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 62 毫秒
1.
1. After administration of styrene (1) to rat, three sulphur-containing urinary metabolites were isolated in the molar ratio of III : IV : V equal to 65 : 34 : 1. 2. These metabolites were identified as N-acetyl-S-(1-phenyl-2-hydroxyethyl)cysteine (III), N-acetyl-S-(2-phenyl-2-hydroxyethyl)cysteine (IV) and N-acetyl-S-(phenacyl)cysteine (V). 3. After a single dose (250 mg/kg, 2.4 mmol/kg) styrene, the totally excreted mercapturic acids in the urine amounted to 10.7 +/- 1.0% (n = 5) of the dose.  相似文献   

2.
Two mercapturic acids, i. e., N-acetyl-S-(1-cyano-2-hydroxyethyl)-l-cysteine (CHEMA) and N-acetyl-S(2-hydroxyethyl)-l-cysteine (HEMA), were isolated from the urine of rats dosed with four successive doses of oxiranecarbonitrile (glycidonitrile, GN), 5 mg/kg, a reactive metabolic intermediate of acrylonitrile (AN). GC-MS analysis of methylated urine extracts from both AN- and GN-dosed rats showed another mercapturate which was identified as N-acetyl-S-(1-cyanoethenyl)-l-cysteine (1-CEMA) methyl ester using an authentic reference sample. The mass spectrum of this compound was very similar to that of a methylated metabolite of AN tentatively identified by Langvardt et al. (1980) as N-acetyl-3-carboxy-5-cyanothiazane (ACCT). In contrast, no ACCT was found in rats dosed with either GN or AN. Hence, there is no evidence for the formation of ACCT or its isomers in rats dosed with AN or GN. The methyl ester of 1-CEMA is formed artificially by dehydration of CHEMA methyl ester in the injector of the gas chromatograph.Details of the synthetic procedures and NMR-spectra are available from the authors on request  相似文献   

3.
1. Three hypotheses have been proposed for the mechanism of metabolism of alkylhalides to hydroxy-alkylmercapturic acids, two of which involve the intermediate step of dehydrohalogenation and formation of an epoxide. 2. After injection of (1-bromoethyl)benzene in rat, the only mercapturic acid appearing in the urine was N-acetyl-S-1-phenylethylcysteine. After injecting (2-bromoethyl)benzene in the rat only N-acetyl-S-2-phenylethylcysteine and N-acetyl-S-(2-phenyl-2-hydroxyethyl)cysteine were found in the urine. 3. Since the principal mercapturic acid formed from both styrene and styrene oxide could not be detected in the urine of rats receiving either 1- or 2-bromoethyl benzene, the intermediate formation of styrene or styrene oxide from the arylalkylhalides does not occur.  相似文献   

4.
1. The administration of beta-bromostyrene to the rat results in a fall in the level of hepatic glutathione. 2. Marmosets, rabbits and rats dosed with beta-bromostyrene excrete two mercapturic acids. One of these, N-acetyl-S-(2-hydroxy-2-phenyl-1-bromoethyl)-cysteine is readily converted into N-acetyl-S-(1-phenyl-2-bromo-2-ethenyl)-cysteine, the structure of which was established by mass spectrometry. 3. Mass spectrometric evidence suggests that the second mercapturic acid is N-acetyl-S-(1-hydroxy-2-phenylethyl)cysteine. 4. Mandelic acid was detected as a metabolite in all three species.  相似文献   

5.
The metabolism of 1,2-dibromo-1-phenylethane (DBPE) was studied in rats. Administration of DBPE orally, in doses of 0.25-1.25 mmol/kg (66-330 mg/kg), to male Wistar rats resulted in the excretion of a single mercapturic acid in urine. The methyl esters of three potential mercapturic acid metabolites were synthesized: N-acetyl-S-(2-oxo-2-phenylethyl)-L-cysteine methyl ester (O),N-acetyl-S-(2-hydroxy-1-phenylethyl)-L-cysteine methyl ester (I), and N-acetyl-S-(2-hydroxy-2-phenylethyl)-L-cysteine methyl ester (II). GC/MS analysis showed that the methyl ester of the excreted mercapturic acid was identical with II. Quantitative measurement of II in urine by GLC showed that, after 24 hr, excretion of the mercapturic acid was almost complete and amounted to 41% of the administered dose. At doses higher than 1.00 mmol/kg, the excretion no longer increased. Inhibition of the oxidative pathways by ip injection of 1-phenylimidazole resulted in an excretion decrease of about 40%. (Pre)treatment with diethyl maleate lowered the excretion of mercapturic acid by 30-60%. Glutathione conjugates synthesized from DBPE and styrene oxide were separated by HPLC. Both compounds can produce the same two pairs of diastereomers, viz. (R)- and (S)-(2-hydroxy-1-phenyl-ethyl)glutathione ((R)-1 and (S)-1), and (R)- and (S)-(2-hydroxy-2-phenylethyl)glutathione ((R)-2 and (S)-2). These could be separated in the order (R)-2, (R)-1, (S)-1, and (S)-2 within 20 min. This method was also applied to examine glutathione conjugates excreted in bile after DBPE administration.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

6.
Biotransformation of styrene and its toxic metabolite, phenyloxirane (1), in mice in vivo was studied. Mice were treated with single intraperitoneal doses of styrene (400 mg/kg of body weight), and with (R)-, (S)-, or racemic styrene oxide (150 mg/kg of body weight). Profiles of neutral and acidic metabolites were determined by GC/MS. Mandelic acid (3) and two mercapturic acids, N-acetyl-S-(2-hydroxy-2-phenylethyl)cysteine (5) and N-acetyl-S-(2-hydroxy-1-phenylethyl)cysteine (6), were found to be major urinary metabolites of both styrene and phenyloxirane. 1-Phenylethane-1,2-diol (2) was the main neutral metabolite. The rate of excretion of this metabolite, as determined by GC, was 5-10 times lower than that of mandelic acid. Several minor acidic metabolites were also identified. Among them, novel phenolic metabolites, namely, 2-(4-hydroxyphenyl)ethanol (7), (4-hydroxyphenyl)acetic acid (11), and two isomeric hydroxymandelic acids (12), are of toxicological significance. Main stereogenic metabolites were isolated as methyl esters from extracts of pooled acidified urine treated with diazomethane. The mandelic acid that was obtained was converted to diastereomeric Mosher's derivatives prior to analysis by NMR. Mercapturic acids were analyzed directly by (13)C NMR. Pure enantiomers of 1 were metabolized predominantly but not exclusively to corresponding enantiomers of 3. Styrene yielded predominantly (S)-mandelic acid. Fractions of mercapturic acids 5 and 6 isolated from urine amounted to 12-15% of the dose for all compounds that were administered. Conversion to mercapturic acids was highly regio- and stereoselective, yielding predominantly regioisomer 5. Styrene, as compared to racemic phenyloxirane, yielded slightly more diastereomers arising from (S)-1 than from (R)-1. These data can be explained by formation of a moderate excess of the less mutagenic (S)-1 in the metabolic activation of styrene in mice in vivo.  相似文献   

7.
We have developed a relatively simple and reproducible HPLC procedure for the determination of urinary metabolites of styrene. Urine samples (pH 2) are extracted with ethyl acetate, the organic layer is evaporated to dryness, and the residues are dissolved in water-methanol (1:1). Samples are analyzed by HPLC with a C18 reversed-phase column and a water (pH 6, 5mM tetrabutylammonium dihydrogen phosphate)-acetonitrile gradient and ultraviolet detection at 225 nm. Using this method, it is possible to determine simultaneously mandelic and phenylglyoxylic acids, N-acetyl-S-(1-phenyl-2-hydroxyethyl)-L-cysteine (M1), and N-acetyl-S-(2-phenyl-2-hydroxyethyl)-L-cysteine (M2). The internal standard is p-hydroxybenzoic acid. Additional validation data are obtained by analysis of urine samples obtained from rats treated with a wide range of styrene doses.  相似文献   

8.
Two workers were accidentally exposed to unusually high styrene concentrations (>1000 ppm) for about 30 min. In addition to the main styrene metabolites, mandelic acid (MA) and phenylglyoxylic acid (PGA), other minor metabolites, including specific mercapturic acids, (R,R)- and (S,R)-N-acetyl-S-(1-phenyl-2-hydroxyethyl)-L-cysteine [(R,R)-M1 and (S,R)-M1] and (R,R)- and (S,R)-N-acetyl-S-(2-phenyl-2-hydroxyethyl)-L-cysteine [(R,R)-M2 and (S,R)-M2], 4-vinylphenol-glucuronide and -sulfate, and phenylglycine, were determined by Liquid Chromatography Electrospray Tandem Mass Spectrometry (LC-ESI-MS/MS) in urine samples collected 12, 24, 36, 48, 75 and 99 h after the episode. The genotypes of microsomal epoxide hydrolase, glutathione-S-transferases M1-1 (GSTM1), T1-1 (GSTT1) and P1-1 (GSTP1) were characterized by PCR-based methods. The two subjects showed similar peak levels of MA and PGA, as well as 4-vinylphenol conjugates, whereas mercapturic acids were five times higher in the subject bearing the GSTM1pos than in the GSTM1null subject. Also, relative proportions of diasteroisomers of mercapturic acids were influenced by the GSTM1 polymorphism.  相似文献   

9.
In mammalian species, including man, N-acetyl-S-(2-hydroxyethyl)-L-cysteine (2-HEMA) is a common urinary metabolite of a large number of structurally different xenobiotic chemicals. It is a common urinary end product of glutathione pathway metabolism of a variety of chemicals possessing electrophilic properties and, in most cases, also a genotoxic potential. Five different chemically reactive intermediates, with different electrophilic properties, may be involved in the formation of 2-HEMA. An inventory of chemicals known to lead to the formation of 2-HEMA, or based on their chemical structure expected to do so, is presented. Furthermore, an attempt is made to evaluate the possibilities and limitations in terms of the potential use of urinary 2-HEMA as a tool in biomonitoring studies. Two other related, sulfur-containing urinary metabolites, i. e. N-acetyl-(S-carboxymethyl)-L-cysteine and thio-diacetic acid, are proposed as possible alternatives to urinary 2-HEMA. It is suggested that 2-HEMA might be seen as a potentially useful and sensitive signal parameter for the assessment of exposure of animals and man to a variety of electrophilic and therefore potentially toxic xenobiotic chemicals.  相似文献   

10.
1. Biotransformation of 1,3-diethenylbenzene (1) in rat gave four major metabolites, namely, 3-ethenylphenylglyoxylic acid (2), 3-ethenylmandelic acid (3), N-acetyl-S-[2-(3-ethenylphenyl)-2-hydroxyethyl]-L-cysteine (4) and N-acetyl-S-[1-(3-ethenylphenyl)-2-hydroxyethyl]-L-cysteine (5) were isolated from urine and identified by n.m.r. and mass spectrometry. 2. Four minor metabolites, 3-ethenylbenzoic acid (6), 3-ethenylphenylacetic acid (7), 3-ethenylbenzoylglycine (8) and 2-(3-ethenylphenyl)ethanol (9) were identified by g.l.c.-mass spectrometric analysis of urine extract derivatized in two different ways. 3. All identified metabolites are derived from 3-ethenylphenyloxirane (10), a reactive metabolic intermediate. No product of any metabolic transformation of second ethenyl group has been identified. However, several minor unidentified metabolites were detected by g.l.c.-mass spectrometry. 4. Total thioether excretion in 24 h urine after a single i.p. dose of 1 amounted to 28.3 +/- 3.5 dose (mean +/- SD). No significant differences in the thioether fraction were observed in the dose range 100-300 mg/kg. 5. Thioether metabolites consisted mainly of mercapturic acids 4 and 5. The ratio of metabolites 5 to 4 was 62:38. Each mercapturic acid consisted of two diastereomers. Their ratio, as determined by quantitative 13C-n.m.r. measurement was 95:5 and 79:21 for mercapturic acids 4 and 5, respectively.  相似文献   

11.
After administration of acrylonitrile, crotononitrile and cinnamonitrile to rats, two types of mercapturic acids were isolated from urine and identified by mass and NMR spectroscopy as N-acetyl-S-(2-cyanoethyl)-L-cysteine (I) and N-acetyl-S-(2-hydroxyethyl)-L-cysteine (II) (methyl-substituted in the case of crotonitrile and phenyl-substituted in the case of cinnamonitrile). After pretreatment of rats with the cytochrome P-450 inhibitor 1-phenylimidazole, no trace of mercapturic acid II was found, whereas a higher amount of mercapturic acid I was excreted. It is suggested that the first type of products result from direct addition of glutathione, whereas the second group of metabolites (II), in which the cyano group has been replaced by a hydroxyl group, are formed via an intermediate epoxide. Substituents on the double bond had a considerable influence on the ratio of the two mercapturic acids formed, and thus presumably on the amount metabolized via an oxidative process: the ratio of the cyano (I) to hydroxy (II) mercapturic acid was 72:28 for AN; introduction of a methyl or a phenyl group resulted in ratios of 91:9 and 98:2, respectively.  相似文献   

12.
The role of polymorphic xenobiotic-metabolizing enzymes in the interindividual variability of phenylhydroxyethyl mercapturic acids (PHEMAs) was investigated in 56 styrene-exposed workers. Ambient monitoring was carried out using passive personal samplers (geometric mean, 157 mg/m3 8-h time-weighted average; geometric standard deviation, 2.90). Biomonitoring was based on mandelic acid and phenylglyoxylic acid in urine spot samples collected at the end of the work shift ("end-of-shift") and prior to the subsequent shift ("next morning"). Four PHEMA diastereoisomers, namely (R,R)-M1, (S,R)-M1, (S,R)-M2, and (R,R)-M2, were determined by HPLC/tandem mass spectrometry. The genotypes of glutathione S-transferases M1-1 (GSTM1), T1-1 (GSTT1) and P1-1 (GSTP1), and microsomal epoxide hydrolase (EPHX) were characterized by PCR-based methods. Workers bearing the GSTM1pos genotype showed PHEMA concentrations five and six times higher (in end-of-shift and next-morning samples, respectively) as compared to GSTM1null people. In GSTM1pos subjects, (R,R)-M1 was the main mercapturate affected by the GSTM1 status, accounting for 54 and 68% of total PHEMAs in end-of-shift and next-morning samples, respectively. Compared to GSTM1null, GSTM1pos subjects excreted more -M1 than -M2 and more (R,R)-M1 and (S,R)-M2 than (S,R)-M1 and (R,R)-M2 diastereoisomers. Thus, GSTM1-1 is the main isoenzyme catalyzing GSH-conjugation of styrene-7,8-oxide in humans and it seems to act in a regio- and stereoselective way. PHEMAs cannot be recommended as biomarkers of exposure to styrene, unless the GSTM1 genotype is considered in data interpretation. Their role as biomarkers of susceptibility deserves further studies.  相似文献   

13.
1. Biotransformation of 1,4-diethenylbenzene (1) in rat was studied. Six urinary metabolites, namely, N-acetyl-S-[2-(4-ethenylphenyl)-2-hydroxyethyl]-L-cysteine (3), N-acetyl-S-[1-(4-ethenylphenyl)-2-hydroxyethyl]-L-cysteine (4), N-acetyl-S-[1-(4-formylphenyl)-2-hydroxyethyl]-L-cysteine (5), 1-(4-ethenylphenyl)ethane-1,2-diol (6), 4-ethenylbenzoic acid (9) and 4-ethenylbenzoyl-glycine (12) were isolated and identified by n.m.r. and mass spectrometry. 2. G.l.c.-mass spectral analysis of the methylated urine extract allowed the identification of four other metabolites, as 4-ethenylphenylacetic acid (11), 4-ethenylphenylacetylglycine (13), 4-ethenylmandelic acid (7), and 4-ethenylphenylglyoxylic acid (8). 3. The structures of the identified metabolites indicate that the main reactive intermediate in the metabolism of 1 is 4-ethenylphenyloxirane (2). The first step in the biotransformation of 1, formation of an oxirane, is very similar to the metabolic activation of styrene. However, subsequent steps lead not only to analogues of styrene metabolites but also to oxidation of the second ethenyl group leading to compound(s) which may contribute to the toxicity of 1, e.g. to the aldehyde 5. 4. Rats dosed with a single i.p. dose of 1 excreted nearly 5.6% of the dose as the glycine conjugate 12, irrespective of the dose. 5. In contrast, the total thioether fraction decreased significantly with increasing dose, being 23 +/- 3, 17 +/- 5 and 12 +/- 1% of dose at 100, 200 and 300 mg/kg, respectively (mean +/- SD).  相似文献   

14.
We investigated the human metabolism of AA to the mercapturic acids N-acetyl-S-(2-carbamoylethyl)-l-cysteine (AAMA) and N-(R/S)-acetyl-S-(2-carbamoyl-2-hydroxyethyl)-l-cysteine (GAMA) which are derived from AA itself and from its oxidative genotoxic metabolite glycidamide (GA), respectively. A healthy male volunteer received a single dose of about 1 mg deuterium-labelled acrylamide (d3-AA), representing 13 μg/kg body weight, in drinking water. Urine samples before dosing and within 46 h after the dose were analysed for d3-AAMA and d3-GAMA by LC-ESI-MS/MS. A first phase of increase in urinary concentration was found to last 18 h with a broad plateau between 8 and 18 h for AAMA, and 22 h for GAMA. Elimination half-lives of both AAMA and GAMA were estimated to be approximately 3.5 h for the first phase and more than 10 h up to few days for the second phase. Total recovery in urine after 24 h was about 51% as the sum of AAMA and GAMA and hereby well in accordance with former studies in rats. After 2 days AAMA, accounting for altogether 52% of the total AA dose, was the major metabolite of AA in humans. GAMA, accounting for 5%, appeared as a minor metabolite of AA. In humans we found a urinary ratio of 0.1 for GAMA/AAMA compared to previously reported values of 0.2 for rats and 0.5 for mice. Therefore, the metabolic fate of AA in humans was more similar to that in rats than in mice as already demonstrated in terms of the haemoglobin adducts. Consequently a genotoxic potency of AA mediated by GA could be supposed to be comparable in rats and humans.  相似文献   

15.
1. The metabolism of several dihalopropanols has been studied in the rat. Irrespective of their structure, each compound produced the same two mercapturic acids, excreted as urinary metabolites. 2. The mercapturic acid metabolites of the dihalopropanols were identified as N-acetyl-S-(2,3-dihydroxypropyl)cysteine and N,N'-bis-acetyl-S,S'-(1,3-bis-cysteinyl)propan-2-ol. Depending on the halogen present, each dihalopropanol produced beta-chlorolactate or beta-bromolactate as oxidative metabolites. 3. From the metabolic pathway of these compounds, it is inferred that an epoxide is an intermediate in their metabolism. 4. The metabolism of 2-chloropropane-1,3-diol has been investigated in the rat and the isolation of one mercapturic acid, N-acetyl-S-(2,3-dihydroxypropyl)cysteine, confirms that an epoxide intermediate is involved.  相似文献   

16.
The present human intervention study investigated the relation between the intake of acrylamide (AA) in diets with minimized, low, and high AA contents and the levels of urinary exposure biomarkers. As biomarkers, the mercapturic acids, N-acetyl-S-(carbamoylethyl)-l-cysteine (AAMA), and N-acetyl-S-(1-carbamoyl-2-hydroxyethyl)-l-cysteine (GAMA) were monitored. The study was performed with 14 healthy male volunteers over a period of 9 days, under controlled conditions excluding any inadvertent AA exposure. Dietary exposure to AA was measured by determining AA contents in duplicates of all meals consumed by the volunteers. The study design included an initial washout period of 3 days on AA-minimized diet, resulting in dietary AA exposure not exceeding 41 ng/kg bw/d. Identical washout periods of 2 days each followed the AA exposure days (day 4, low exposure, and day 7, high exposure). At the respective AA intake days, volunteers ingested 0.6–0.8 (low exposure) or 1.3–1.8 (high exposure) μg AA/kg bw/d with their food. Both low and high AA intakes resulted in an AAMA output within 72 h corresponding to 58 % of the respective AA intake. At the end of the initial 3-day washout period, an AAMA baseline level of 93 ± 31 nmol/d was recorded, suggestive for an assumed net AA baseline exposure level of 0.2–0.3 μg AA/kg bw/d.  相似文献   

17.
N-Acetyl-L-cysteine was reacted with 2-(2-, 3-, or 4-methylphenyl)-oxiranes to give mixtures of the two possible regio isomers N-acetyl-S-[1-(2-, 3-, or 4-methylphenyl)-2-hydroxyethyl]-L-cysteine and N-acetyl-S-[2-(2-, 3-, or 4-methylphenyl)-2-hydroxyethyl]-L-cysteine, respectively. These were isolated in pure form by h.p.l.c.. The diastereomers were characterized by n.m.r. and mass spectrometry. The 2-, 3- and 4-ethenyl-methylbenzenes and the 2-(2-, 3-, and 4-methylphenyl)-oxiranes were injected i.p. into rats. G.l.c.-mass spectrometry showed similar patterns of acidic metabolites in the urine. Comparison with authentic mass spectra showed that the N-acetyl-S-[1-(methylphenyl)-2-hydroxyethyl]-L-cysteines accounted for over 80% of the mercapturic acids.  相似文献   

18.
1-Cyano-3,4-epithiobutane (CEB), a naturally occurring nitrile derived from cruciferous plants, causes nephrotoxicity in male Fischer 344 rats. Nephrotoxicity induced by CEB is dependent on glutathione (GSH) conjugation and bioactivation. Conjugation with GSH and subsequent metabolism leads to the formation of specific urinary metabolites. The objectives of the present study were to identify CEB-derived urinary metabolites and quantify urinary non-protein thiols and thioethers in male Fischer 344 rats. Animals received 125 mg kg(-1) of CEB alone or following pretreatment with one of three selective inhibitors of GSH metabolism: acivicin, probenecid or aminooxyacetic acid. Total non-protein urinary thiol and urinary thioether concentrations were elevated in all treated groups at 12 and 24 h; however, elevations in non-protein thiols were not significantly greater in rats administered CEB alone as compared to negative controls. A single predominant urinary metabolite was identified as the CEB-derived mercapturic acid N-acetyl-S-(4-cyano-thio-1-butyl)-cysteine. Evidence for other CEB-derived metabolites was also demonstrated. These findings represent the identification of a unique compound and provide further evidence for the importance of GSH conjugation as a significant pathway in CEB metabolism.  相似文献   

19.
The oxidative metabolism of 1-bromopropane in the rat   总被引:3,自引:0,他引:3  
1. The metabolism of 1-bromopropane in the rat has been re-investigated. The previously known metabolites have been isolated and confirmed as the three mercapturic acids N-acetyl-S-propyl cysteine, N-acetyl-S-propyl cysteine-S-oxide and N-acetyl-S-(2-hydroxypropyl)cysteine. 2. Three further metabolites have been isolated from the urine of rats treated with 4-bromopropane. These have been identified as 3-bromopropionic acid and the mercapturic acids N-acetyl-S-(3-hydroxypropyl)cysteine and N-acetyl-S-(2-carboxyethyl)cysteine. 3. The metabolites of 3-bromopropanol and 3-chloropropanol in the rat have been shown to be the mercapturic acids N-acetyl-S-(3-hydroxypropyl)cysteine and N-acetyl-S-(2-carboxyethyl)cysteine and the corresponding 2-carboxyethyl halide. 4. Studies with 1-bromopropane and the 3-halopropanols in vitro indicate that oxidation of C3 and C2 of 1-bromopropane occurs before conjugation of the alkyl group with glutathione. The implications of these studies are discussed in relation to the mechanism of the biosynthesis of the S-(2-hydroxyalkyl)mercapturic acid metabolites derived from the alkyl halides.  相似文献   

20.
1,2-Dibromopropane was administered orally in doses of 50-350 mg/kg to male Wistar rats. Four mercapturic acids were identified in urine by GC/MS, viz. N-acetyl-S-(2-oxopropyl)-L-cysteine (I), N-acetyl-S-(2-hydroxypropyl)-L-cysteine (II), N-acetyl-S-(1-carboxyethyl)-L-cysteine (III), and N-acetyl-S-(2-bromo-2-propenyl)-L-cysteine (IV). Mercapturic acid IV was a minor metabolite which could only be measured at doses of 200 mg/kg or higher. In 24 hr, urinary excretion of mercapturic acids amounted to about 36% of the dose (11% I, 21% II, 4% III, 0.2% IV). No dose dependency was found up to the highest dose. A unified scheme is proposed for the metabolism of 1,2-dibromopropane in the rat, which accounts for the identified mercapturic acids. The role of direct glutathione conjugation in the route leading to the major metabolite II, presumably involving thiiranium ion formation, is discussed. This route probably is biologically not very important because of the absence of detectable activity of 1,2-dibromopropane toward glutathione S-transferases in vitro, the very low mutagenicity of 1,2-dibromopropane, and the high mutagenic activity of N-acetyl-S-(2-bromopropyl)-L-cysteine methyl ester which was studied as a model compound for direct 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