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 31 毫秒
1.
Naphthalene-1,2-oxide (NPO), 1,2-naphthoquinone (1,2-NPQ) and 1,4-naphthoquinone (1,4-NPQ) are the major metabolites of naphthalene that are thought to be responsible for the cytotoxicity and genotoxicity of this chemical. We measured cysteinyl adducts of these metabolites in hemoglobin (Hb) and albumin (Alb) from F344 rats dosed with 100-800 mg naphthalene per kg body weight. The method employs cleavage and derivatization of these adducts by trifluoroacetic anhydride and methanesulfonic acid followed by gas chromatography-mass spectrometry in negative ion chemical ionization mode. Cysteinyl adducts of both proteins with NPO, and 1,2- and 1,4-NPQ (designated NPO-Hb and -Alb, 1,2-NPQ-Hb and -Alb, and 1,4-NPQ-Hb and -Alb, respectively) were produced in a dose-dependent manner. Of the two structural isomers resulting from NPO, levels of NPO1 adducts were greater than those of NPO2 adducts in both Hb and Alb, indicating that aromatic substitution is favored in vivo at positions 1 over 2. Of the quinone adducts, 1,2-NPQ-Hb and -Alb were produced in greater quantities than 1,4-NPQ-Hb and -Alb, indicating either that the formation of 1,2-NPQ from NPO is favored or that more than one pathway leads to the formation of 1,2-NPQ. The shapes of the dose-response curves were generally nonlinear at doses above 200 mg naphthalene per kg body weight. However, the nature of nonlinearity differed, showing evidence of supralinearity for NPO-Hb, NPQ-Hb and NPQ-Alb and of sublinearity for NPO-Alb. Low background levels of 1,2-NPQ-Hb and -Alb and 1,4-NPQ-Hb and -Alb were detected in control animals without known exposure to naphthalene. However, the corresponding NPO-Hb and -Alb adducts were not detected in control animals.  相似文献   

2.
Both 1,2-naphthoquinone (1,2-NPQ) and 1,4-naphthoquinone (1,4-NPQ) are reactive metabolites of naphthalene that are thought to be responsible for the naphthalene-induced cytotoxicity and genotoxicity. The aim of this study was to investigate the cumulative tissue dose of 1,2-NPQ and 1,4-NPQ in human serum derived from blood donors in Taiwan via measurements of albumin adducts by a methodology, which employs trifluoroacetic acid anhydride and methanesulfonic acid to selectively cleave cysteinyl adducts on proteins. Both 1,2-NPQ and 1,4-NPQ adducts were detected in all male and female subjects (n = 22). The median levels of 1,2-NPQ adduct in human subjects were estimated to be 268 (range 139-857) and 203 (range 128-1352) (pmol/g) in male (n = 11) and female (n = 11) subjects, respectively. In contrast, the median levels of 1,4-NPQ adduct were estimated to be 45.0 (range 22.0-117) and 38.9 (range 21.5-172) (pmol/g) in male and female subjects, respectively. We noticed that levels of 1,2-NPQ adduct were significantly correlated with those of 1,4-NPQ adduct (correlation coefficient r = 0.643, p < 0.01). Results from in vitro experiments confirmed that the production of naphthoquinones-derived adducts on serum albumin increased with increased concentration of naphthoquinones (0-100 μM). Linear relationships were observed over the range of concentration. Time-course experiments suggested that both 1,2-NPQ and 1,4-NPQ-derived adducts rapidly reached maximum values at 10 min mark and remained constant thereafter. The reaction rate constant analyses indicated that the second-order rate constants, representing in vitro reactions between naphthoquinones and cysteine residues of serum albumin, were estimated to be 0.0044/0.0002 L(g protein)−1 h−1, respectively. Overall, the cumulative tissue doses of 1,4-NPQ (217-316 nM h) in male and female subjects were ∼3-fold greater than those of 1,2-NPQ (76-98 nM h) in the study population. The initial concentrations of serum 1,2-NPQ and 1,4-NPQ in the study population were estimated to be between 145-188 and 807-1175 nM, respectively. We conclude that the relatively large amounts of naphthoquinones present in human serum may point to toxicological consequences.  相似文献   

3.
Naphthalene is considered by the US Environmental Protection Agency to be a carcinogenic compound based on inhalation studies in rats. The primary metabolite of naphthalene is naphthalene 1,2-arene oxide. This unstable intermediate can lead to formation of 1-naphthol and naphthalene-1,2-dihydrodiol. Secondary metabolites include 1,2-dihydroxynaphthalene (1,2-DHN), which can be further oxidized to 1,2-naphthoquinone (1,2-NQ). Based on the metabolism of naphthalene and its similarity to the metabolic activation of carcinogenic natural estrogens, synthetic estrogens and benzene, we hypothesize that naphthalene is activated to initiate cancer by reaction of 1,2-NQ with DNA to form the depurinating adducts 1,2-DHN-4-N3Ade and 1,2-DHN-4-N7Gua. These adducts were synthesized by reaction of 1,2-NQ with Ade or dG in acetic acid/water/DMF (1:1:1). 1,2-NQ was reacted with DNA, and the depurinating 1,2-DHN-4-N3Ade and 1,2-DHN-4-N7Gua adducts were analyzed by ultraperformance liquid chromatography/tandem mass spectrometry and HPLC with electrochemical detection. After the reaction of 1,2-NQ with DNA, the N3Ade and N7Gua adducts were found. Similarly, when 1,2-DHN was activated by tyrosinase in the presence of DNA, higher amounts of the N3Ade and N7Gua adducts were detected. These same adducts were also formed when 1,2-DHN was activated by prostaglandin H synthase or 3-methylcholanthrene-induced rat liver microsomes in the presence of DNA. These depurinating adducts are analogous to those obtained from the ortho-quinones of natural estrogens, synthetic estrogens and benzene. These results suggest that reaction of ortho-quinones with DNA by 1,4-Michael addition is a general mechanism of weak carcinogenesis that occurs with naphthalene and a number of other aromatic compounds.  相似文献   

4.
Naphthalene is a volatile polycyclic aromatic hydrocarbon generated during combustion and is a ubiquitous chemical in the environment. Short term exposures of rodents to air concentrations less than the current OSHA standard yielded necrotic lesions in the airways and nasal epithelium of the mouse, and in the nasal epithelium of the rat. The cytotoxic effects of naphthalene have been correlated with the formation of covalent protein adducts after the generation of reactive metabolites, but there is little information about the specific sites of adduction or on the amino acid targets of these metabolites. To better understand the chemical species produced when naphthalene metabolites react with proteins and peptides, we studied the formation and structure of the resulting adducts from the incubation of model peptides with naphthalene epoxide, naphthalene diol epoxide, 1,2-naphthoquinone, and 1,4-naphthoquinone using high resolution mass spectrometry. Identification of the binding sites, relative rates of depletion of the unadducted peptide, and selectivity of binding to amino acid residues were determined. Adduction occurred on the cysteine, lysine, and histidine residues, and on the N-terminus. Monoadduct formation occurred in 39 of the 48 reactions. In reactions with the naphthoquinones, diadducts were observed, and in one case, a triadduct was detected. The results from this model peptide study will assist in data interpretation from ongoing work to detect peptide adducts in vivo as markers of biologic effect.  相似文献   

5.
Metabolism of naphthalene by Cunninghamella elegans.   总被引:11,自引:7,他引:4       下载免费PDF全文
Cunninghamella elegans grown on Sabouraud dextrose broth in the presence of naphthalene produced six metabolites. Each product was isolated and identified by conventional chemical techniques. The major metabolites were 1-naphthol (67.9%) and 4-hydroxy-1-tetralone (16.7%). Minor products isolated were 1,4-naphthoquinone (2.8%), 1,2-naphthoquinone (0.2%), 2-naphthol (6.3%), and trans-1,2-dihydroxy-1,2-dihydronaphthalene (5.3%). C. elegans oxidized both 1-naphthol and 1,4-naphthoquinone to 4-hydroxy-1-tetralone. The results suggest that C. elegans oxidizes naphthalene by a sequence of reactions similar to those reported for the mammalian metabolism of this hydrocarbon.  相似文献   

6.
1,3-Butadiene and isoprene (2-methyl-1,3-butadiene) are chemically related substances that are carcinogenic to rodents. The overall aim of this work is to elucidate the role of the genotoxic action of diepoxide metabolites in the carcinogenesis of the dialkenes. In vivo doses of the diepoxide metabolites were measured through reaction products with hemoglobin (Hb adducts) in studies of induced micronuclei (MN) in rodents. In the reaction with N-terminal valine in Hb, diepoxybutane and isoprenediepoxide form ring-closed adducts, pyrrolidines [N,N-(2,3-dihydroxy-1,4-butadiyl)valine and N,N-(2,3-dihydroxy-2-methyl-1,4-butadiyl)valine, respectively]. The method applied for Hb-adduct measurement is based on tryptic degradation of the protein and liquid chromatography electrospray ionisation tandem mass spectrometry (LC-ESI-MS/MS) analysis. Mice were given single i.p. injections of the monoepoxides of butadiene and isoprene, 1,2-epoxy-3-butene or 1,2-epoxy-2-methyl-3-butene, respectively. Rats were treated in the same way with 1,2-epoxy-3-butene. In mice pyrrolidine adduct levels increased with increasing administered doses of the monoepoxides. The in vivo dose of diepoxybutane was on average twice as high (0.29+/-0.059 mMh) as the in vivo dose of isoprenediepoxide (0.15+/-0.053 mMh) per administered dose (mmol/kg body weight) of the monoepoxides. In mice the genotoxic effects of the two monoepoxides, measured as the increase in the frequencies of micronuclei (MN), were approximately linearly correlated to the in vivo doses of the diepoxides (except at the highest dose of diepoxybutane). In rats the pyrrolidine-adduct levels from diepoxybutane were below the limit of quantification at all administered doses of 1,2-epoxy-3-butene and no significant increase was observed in the frequency of MN. Measurement of the ring-closed adducts to N-termini in Hb by the applied method permits analysis of in vivo doses of diepoxybutane and isoprenediepoxide, which may be further used for the elucidation of the mechanisms of carcinogenesis of butadiene and isoprene.  相似文献   

7.
Naphthalene has been shown to be a weak carcinogen in rats. To investigate its mechanism of metabolic activation and cancer initiation, mice were topically treated with naphthalene or one of its metabolites, 1-naphthol, 1,2-dihydrodiolnaphthalene (1,2-DDN), 1,2-dihydroxynaphthalene (1,2-DHN), and 1,2-naphthoquinone (1,2-NQ). After 4 h, the mice were sacrificed, the treated skin was excised, and the depurinating and stable DNA adducts were analyzed. The depurinating adducts were identified and quantified by ultraperformance liquid chromatography/tandem mass spectrometry, whereas the stable adducts were quantified by 32P-postlabeling. For comparison, the stable adducts formed when a mixture of the four deoxyribonucleoside monophosphates was treated with 1,2-NQ or enzyme-activated naphthalene were also analyzed. The depurinating adducts 1,2-DHN-1-N3Ade and 1,2-DHN-1-N7Gua arise from reaction of 1,2-NQ with DNA. Similarly, the major stable adducts appear to derive from the 1,2-NQ. The depurinating DNA adducts are, in general, the most abundant. Therefore, naphthalene undergoes metabolic activation to the electrophilic ortho-quinone, 1,2-NQ, which reacts with DNA to form depurinating adducts. This is the same mechanism as other weak carcinogens, such as the natural and synthetic estrogens, and benzene.  相似文献   

8.
A gas chromatography-mass spectrometry (GC-MS) method was developed for measuring 1,2-dihydroxynaphthalene (1,2-DHN) and 1,4-dihydroxynaphthalene (1,4-DHN) in urine. The method involves enzymatic digestion of urinary conjugates to release the DHNs which were then analyzed as trimethylsilyl derivatives by GC-MS. For 1,2-DHN and 1,4-DHN, respectively, the assay limits of detection were 0.21 and 0.15 microg/l, the assay limits of quantitation were 0.69 and 0.44 microg/l, and the coefficients of variation were 14.7 and 10.9%. This method was successfully applied to determine urinary levels of 1,2-DHN and 1,4-DHN in coke workers (14 top workers and 13 side-bottom workers) and 21 matching control workers from the steel industry of northern China. The geometric mean (GM) levels of 1,2-DHN were approximately 100 and 30 times higher than those of 1,4-DHN in exposed and control subjects, respectively. The GM levels 1,2-DHN and 1,4-DHN were significantly higher for coke workers (1,2-DHN: top workers--552 microg/l, side-bottom workers--260 microg/l; 1,4-DHN: top workers--3.42 microg/l, side-bottom workers--3.56 microg/l) than for controls (1,2-DHN: 38.8 microg/l; 1,4-DHN: 1.21 microg/l) (por=0.623; p<0.0001). Also, levels of 1,2-DHN were significantly correlated with those of serum albumin adducts of l,2-naphthoquinone (rs=0.492, p=0.0004). These results indicate that 1,2- and 1,4-DHN are good biomarkers for assessment of naphthalene exposure in coke workers. Since the DHNs are precursors of the naphthoquinones, which have been implicated as toxic products of naphthalene metabolism, measurements of urinary DHNs may have toxicological significance.  相似文献   

9.
The electrophilic metabolites of the polyaromatic hydrocarbon naphthalene have been shown to bind covalently to proteins and covalent adduct formation correlates with the cytotoxic effects of the chemical in the respiratory system. Although 1,2-naphthalene epoxide, naphthalene diol epoxide, 1,2-naphthoquinone, and 1,4-napthoquinone have been identified as reactive metabolites of interest, the role of each metabolite in total covalent protein adduction and subsequent cytotoxicity remains to be established. To better understand the target residues associated with the reaction of these metabolites with proteins, mass spectrometry was used to identify adducted residues following (1) incubation of metabolites with actin and protein disulfide isomerase (PDI), and (2) activation of naphthalene in microsomal incubations containing supplemental actin or PDI. All four reactive metabolites bound to Cys, Lys or His residues in actin and PDI. Cys(17) of actin was the only residue adducted by all metabolites; there was substantial metabolite selectivity for the majority of adducted residues. Modifications of actin and PDI, following microsomal incubations containing (14)C-naphthalene, were detected readily by 2D gel electrophoresis and phosphor imaging. However, target modifications on tryptic peptides from these isolated proteins could not be readily detected by MALDI/TOF/TOF and only three modified peptides were detected using high resolution-selective ion monitoring (HR-SIM). All the reactive metabolites investigated have the potential to modify several residues in a single protein, but even in tissues with very high rates of naphthalene activation, the extent of modification was too low to allow unambiguous identification of a significant number of modified residues in the isolated proteins.  相似文献   

10.
The microsomal oxidation of 1,2-[14C]- and 1,4-[14C]dichlorobenzene (DICB) was investigated with special attention for possible differences in biotransformation that might contribute to the isomer-specific hepatotoxicity. Major metabolites of both isomers were dichlorophenols (2,5-DICP for 1,4-DICB and 2,3- and 3,4-DICP for 1,2-DICB, respectively) and dichlorohydroquinones. The formation of polar dihydrodiols appeared to be a major route for 1,2-DICB but not 1,4-DICB. Both the hepatotoxic 1,2-DICB and the non-hepatotoxic 1,4-DICB were oxidized to metabolites that covalently interacted with protein and only to a small extent with DNA. Protein binding could be inhibited by the addition of the reducing agent ascorbic acid with a concomitant increase in the formation of hydroquinones and catechols, indicating the involvement of reactive benzoquinone metabolites in protein binding. However, in the presence of ascorbic acid, a substantial amount of protein-bound metabolites of 1,2-DICB was still observed, in contrast to 1,4-DICB where binding was nearly completely inhibited. This latter effect was ascribed to the direct formation of reactive benzoquinone metabolites in a single P450-mediated oxidation of para-substituted dichlorophenols (such as 3,4-DICP) in the case of 1,2-DICB. In contrast, the major phenol isomer derived from 1,4-DICB (i.e. 2,5-DICP) is oxidized to its hydroquinone derivative, which needs prior oxidation in order to generate the reactive benzoquinone species. Residual protein binding in the presence of ascorbic acid could also indicate the involvement of reactive arene oxides in the protein binding of 1,2-DICB, but not of 1,4-DICB. However, MO computer calculations did not provide indications for differences in chemical reactivity and/or stability of the various arene oxide/oxepin tautomers that can be formed from either 1,2-DICB or 1,4-DICB. In conclusion, reactive intermediates in the secondary metabolism of 1,2-DICB lead to more covalent binding than those derived from 1,4-DICB, which correlates very well with their reported hepatotoxic potency.  相似文献   

11.
The hepatic microsomal metabolism of 1-naphthol, 1,2- and 1,4-naphthoquinone has been shown to generate active oxygen species by using electron spin resonance spin-trapping techniques. 1-Naphthol, in the presence of NADPH, and 1,2- and 1,4-naphthoquinone, with either NADH or NADPH, caused a stimulation in both the rate of microsomal oxygen consumption and the formation of superoxide spin adduct, 5,5-dimethyl-2-hydroxyperoxypyrrolidino-1-oxyl (DMPO-OOH). Superoxide dismutase, but not catalase, prevented the formation of this spin adduct, further supporting the suggestion that the superoxide free radical was the major oxy-radical formed during the microsomal metabolism of 1-naphthol and the naphthoquinones. These results are compatible with the suggestion that 1-naphthol may exert its toxicity to isolated hepatocytes and other cellular systems by metabolism to naphthoquinones followed by their redox cycling with concomittant generation of active oxygen species in particular superoxide free radicals.  相似文献   

12.
It has previously been shown that rats pre-treated with butylated hydroxyanisole (BHA), a well-known inducer of the enzyme DT-diaphorase, are protected against the toxic effects of 2-methyl-1,4-naphthoquinone but are made more susceptible to the harmful action of 2-hydroxy-1,4-naphthoquinone. In the present experiments, the effects of BHA have been compared with those of other inducers of DT-diaphorase. Rats were dosed with BHA, butylated hydroxytoluene (BHT), ethoxyquin (EQ), dimethyl fumarate (DMF) or disulfiram (DIS) and then challenged with a toxic dose of the naphthoquinones. All the inducers protected against the haemolytic anaemia induced by 2-methyl-1,4-naphthoquinone in rats, with BHA, BHT and EQ being somewhat more effective than DMF and DIS. A similar order of activity was recorded in the relative ability of these substances to increase hepatic activities of DT-diaphorase, consistent with a role for this enzyme in facilitating conjugation and excretion of this naphthoquinone. In contrast, all the compounds increased the haemolytic activity of 2-hydroxy-1,4-naphthoquinone. DMF and DIS were significantly more effective in this regard than BHA, BHT and EQ. DMF and DIS also caused a much greater increase in levels of DT-diaphorase in the intestine, suggesting that 2-hydroxy-1,4-naphthoquinone is activated by this enzyme in the gut. BHA, BHT and EQ had no effect on the nephrotoxicity of 2-hydroxy-1,4-naphthoquinone, but the severity of the renal lesions was decreased in rats pre-treated with DMF and DIS. The results of the present experiments show that modulation of tissue levels of DT-diaphorase may not only alter the severity of naphthoquinone toxicity in vivo, but may also change the relative toxicity of these substances to different target organs.  相似文献   

13.
It has previously been shown that rats pre-treated with butylated hydroxyanisole (BHA), a well-known inducer of the enzyme DT-diaphorase, are protected against the harmful effects of 2-methyl-1,4-naphthoquinone. This is consistent with a role for diaphorase in the detoxification of this quinone, but it is not known if increased tissue levels of this enzyme give protection against other naphthoquinone derivatives. In the present study, rats were dosed with BHA and then challenged with a toxic dose of 2-hydroxy-1,4-naphthoquinone, a substance that causes haemolytic anaemia and renal damage in vivo. Pre-treatment with BHA had no effect upon the nephrotoxicity of 2-hydroxy-1,4-naphthoquinone, but the severity of the haemolysis induced by this compound was increased in the animals given BHA. DT-Diaphorase is known to promote the redox cycling of 2-hydroxy-1,4-naphthoquinone in vitro, with concomitant formation of 'active oxygen' species. The results of the present experiment suggest that activation of 2-hydroxy-1,4-naphthoquinone by DT-diaphorase may also occur in vivo and show that increased tissue levels of DT-diaphorase are not always associated with naphthoquinone detoxification.  相似文献   

14.
DT-diaphorase catalysed the reduction of 1,4-naphthoquinones with hydroxy, methyl, methoxy and glutathionyl substituents at the expense of reducing equivalents from NADPH. The initial rates of quinone reduction did not correlate with either the half-wave reduction potential (E1/2) value (determined by h.p.l.c. with electrochemical detection against an Ag/AgCl reference electrode) or the partition coefficient of the quinones. After their reduction by DT-diaphorase the 1,4-naphthoquinone derivatives autoxidized at distinct rates, the extent of which was influenced by the nature of the substituents. Thus for the 1,4-naphthoquinone series the following order of rate of autoxidation was found: 5-hydroxy-1,4-naphthoquinone greater than 3-glutathionyl-1,4-naphthoquinone greater than 5-hydroxy-3-glutathionyl-1,4-naphthoquinone greater than 1,4-naphthoquinone greater than 2-hydroxy-1,4-naphthoquinone. For the 2-methyl-1,4-naphthoquinone (menadione) series the following order was observed: 5-hydroxy-2-methyl-1,4-naphthoquinone greater than 3-glutathionyl-5-hydroxy-2-methyl-1,4-naphthoquinone greater than 3-glutathionyl-2-methyl-1,4-naphthoquinone greater than 2-methyl-1,4-naphthoquinone greater than 3-hydroxy-2-methyl-1,4-naphthoquinone. The autoxidized naphthohydroquinone derivatives were re-reduced by DT-diaphorase, thus closing a cycle of enzymic reduction in equilibrium autoxidation. This was expressed as an excess of NADPH oxidized over the initial concentration of quinone present as well as H2O2 formation. These findings demonstrate that glutathionyl conjugates of 1,4-naphthoquinone and 2-methyl-1,4-naphthoquinone and those of their respective 5-hydroxy derivatives are able to act as substrates for DT-diaphorase and that they also autoxidize at rates higher than those for the unsubstituted parent compounds. These results are discussed in terms of the cellular role of DT-diaphorase in the reduction of hydroxy- or glutathionyl-substituted naphthoquinones as well as the further conjugation of these hydroquinones with glucuronide or sulphate within the cellular milieu, thereby facilitating their disposal from the cells.  相似文献   

15.
A new series of 1,2-naphthoquinone derivatives was synthesized by various synthetic methods and evaluated for their ability to inhibit protein tyrosine phosphatase 1B (PTP1B). 1,2-Naphthoquinone derivatives with substituent at R(4) position showed submicromolar inhibitory activity, and compound 24 demonstrated 10- to 60-fold selectivity against the tested phosphatases. Also, several 4-aryl-1,2-naphthoquinone derivatives with substituents at R(3), R(6), R(7), or/and R(8) showed submicromolar inhibitory activity and good plasma stability.  相似文献   

16.
The effect of hydroxy substitution on 1,4-naphthoquinone toxicity to cultured rat hepatocytes was studied. Toxicity of the quinones decreased in the series 5,8-dihydroxy-1,4-naphthoquinone greater than 5-hydroxy-1,4-naphthoquinone greater than 1,4-naphthoquinone greater than 2-hydroxy-1,4-naphthoquinone, and intracellular GSSG formation decreased in the order 5,8-dihydroxy-1,4-naphthoquinone greater than 5-hydroxy-1,4-naphthoquinone much greater than 1,4-naphthoquinone much greater than 2-hydroxy-1,4-naphthoquinone. The electrophilicity of the quinones decreased in the order 1,4-naphthoquinone much greater than 5-hydroxy-1,4-naphthoquinone greater than 5,8-dihydroxy-1,4-naphthoquinone much greater than 2-hydroxy-1,4-naphthoquinone. Treatment of the hepatocytes with BSO (buthionine sulfoximine) or BCNU (1,3-bis-2-chloroethyl-1-nitrosourea) increased 5-hydroxy-1, 4-naphthoquinone and 5,8-dihydroxy-1,4-naphthoquinone toxicity, whereas neither BSO nor BCNU largely affected 1,4-naphthoquinone and 2-hydroxy-1, 4-naphthoquinone toxicity. Dicumarol increased the toxicity of 1,4-naphthoquinone dramatically and somewhat the toxicity of 2-hydroxy-1,4- naphthoquinone, whereas 5-hydroxy-1,4-naphthoquinone and 5,8-dihydroxy-1,4-naphthoquinone toxicity increased only slightly. The toxicity of 5,8-dihydroxy-1,4-naphthoquinone decreased dramatically in reduced O2 concentration, whereas 1,4-naphthoquinone, 5-hydroxy-1,4-naphthoquinone, and 2-hydroxy-1,4-naphthoquinone toxicity was not largely affected. It was concluded that 5,8-dihydroxy-1,4-naphthoquinone toxicity is due to free radical formation, whereas the toxicity of 1,4-naphthoquinone and of 5-hydroxy-1,4-naphthoquinone also has an electrophilic addition component. The toxicity of 2-hydroxy-1,4-naphthoquinone could not be fully explained by either of these phenomena.  相似文献   

17.
The reactive industrial chemicals acrylamide (AA) and N-methylolacrylamide (MAA) are neurotoxic and carcinogenic in animals, MAA showing a lower potency than AA. The causative agent in AA-induced carcinogenesis is assumed to be the epoxy metabolite, glycidamide (GA), which in contrast to AA gives rise to stable adducts to DNA. The causative agent in MAA induced carcinogenesis is so far not studied. The two AAs were studied in mice and rats using analysis of hemoglobin (Hb) adducts as a measure of in vivo doses and the in vivo micronucleus (MN) assay as an end-point for chromosome damage. Male CBA mice were treated by intraperitoneal (i.p.) injection of three different doses and male Sprague-Dawley rats with one dose of each AA. Identical adducts were monitored from the two AAs [N-(2-carbamoylethyl)valine] and the respective epoxide metabolites [N-(2-carbamoyl-2-hydroxyethyl)valine]. Per unit of administered amount, AA gives rise to higher (three to six times) Hb adduct levels than MAA in mice and rats. Mice exhibit, compared with rats, higher in vivo doses of the epoxy metabolites, indicating that AAs were more efficiently metabolized in the mice. In mouse the two AAs induced dose-dependent increases in both Hb adduct level and MN frequency in peripheral erythrocytes. Per unit of administered dose MAA showed only half the potency for inducing micronuclei compared with AA, although the MN frequency per unit of in vivo dose of measured epoxy metabolite was three times higher for MAA than for AA. No increase in MN frequency was observed in rat bone marrow erythrocytes, after treatment with either AA. This is compatible with a lower sensitivity of the rat than of the mouse to the carcinogenic action of these compounds.  相似文献   

18.
The purpose of this study is to examine the differences in the induction of cytotoxic effects and poly(ADP-ribose) polymerase-1 activation in human MCF-7 breast cancer cells by quinonoid derivatives of naphthalene, including 1,2-naphthalenediol (NCAT), 1,4-naphthalenediol (NHQ), 1,2-naphthoquinone (1,2-NQ), and 1,4-naphthoquinone (1,4-NQ). Results from the cytotoxic response analyses in cells indicated that all naphthalene quinonoids induced cell death in MCF-7 cells at concentrations ranging from 0.1 to 100microM where NHQ and 1,4-NQ were more efficient than NCAT and 1,2-NQ in the induction of cell death. Results from Western blot analyses confirmed that treatment of cells with NCAT and NHQ resulted in up-regulation of p53 protein expression and a significant shift in bax/bcl2 ratio, suggesting the induction of p53-dependent apoptosis in MCF-7 cells. Additionally, we observed that all naphthalene quinonoids induced increases in reactive oxygen species (ROS) formation and glutathione (GSH) depletion in MCF-7 cells. The induction of ROS formation and GSH depletion in cells by naphthalene quinonoids decreases in the rank order 1,4-NQ>NHQ>1,2-NQ approximately equal to NCAT. Further investigation indicated that least-squares estimates of the overall rates of elimination (k(e)) of naphthalene quinonoids in MCF-7 cells decreased in the rank order 1,4-NQ>1,2-NQ>NHQ>NCAT. Values of k(e) were estimated to be between 0.280h(-1)(T(1/2)=151min) and 13.8h(-1)(T(1/2)=3.05min). These results provide evidence that the para-isomeric form of naphthalene quinonoids tend to induce acute production of ROS and alterations in intracellular redox status in cells, leading to the subsequent cell death. Further, all naphthalene quinonoids induced decreases in intracellular NAD(P)H and NAD(+) in MCF-7 cells at non-cytotoxic concentrations. The reduction of intracellular NAD(P)H in cells exposed to NCAT and 1,2-NQ was blocked by two types of poly(ADP-ribose) polymerase (PARP) inhibitors whereas PARP inhibitors did not prevent the reduction of NAD(P)H in cells exposed to NHQ and 1,4-NQ. Further investigation confirmed that increases in the number of DNA single-strand breaks were detected in MCF-7 cells exposed to NCAT and 1,2-NQ as measured by the single-cell gel electrophoresis (Comet) assay whereas NHQ and 1,4-NQ did not induce increases in the number of single-strand breaks in MCF-7 cells. Overall, results from our investigation suggest that while NHQ and 1,4-NQ are more efficient in the induction of cell death, NCAT and 1,2-NQ are prone to induce depletion of NAD(P)H and NAD(+) mediated by PARP-1 activation through formation of DNA single-strand breaks in human cultured cells.  相似文献   

19.
20.
The ability of naphthalene 1,2-oxide to diffuse across intact cellular membranes, the subsequent biotransformation of this epoxide and its potential to produce losses in cellular viability have been examined in incubations of isolated hepatocytes. Addition of 1R,2S- or 1S,2R-naphthalene oxide enantiomers (15, 30 and 60 microM) to isolated hepatocytes resulted in a rapid depletion of intracellular glutathione. Depletion of glutathione was concentration dependent and maximal at 5-15 min. Addition of either of the enantiomeric oxides at 60 microM resulted in the loss of more than 20 nmol glutathione/10(6) cells (1 ml cells); thus more than a third of the added epoxide was available for conjugation with intracellular glutathione. The time course and concentration dependence of glutathione depletion corresponded to the rapid, concentration-dependent formation of naphthalene oxide glutathione conjugates. The levels of glutathione adduct were highest 1 min after addition of naphthalene oxide and declined to 25% of this level after 30 min. Loss of glutathione conjugates from incubations correlated with the formation of N-acetylcysteine adducts. In contrast, the levels of glutathione adducts added exogenously to hepatocytes were relatively stable over a 120-min incubation suggesting that although further metabolism of naphthalene oxide glutathione adducts formed intracellularly is possible, extracellular glutathione adducts cannot penetrate the hepatocellular membrane. Small amounts of radiolabel from [3H]naphthalene 1,2-oxide were bound covalently to macromolecules in hepatocytes; the rate of this binding slowed rapidly after the first minute of incubation. Severe blebbing of the surface of the hepatocytes was noted in cells incubated for 30 min with 480 microM naphthalene oxide. Many of the cells were vacuolated at 60 min and progressed to frank necrosis with pyknotic nuclei and inability to exclude trypan blue. Cells incubated with 1-naphthol responded in a qualitatively similar fashion to those cells incubated with epoxide; however, hepatocytes incubated with 1-naphthol progressed to frank cellular necrosis at a slower rate. In hepatocytes partially depleted of glutathione by pretreatment with buthionine sulfoximine, addition of 1S,2R-naphthalene oxide at a rate of 1 nmol/min/10(6) cells resulted in significant losses in cell viability. In contrast, no losses in cell viability were observed with the enantiomer, 1R,2S-naphthalene oxide. Both epoxides produced similar losses in cellular glutathione levels.(ABSTRACT TRUNCATED AT 400 WORDS)  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