases was investigated in rat liver after administration of various peroxisome proliferators and related compounds. Treatment of rats with di-(2-ethylhexyl)-phthalate, di-(2-ethylhexyl)-adipate or tiadenol induced hydrolases I and II, while acetylsalicylic acid induced only hydrolase II. Among the various phenoxyacetic acid derivatives and related compounds, 2,4,5-trichlorophenoxyacetic acid, 2-(4-chlorophenoxy)-2-methylacetic acid, 2-(2-chlorophenoxy)-2-methylpropionic acid and clofibric acid induced both hydrolases I and II, whereas 2, 4-dichlorophenoxyacetic acid induced only hydrolase II. All nine of the above-mentioned inducers also markedly increased the activity of peroxisomal beta-oxidation. Other compounds tested (2-chlorophenoxyacetic acid, 4-chlorophenoxyacetic acid, 4-chlorophenol, phenoxyacetic acid and phenoxy-2-methylacetic acid) were ineffective as inducers. These results suggest that inducers of acyl-CoA hydrolase II also enhance peroxisomal beta-oxidation activity, but do not necessarily induce acyl-CoA hydrolase I. The structure-inducing activity relationships of these compounds are discussed.  相似文献   

7.
Glutathione S-transferases (GSTs) are multifunctional detoxification proteins that protect the cell from electrophilic compounds. Overexpression of GSTs in cancer results in resistance to chemotherapeutic agents and inhibition of the over expressed GST has been suggested as an approach to combat GST-induced resistance. The inhibition of human recombinant GSTs by natural plant products was investigated in this study. Using 1-chloro-2,4 dinitrobenzene (CDNB) as a substrate, ellagic acid and curcumin were shown to inhibit GSTs A1-1, A2-2, M1-1, M2-2 and P1-1 with IC(50) values ranging from 0.04 to 5 microM whilst genistein, kaempferol and quercetin inhibited GSTs M1-1 and M2-2 only. The predominant mode of inhibition with respect to the G and H-sites were mixed inhibition and uncompetitive to a lesser extent. The K(i) (K(i)(')) values for ellagic acid and curcumin with respect to GSH and CDNB were in the range 0.04-6 microM showing the inhibitory potency of these polyphenolic compounds. Ellagic acid and curcumin also showed time- and concentration-dependent inactivation of GSTs M1-1, M2-2 and P1-1 with curcumin being a more potent inactivator than ellagic acid. These results facilitate the understanding of the interaction of human GSTs with plant polyphenolic compounds with regards to their role as chemomodulators in cases of GST-overexpression in malignancies.  相似文献   

8.
The herbicieds 2-methyl-4-chlorophenoxy acetic acid (MCPA) and 2-(2-methyl-4-chlorophenoxy) propionic acid (MCPP or mecoprop) were tested for 90 days in rats. The compounds were added to the diet at levels of 0, 50, 400, and 3200 ppm. Growth, food intake, mortality, haematology, blood and liver chemistry, organ weights and histopathology were used as criteria. The main effects of both compounds were growth retardation and elevated relative kidney weights at levels of 400 ppm and more. The 50 ppm dose level can be considered as a no-toxic-effect level in the 90-day study. In subacute dermal studies in rabbits during 3 weeks the dosages were 0, 0.5, 1.0 and 2 g MCPA or MCPP per kg body weight. Therafter followed a recovery period of 2 weeks. Growth, mortality, skin reaction, haematology, organ weights (MCPP) and histopathology were recorded and determined. Both compounds caused slight to moderate erythema at all dose levels, whereas elasticity of the skin was decreased. In both experiments the skin returned to normal during the recovery period. Weight loss was observed at all dose levels. In the MCPA experiment high mortality and histopathological changes in the liver, kidneys, spleen and thymus were recorded at the two highest dose levels. The cause of this could have been either the treatment with MCPA or a dysbacteria infection which developed during the experiment. Oral and intraperitoneal acute toxicity of MCPP for the rat were found to be 1210 and 402 mg/kg, respectively. After a single oral or dermal application of MCPA to the rabbit, the compound was excreted unchanged in the urine.  相似文献   

9.
The reaction of (E)-2,4-pentadienoic acid with aqueous bromine was reinvestigated to affirm the formation of (E)-5-bromo-4-hydroxy-2-pentenoic acid, whose structure was confirmed by the spectroscopic methods as well as the chemical modification.  相似文献   

10.
The effects of polysaccharide peptide (PSP), an immunomodulator isolated from Coriolus versicolor COV-1, on glutathione (GSH) and GSH-related enzymes was investigated in C57 mouse. Administration of PSP (1-4 micromole/kg, i.p.) produced a transient, dose-dependent depletion (10-37%) of hepatic GSH, with no effect on serum glutamic-pyruvic transaminase (SGPT) activity. Blood GSH was depleted (6-25%) at 3 h, followed by a rebound increase above the control GSH level (20%) at 18 h. The GSSG/GSH ratio, a measure of oxidative stress, was increased 3 h after PSP treatment but returned to normal levels at 24 h. Sub-chronic treatment of PSP (1-4 micromole/kg/day, i.p.) for seven days did not produce any significant changes in hepatic GSH levels and the GSSG/GSH ratio when measured 24 h after the final dose of PSP. PSP had little effect on glutathione transferase (GST), glutathione reductase (GSSG reductase) and glutathione peroxidase (GPX) activities in the liver. However, a dose-dependent increase in blood GPX activity (30-48%) was observed at 3h, which coincided with the increase in the GSSG/GSH ratio. The increase in blood GPX activity may be a responsive measure to deal with the transient oxidative stress induced by PSP treatment. The results showed that PSP only caused a transient perturbation on hepatic glutathione without affecting the GSH-related enzymes such as GST, GSSG reductase and GPX. The observed changes in blood GSH simply reflected the intra-organ translocation of glutathione, as the glutathione-related enzymes were not significantly affected by PSP treatment.  相似文献   

11.
Specific activities and apparent Michaelis-Menten kinetic parameters were determined for glutathione (GSH) S-transferase activity (E.C. 2.5.1.18) in rat liver cytosol, towards styrene oxide (STOX), 1,2-butylene oxide (BOX) and 1-chloro-2,4-dinitrobenzene (CDNB) as electrophilic substrates, before and after pretreatment with the drug-metabolizing enzyme inducers phenobarbital (PB), 3-methylcholanthrene (MC) and 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD). The measured GSH S-transferase activities appear to obey Michaelis-Menten kinetics. In non-induced animals the apparent Km values of the transferase activities were equal for STOX vs GSH, but they differed by a factor of 2 for CDNB vs GSH and by a factor of 14 for BOX vs GSH. The apparent Vmax values in each combination of GSH and electrophilic substrate were equal, but differed by one order of magnitude for the mutual substrate combinations. Pretreatment of the rats with MC resulted in enhancement of all measured activities expressed in terms of cytosol protein, while TCDD only enhanced the activities expressed as per gram body wt. PB enhanced both activities when STOX was employed as substrate, but when CDNB was used as the substrate, only the activity per gram body wt increased. All pretreatments increased the Vmax values using CDNB as the substrate, while PB and MC had an enhancing effect using STOX; the Vmax using BOX was enhanced after TCDD administration only. The Km values using BOX as the substrate was lowered after MC pretreatment; TCDD pretreatment decreased the Km using STOX, while it increased the Km using CDNB. It is concluded that the GSH S-transferase system is inducible, but in contrast to the induction of the mixed function oxidase system, qualitative differences between the inducing effects of PB and MC were not observed. Use of TCDD as inducing agent, however. resulted in a different induction pattern, which may indicate that during induction with this agent different types of GSH S-transferases are involved.  相似文献   

12.
6-甲基-4-(1H)-吡啶酮-3-羧酸的制备   总被引:1,自引:0,他引:1  
目的制备 6 甲基 4 (1H) 吡啶酮 3 羧酸。方法以 4 羟基 6 甲基 2 吡喃酮、N ,N 二甲基甲酰胺二甲氧基缩醛为起始原料 ,经两步反应制得目标化合物。结果与结论经熔点测定及1H NMR、MS分析确证目标产物结构 ,总收率为 39 6 % ,高于文献收率  相似文献   

13.
对头孢布烯的关键中间体2-(2-苄氧羰基氨基噻唑-4-基)-5-(3-甲基-2-丁烯氧羰基)-2-戊烯酸(1)的合成工艺进行了研究。选用国内易得的(2-氨基噻唑-4-基)乙酸甲酯(2)作为起始原料,经过氨基保护、M ichae l加成-消除和选择性酯化三步反应制得目标化合物1,反应总收率63.0%。该工艺操作简单,生产成本低,适合工业化生产。  相似文献   

14.
Trichloroethylene (TCE, CAS 79-01-6) is a widely used industrial chemical, and a common environmental pollutant. TCE is a well-known carcinogen in rodents and is classified as “probably carcinogenic to humans”. Several analytical methods have been proposed for detection of TCE metabolites in biological media utilizing derivatization-free techniques; however, none of them is suitable for simultaneous detection of both oxidative metabolites and glutathione conjugates of TCE in small volume biological samples. Here, we report a new combination of methods for assessment of major TCE metabolites: dichloroacetic acid (DCA), trichloroacetic acid (TCA), S-(1,2-dichlorovinyl)-L-cysteine (DCVC), and S-(1,2-dichlorovinyl) glutathione (DCVG). First, DCA and TCA were extracted with ether. Second, the remaining aqueous fraction underwent solid phase extraction for DCVC and DCVG. Then, DCA and TCA were measured by hydrophilic interaction liquid chromatography ion exchange negative electrospray ionization tandem mass spectrometry, while DCVC and DCVG were measured by reverse phase positive electrospray ionization tandem mass spectrometry. This method was applied successfully to measure all 4 TCE metabolites in as little as 50 μl of serum from mice orally exposed to TCE (2100 mg/kg, 2 h). Serum concentrations (mean ± standard deviation) of the TCE metabolites obtained with this method are comparable or equivalent to those previously reported in the literature: DCA, 0.122 ± 0.014 nmol/ml (limit of detection: 0.01 nmol/ml); TCA, 256 ± 30 nmol/ml (0.4 nmol/ml); DCVG, 0.037 ± 0.015 nmol/ml (0.001 nmol/ml); DCVC, 0.0024 ± 0.0009 nmol/ml (0.001 nmol/ml). This method opens new opportunities to increase throughput and decrease number of animals required for mechanistic studies on TCE in rodents.  相似文献   

15.
Incubation of rat liver cell-free extracts with an NADPH-generating system and with nifurtimox or benznidazole (two nitroheterocyclic drugs used in the treatment of Chagas' disease) produced oxidation of reduced glutathione (GSH) and increased lipid peroxidation, as shown by the generation of thiobarbituric-acid-reacting intermediates. Nifurtimox and benznidazole inhibited GSSG-reductase, but not GSH-peroxidase, the former inhibition contributing to GSH depletion. In every case, nifurtimox was more effective than benznidazole. Addition of GSH or free-radical scavengers (catalase, superoxide dismutase, mannitol, sodium benzoate or L-histidine) prevented the effect of nifurtimox on lipid peroxidation reactions. These results support the assumption [M. Dubin, S. N. J. Moreno, E. E. Martino, R. Docampo and A. O. M. Dubin, Biochem. Pharmac.32, 483 (1983)] that, in the rat liver, GSH exerts a protective action against oxygen radicals generated by the nitroheterocyclic drugs.  相似文献   

16.
(RS)-2-Amino-3-(5-tert-butyl-3-hydroxy-4-isoxazolyl)propionic acid (ATPA), an analogue of (RS)-2-amino-3-(3-hydroxy-5-methyl-4-isoxazolyl)propionic acid (AMPA), has previously been shown to be a relatively weak AMPA receptor agonist and a very potent agonist at the GluR5 subtype of kainic acid-preferring (S)-glutamic acid ((S)-Glu) receptors. We report here the separation of (+)- and (−)-ATPA, obtained at high enantiomeric purity (enantiomeric excess values of 99.8% and >99.8%, respectively) using chiral chromatography, and the unequivocal assignment of the stereochemistry of (S)-(+)-ATPA and (R)-(−)-ATPA. (S)- and (R)-ATPA were characterized in receptor binding studies using rat brain membranes, and electrophysiologically using the rat cortical wedge preparation and cloned AMPA-preferring (GluR1, GluR3, and GluR4) and kainic acid-preferring (GluR5, GluR6, and GluR6+ KA2) receptors expressed in Xenopus oocytes. In the cortical wedge, (S)-ATPA showed AMPA receptor agonist effects (EC50=23 μM) approximately twice as potent as those of ATPA. (R)-ATPA antagonized depolarizations induced by AMPA (Ki=253 μM) and by (S)-ATPA (Ki=376 μM), and (R)-ATPA antagonized the biphasic depolarizing effects induced by kainic acid (Ki=301 μM and 1115 μM). At cloned AMPA receptors, (S)-ATPA showed agonist effects at GluR3 and GluR4 with EC50 values of approximately 8 μM and at GluR1 (EC50=22 μM), producing maximal steady state currents only 5.4–33% of those evoked by kainic acid. (R)-ATPA antagonized currents evoked by kainic acid at cloned AMPA receptor subtypes with Ki values of 33–75 μM. (S)-ATPA produced potent agonist effects at GluR5 (EC50=0.48 μM). Due to desensitization of GluR5 receptors, which could not be fully prevented by treatment with concanavalin A, (S)-ATPA-induced agonist effects were normalized to those of kainic acid. Under these circumstances, maximal currents produced by (S)-ATPA and kainic acid were not significantly different. (R)-ATPA did not attenuate currents produced by kainic acid at GluR5, and neither (S)- nor (R)-ATPA showed significant effects at GluR6. (S)-ATPA as well as AMPA showed weak agonist effects at heteromeric GluR6+KA2 receptors, whereas (R)-ATPA was inactive. Thus, (S)- and (R)-ATPA may be useful tools for mechanistic studies of ionotropic non-NMDA (S)-Glu receptors, and lead structures for the design of new subtype-selective ligands for such receptors.  相似文献   

17.
Cephalothin, penicillin G and probenecid inhibited GSH organic nitrate ester reductase (ONER) and several other enzymatic activities of GSH-S-transferases (EC 2.5.1.18) from rat and guinea pig liver. Erythrityl tetranitrate, a substrate for ONER, inhibited the aryl and alkyl transferase activities of two guinea pig liver GSH-S-transferases. These findings support the concept that ONER is one of the several activities possessed by the GSH-S-transferases. In an examination of possible in vivo action, parenteral administration of these inhibitors 2–30 min prior to i.v. administration of [14C]glyceryl trinitrate resulted in a 50–100 per cent increase in the half-time of the metabolism phase of [14C]glyceryl trinitrate clearance from the blood and postponed the appearance of metabolites. This presumably occurs through the in vivo inhibition of GSH-ONER activity of the GSH-S-transferases and suggests a possible means of prolonging the pharmacologie action of nitrate esters.  相似文献   

18.
Bladder cancer is highly recurrent following specific transurethral resection and intravesical chemotherapy, which has prompted continuing efforts to develop novel therapeutic agents and early-stage diagnostic tools. Specific changes in protein expression can provide a diagnostic marker. In our present study, we investigated changes in protein expression during urothelial carcinogenesis. The carcinogen BBN was used to induce mouse bladder tumor formation. Mouse bladder mucosa proteins were collected and analyzed by 2D electrophoresis from 6 to 20 weeks after commencing continuous BBN treatment. By histological examination, the connective layer of the submucosa showed gradual thickening and the number of submucosal capillaries gradually increased after BBN treatment. At 12-weeks after the start of BBN treatment, the urothelia became moderately dysplastic and tumors arose after 20-weeks of treatment. These induced bladder lesions included carcinoma in situ and connective tissue invasive cancer. In protein 2D analysis, the sequentially downregulated proteins from 6 to 20 weeks included GSTM1, L-lactate dehydrogenase B chain, keratin 8, keratin 18 and major urinary proteins 2 and 11/8. In contrast, the sequentially upregulated proteins identified were GSTO1, keratin 15 and myosin light polypeptide 6. Western blotting confirmed that GSTM1 and NQO-1 were decreased, while GSTO1 and Sp1 were increased, after BBN treatment. In human bladder cancer cells, 5-aza-2′-deoxycytidine increased the GSTM1 mRNA and protein expression. These data suggest that the downregulation of GSTM1 in the urothelia is a biomarker of bladder carcinogenesis and that this may be mediated by DNA CpG methylation.  相似文献   

19.
2, 3, 5, 6-Tetrachloro-1, 4-benzoquinone (TCBQ) is a metabolite of pentachlorophenol known to react with cysteines of glutathione transferases (GSTs). TCBQ treatment of rat kidney rGSTA1-2 and rGSTA1-1 abolishes 70-80% conjugation of glutathione (GSH) to 1-chloro-2, 4-dinitrobenzene and results in strongly correlated quenching of intrinsic fluorescence of Trp-20 (R>0.96). rGSTA2-2 is only inhibited by 25%. Approximately 70% (rGSTA1-1) and 60% (rGSTA1-2) conjugation activity is abolished at TCBQ: GST stoichiometries near 1:1. The inactivation follows a Kitz/Wilson model with K(D) of 4.77+/-2.5microM for TCBQ and k(3) for inactivation of 0.036+/-0.01min(-1). A single tryptic peptide labelled with TCBQ was isolated from kidney rGSTA1-2 containing Cys-17 which we identify as the site of modification. Treatment with more than stoichiometric amounts of TCBQ modified other residues but resulted in only modest further inhibition of catalysis. We interpret these findings in terms of localised steric effects on the relatively rigid alpha-helix 1 adjacent to the catalytic site of subunit 1 possibly affecting the Alpha class-specific alpha-helix 9 which acts as a "lid" on the hydrophobic part of the active site. Homology modelling of rGSTA1-1 modified at Cys-17 of one subunit revealed only modest structural perturbations in the second subunit and tends to exclude global structural effects.  相似文献   

20.
The objective of this study was to investigate the structural requirements necessary for inhibition of glutathione S-transferase P1-1 (GSTP1-1) and GS-X pump (MRP1 and MRP2) activity by structurally related flavonoids, in GSTP1-1 transfected MCF7 cells (pMTG5). The results reveal that GSTP1-1 activity in MCF7 pMTG5 cells can be inhibited by some flavonoids. Especially galangin was able to inhibit almost all cellular GSTP1-1 activity upon exposure of the cells to a concentration of 25microM. Other flavonoids like kaempferol, eriodictyol and quercetin showed a moderate GSTP1-1 inhibitory potential. For GSTP1-1 inhibition, no specific structural requirements necessary for potent inhibition could be defined. Most flavonoids appeared to be potent GS-X transport inhibitors with IC(50) values ranging between 0.8 and 8microM. Luteolin and quercetin were the strongest inhibitors with IC(50) values of 0.8 and 1.3microM, respectively. Flavonoids without a C2-C3 double bond like eriodictyol, taxifolin and catechin did not inhibit GS-X pump activity. The results of this study demonstrate that the structural features necessary for high potency GS-X pump inhibition by flavonoids are (1) the presence of hydroxyl groups, especially two of them generating the 3',4'-catechol moiety; and (2) a planar molecule due to the presence of a C2-C3 double bond. Other factors, like lipophilicity and the total number of hydroxyl groups do not seem to be dominating the flavonoid-mediated GS-X pump inhibition. To identify the GS-X pump responsible for the DNP-SG efflux in MCF7 cells, the effects of three characteristic flavonoids quercetin, flavone and taxifolin on MRP1 and MRP2 activity were studied using transfected MDCKII cells. All three flavonoids as well as the typical MRP inhibitor (MK571) affected MRP1-mediated transport activity in a similar way as observed in the MCF7 cells. In addition, the most potent GS-X pump inhibitor in the MCF7 cells, quercetin, did not affect MRP2-mediated transport activity. These observations clearly indicate that the GS-X pump activity in the MCF7 cells is likely to be the result of flavonoid-mediated inhibition of MRP1 and not MRP2. Altogether, the present study reveals that a major site for flavonoid interaction with GSH-dependent toxicokinetics is the GS-X pump MRP1 rather than the conjugating GSTP1-1 activity itself. Of the flavonoids shown to be most active especially quercetin is frequently marketed in functional food supplements. Given the physiological levels expected to be reached upon supplement intake, the IC(50) values of the present study point at possible flavonoid-drug and/or flavonoid-xenobiotic interactions especially regarding transport processes involved in toxicokinetics.  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