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 468 毫秒
1.
Adducts of benzo[a]pyrene-diolepoxide (BPDE) with blood nucleophiles have been used as biomarkers of exposure to polycyclic aromatic hydrocarbons (PAHs). The most popular such assay is a competitive enzyme-linked immunosorbent assay (ELISA) that employs monoclonal antibody 8E11 to detect benzo[a]pyrene tetrols following hydrolysis of BPDE adducts from lymphocyte DNA or human serum albumin (HSA). Here we used 8E11 as the capture antibody in a sandwich ELISA to detect BPDE-HSA adducts directly in 1-mg samples of HSA or 20 μl of serum/plasma. The assay employs an anti-HSA antibody for detection, and this is amplified by an avidin/biotinylated horseradish peroxidase complex. The sandwich ELISA has advantages of specificity and simplicity and is approximately 10 times more sensitive than the competitive ELISA. To validate the assay, HSA samples were assayed from three populations with known high PAH exposures (coke oven workers), medium PAH exposures (steel factory control workers), and low PAH exposures (volunteer subjects) (n = 30). The respective geometric mean levels of BPDE-HSA adducts—67.8, 14.7, and 1.93 ng/mg HSA (1010, 220, and 28.9 fmol BPDE equiv/mg HSA)—were significantly different (P < 0.05). The sandwich ELISA will be useful for screening PAH exposures in large epidemiologic studies and can be extended to other adducts for which capture antibodies are available.  相似文献   

2.
Activation of aryl hydrocarbon receptor (AhR) by 30 polycyclic aromatic hydrocarbons (PAHs) was determined in the chemical-activated luciferase expression (CALUX) assay, using two exposure times (6 and 24h), in order to reflect the metabolization of PAHs. AhR-inducing potencies of PAHs were expressed as induction equivalency factors (IEFs) relative to benzo[a]pyrene and 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD). In 24h exposure assay, the highest IEFs were found for benzo[k]fluoranthene, dibenzo[a,h]anthracene and dibenzo[a,k]fluoranthene (approximately three orders of magnitude lower than TCDD) followed by dibenzo[a,j]anthracene, benzo[j]fluoranthene, indeno[1,2,3-cd]pyrene, and naphtho[2,3-a]pyrene. The 6h exposure to PAHs led to a significantly higher AhR-mediated activity than the 24h exposure (generally by two orders of magnitude), probably due to the high rate of PAH metabolism. The strongest AhR inducers showed IEFs approaching that of TCDD. Several PAHs, including some strong mutagens, such as dibenzo[a,l]pyrene, cyclopenta[cd]pyrene, and benzo[a]perylene, elicited only partial agonist activity. Calculation of IEFs based on EC25 values and/or 6h exposure data is suggested as an alternative approach to estimation of toxic potencies of PAHs with high metabolic rates and/or the weak AhR agonists. The IEFs, together with the recently reported relative mutagenic potencies of PAHs [Mutat. Res. 371 (1996) 123; Mutat. Res. 446 (1999) 1] were combined with data on concentrations of PAHs in extracts of model environmental samples (river sediments) to calculate AhR-mediated induction equivalents and mutagenic equivalents. The highest AhR-mediated induction equivalents were found for benzo[k]fluoranthene and benzo[j]fluoranthene, followed by indeno[1,2,3-cd]pyrene, dibenzo[a,h]anthracene, benzo[a]pyrene, dibenzo[a,j]anthracene, chrysene, and benzo[b]fluoranthene. High mutagenic equivalents in the river sediments were found for benzo[a]pyrene, dibenzo[a,e]pyrene, and naphtho[2,3-a]pyrene and to a lesser extent also for benzo[a]anthracene, benzo[b]fluoranthene, indeno[1,2,3-cd]pyrene, benzo[j]fluoranthene, dibenzo[a,e]fluoranthene and dibenzo[a,i]pyrene. These data illustrate that AhR-mediated activity of PAHs, including the highly mutagenic compounds, occurring in the environment but not routinely monitored, could significantly contribute to their adverse effects.  相似文献   

3.
A number of highly toxic environmental pollutants including certain polychlorinated dibenzo-p-dioxins (PCDD), polychlorinated dibenzofurans (PCDF), and 'dioxin-like' polychlorinated biphenyls (PCB) are among the most potent agonists of the aryl hydrocarbon receptor (AHR). Induction of cytochrome P4501A1 (CYP1A1) in mammalian cell culture is widely used as a functional parameter for AHR activation providing an estimate for 'dioxin-like' inducing equivalents in extracts from environmental samples. Since a number of polycyclic aromatic hydrocarbons (PAHs) also act as AHR-agonists, the CYP1A1-inducing potencies, measured as induction of 7-ethoxyresorufin O-deethylase (EROD) activity in rat hepatocyte cultures were analyzed for 16 PAHs frequently present in environmental samples. Among these, seven PAHs including benzo[a]pyrene were relatively potent inducers allowing the determination of Induction Equivalency Factors (IEF). For three PAHs including benzo[k]fluoranthene which acted as weak inducers, IEFs were estimated, while six PAHs including acenaphthylene were classified as inactive. Based on different efficacies the concentration-response characteristics of CYP1A1 induction were analyzed in more detail for benzo[a]pyrene, benzo[k]fluoranthene, and acenaphthylene. Benzo[k]fluoranthene was markedly less effective than benzo[a]pyrene as inducer of EROD activity but even more effective than benzo[a]pyrene as inducer of CYP1A1 protein and mRNA. Acenaphthylene was highly more effective on the level of mRNA than on the levels of protein or EROD activity. Further analysis revealed that the low efficacy of acenaphthylene as inducer of CYP1A1 protein and EROD activity is due to its marked cytotoxicity while no clear-cut explanation was found for the differences in efficacy between benzo[k]fluoranthene and benzo[a]pyrene. The EROD-inducing potency of a mixture of 16 PAH was about 2-fold higher than that calculated on the basis of IEFs of the individual constituents of the mixture.  相似文献   

4.
An ombrotrophic peat core was collected in 2005 from Etang de la Gruère, Jura Mountains, Switzerland. The concentrations of nine among the U.S. Environmental Protection Agency priority polycyclic aromatic hydrocarbons (PAHs) (i.e., acenaphthene, phenanthrene, fluorene, pyrene, fluoranthene, benzo[jbk]fluoranthene, benzo[a]pyrene, benzo[ghi]perylene, and indeno[1,2,3-cd]pyrene) were determined in both bulk peat and corresponding humic acids (HA) samples by gas chromatography equipped with a mass spectrometry detector (GC-MS). The maximum PAHs concentrations in peat (around 1,250 μg Σ PAHs kg?1 dry matter) were found at 28–30 cm of depth, which correspond to ca. 1920–1930, when coal inputs to Switzerland reached their maximum level. Amongst the nine PAHs analyzed in the peat samples, pyrene (Pyr) was the predominant species, accounting for ca. 20–100% of the total PAHs throughout the profile. In the HA fraction, that represents 24.7% (average value) of the bulk peat, only phenanthrene (Phe), and sporadically Pyr and fluoranthene (Fth), were detected. In particular, HA showed Phe concentrations that were ten–150 times higher than corresponding bulk peat samples, thus suggesting its preservation against biodegradation due to the incorporation into HA molecules.  相似文献   

5.
This study was done to determine the concentration of PAHs in urban soil of Delhi (India). Surface top soil (up to 10 cm depth) samples were collected from four different sampling sites including industrial, roadside, residential, and agricultural areas of Delhi and 16 USEPA priority polycyclic aromatic hydrocarbons (PAHs) were evaluated. Total PAH concentrations at industrial, roadside, residential, and agricultural sites were 11.46 ± 8.39, 6.96 ± 4.82, 2.12 ± 1.12, and 1.55 ± 1.07 mg/kg (dry weight), respectively, with 3–7 times greater concentrations in industrial and roadside soils than that in residential and agricultural soils. The PAH pattern was dominated by 4- and 5-ring PAHs (contributing >50% to the total PAHs) at industrial and roadside sites with greater concentration of fluoranthene, chrysene, benzo[b]fluoranthene, benzo[k]fluoranthene, benzo[a]anthracene, benzo[ghi]perylene, and pyrene, whereas, residential and agricultural sites showed a predominance of low molecular weight 2- and 3-ring PAHs (fluoranthene, acenaphthene, naphthalene, chrysene, and anthracene). Isomeric pair ratios suggested biomass combustion and fossil fuel emissions as the main sources of PAHs. The toxic equivalency factors (TEFs) showed that carcinogenic potency (benzo[a]pyrene-equivalent concentration (B[a]Peq) of PAH load in industrial and roadside soils was ~10 and ~6 times greater than the agricultural soil.  相似文献   

6.
Genotoxicity of polycyclic aromatic hydrocarbons in Escherichia coli PQ37.   总被引:2,自引:0,他引:2  
In the present investigation, 32 polycyclic aromatic hydrocarbons (PAHs) were tested for genotoxicity in E. coli PQ37 using the standard tube assay of the SOS chromotest. PAHs such as benzo[ghi]fluoranthene, benzo[j]fluoranthene, benzo[a]pyrene, chrysene, dibenzo[a,l]pyrene, fluoranthene and triphenylene exhibited high genotoxicity when incubated in the presence of an exogenous metabolic activation mixture. The results were compared to those obtained with the Salmonella/microsome test.  相似文献   

7.
《Luminescence》2003,18(2):97-102
Benzo[a]pyrene causes cancer at cellular level and is widely present in the environment. Conventional spectroscopic methods for analysis of this compound need a pre‐separation procedure due to severe spectral overlap from other polycyclic aromatic hydrocarbons. We report a simple method that avoids spectral overlap of benzo[a]pyrene from other impurities or polycyclic aromatic hydrocarbons (PAHs), thus it can easily identify benzo[a]pyrene in a complex PAH mixture. The method could easily identify benzo[a]pyrene in an 18‐component PAH mixture. Calibration plots in methanol solution and in micellar media show a good linearity (R > 0.9997) in the benzo[a]pyrene concentration range generally found in the environment. The method gives a detection limit of 1.52 × 10?9 mol/L in CTAB micellar medium and 2.55 × 10?9 mol/L in methanol solution. The proposed method is selective, sensitive and fast. The fluorescence response of benzo[a]pyrene is found to be a potential candidate to sense the critical micellar concentration (CMC) of CTAB micelles. Copyright © 2003 John Wiley & Sons, Ltd.  相似文献   

8.
《Process Biochemistry》2014,49(10):1723-1732
The removal and transformation of seven high molecular weight polycyclic aromatic hydrocarbons (PAHs), namely benz[a]anthracene, benzo[b]fluoranthene, benzo[k]fluoranthene, benzo[a]pyrene, dibenzo[a,h]anthracene, indeno[1,2,3-c,d]pyrene and benzo[g,h,i]perylene, by a freshwater microalga Selenastrum capricornutum under gold and white light irradiation was studied. The two light sources did not result in significant differences in the biodegradation of the selected PAHs in live algal cells, but white light was more effective in promoting photodegradation than was gold light in dead cells. The removal efficiency of seven PAHs, as well as the difference between live and dead microalgal cells, was PAH compound-dependent. Benz[a]anthracene and benzo[a]pyrene were highly transformed in live and dead algal cells, and dead cells displayed greater transformation levels than live cells. Further investigation comparing the transformation of single PAH compound, benzo[a]pyrene, by S. capricornutum and another green microalgal species, Chlorella sp., demonstrated that the transformation in dead cells was similar, indicating the process was algal-species independent. Dead algal cells most likely acted as a photosensitizer and accelerated the photodegradation of PAHs.  相似文献   

9.
The genotoxicity of 15 polycyclic aromatic hydrocarbons was determined with the alkaline version of the comet assay employing V79 lung fibroblasts of the Chinese hamster as target cells. These cells lack the enzymes necessary to convert PAHs to DNA-binding metabolites. Surprisingly, 11 PAHs, i.e., benzo[a]pyrene (BaP), benz[a]anthracene, 7,12-dimethylbenz[a]anthracene, 3-methylcholanthrene, fluoranthene, anthanthrene, 11H-benzo[b]fluorene, dibenz[a,h]anthracene, pyrene, benzo[ghi]perylene and benzo[e]pyrene caused DNA strand breaks even without external metabolic activation, while naphthalene, anthracene, phenanthrene and naphthacene were inactive. When the comet assay was performed in the dark or when yellow fluorescent lamps were used for illumination the DNA-damaging effect of the 11 PAHs disappeared. White fluorescent lamps exhibit emission maxima at 334.1, 365.0, 404.7, and 435.8 nm representing spectral lines of mercury. In the case of yellow fluorescent lamps these emissions were absent. Obviously, under standard laboratory illumination many PAHs are photo-activated, resulting in DNA-damaging species. This feature of PAHs should be taken into account when these compounds are employed for the initiation of skin cancer. The genotoxicity of BaP that is metabolically activated in V79 cells stably expressing human cytochrome P450-dependent monooxygenase (CYP1A1) as well as human epoxide hydrolase (V79-hCYP1A1-mEH) could not be detected with the comet assay performed under yellow light. Likewise the DNA-damaging effect of r-7,t-8-dihydroxy-t-9,10-epoxy-7,8,9,10-tetrahydrobenzo[a]pyrene (anti-BaPDE) observed with the comet assay was only weak. However, upon inhibition of nucleotide excision repair (NER), which is responsible for the removal of stable DNA adducts caused by anti-BaPDE, the tail moment rose 3.4-fold in the case of BaP and 12.9-fold in the case of anti-BaPDE. These results indicate that the genotoxicity of BaP and probably of other compounds producing stable DNA adducts are reliably detected with the comet assay only when NER is inhibited.  相似文献   

10.
Laccases produced by white rot fungi are capable of rapidly oxidizing benzo[a]pyrene. We hypothesize that the polycyclic aromatic hydrocarbon (PAH)-degrading bacteria producing laccase can enhance the degree of benzo[a]pyrene mineralization. However, fungal laccases are glycoproteins which cannot be glycosylated in bacteria, and there is no evidence to show that bacterial laccases can oxidize benzo[a]pyrene. In this study, the in vitro oxidation of PAHs by crude preparations of the bacterial laccase, CueO, from Escherichia coli was investigated. The results revealed that the crude CueO catalyzed the oxidation of anthracene and benzo[a]pyrene in the same way as the fungal laccase from Trametes versicolor, but showed specific characteristics such as thermostability and copper dependence. In the presence of 2,2'-azino-bis-(3-ethylbenzothiazoline-6-sulfonic acid), high amounts of anthracene and benzo[a]pyrene, 80% and 97%, respectively, were transformed under optimal conditions of 60°C, pH 5, and 5 mmol l(-1) CuCl(2) after a 24-h incubation period. Other PAHs including fluorene, acenaphthylene, phenanthrene, and benzo[a]anthracene were also oxidized by the crude CueO. These findings indicated the potential application of prokaryotic laccases in enhancing the mineralization of benzo[a]pyrene by PAH-degrading bacteria.  相似文献   

11.
A fragment of the human β-chain of hemoglobin (HEM), hHEMβ111-146, was shown to have broad antimicrobial properties. The 3.9-kDa peptide was postulated to occur in high concentrations in placenta tissue. We established a reliable method to quantify hHEMβ111-146 in placenta tissue. Our methodology consists of a tissue extraction step (step 1), a chromatographic enrichment step (step 2), and a final quantification step (step 3) by enzyme-linked immunosorbent assay (ELISA). The specificity of the ELISA reaction was confirmed by parallel analysis of the samples via Western blot (step 4). The ELISA measured the absorbance of a tetramethylbenzidine substrate at 450 nm. It showed no cross-reactivity with the corresponding γ- and α-HEM regions and low cross-reactivity with the β-HEM region and full-length HEM. The sample preparation procedure enabled a prepurification of hHEMβ111-146, completely eliminating cross-reactive proteins and HEM peptides. The linear range of detection in step 3 was 20-200 ng/well (200-2000 μg/L) with a limit of quantification of 23 ng/well (230 μg/L) and a limit of detection of 7 ng/well (70 μg/L). The assay was characterized by good linearity (r> 0.99), intraday precision (coefficient of variation [CV] = 2.2-8.3%), interday precision (CV = 1.8-9.1%), and accuracy (76-109%). The mean recovery of the ELISA was determined to be 97%, and the overall recovery during steps 1-3 was found to be 40.3 ± 2.5%. We measured concentrations from 0.28 to 0.74 mg/g placenta tissue of the hHEMβ111-146 in different placenta samples with an average concentration of 0.57 mg/g. This abundant concentration supports an important physiological role of hHEMβ111-146 in the placenta infective barrier.  相似文献   

12.
Yan J  Wang L  Fu PP  Yu H 《Mutation research》2004,557(1):99-108
The photomutagenicity of 16 polycyclic aromatic hydrocarbons (PAHs), all on the United States Environmental Protection Agency (US EPA) priority pollutant list, was studied. Concomitant exposing the Salmonella typhimurium bacteria strain TA102 to one of the PAHs and light (1.1 J/cm2 UVA+2.1 J/cm2 visible) without the activation enzyme S9, strong photomutagenic response is observed for anthracene, benz[a]anthracene, benzo[ghi]perylene, benzo[a]pyrene, indeno[1,2,3-cd]pyrene, and pyrene. Under the same conditions, acenaphthene, acenaphthylene, benzo[k]fluoranthene, chrysene, and fluorene are weakly photomutagenic. Benzo[b]fluoranthene, fluoranthene, naphthalene, phenanthrene, and dibenz[a,h]anthracene are not photomutagenic. These results indicate that PAHs can be activated by light and become mutagenic in Salmonella TA102 bacteria. At the same time, the mutagenicity for all the 16 PAHs was examined with the standard mutagenicity test with 10% S9 as the activation system. Benzo[b]fluoranthene, benzo[k]fluoranthene, chrysene, acenaphthylene, and fluorene are weakly mutagenic, while the rest of the PAHs are not. In general, the photomutagenicity of PAHs in TA102 does not correlate with their S9-activated mutagenicity in either TA102 or TA98/TA100 since they involve different activation mechanisms.  相似文献   

13.
High-molecular-weight (HMW) polycyclic aromatic hydrocarbons (PAHs) are pollutants that persist in the environment due to their low solubility in water and their sequestration by soil and sediments. The addition of a water-immiscible, nonbiodegradable, and biocompatible liquid, silicone oil, to a soil slurry was studied to promote the desorption of PAHs from soil and to increase their bioavailability. First, the transfer into silicone oil of phenanthrene, pyrene, chrysene, and benzo[a]pyrene added to a sterilized soil (sandy soil with 0.65% total volatile solids) was measured for 4 days in three two-liquid-phase (TLP) slurry systems each containing 30% (w/v) soil but different volumes of silicone oil (2.5%, 7.5%, and 15% [v/v]). Except for chrysene, a high percentage of these PAHs was transferred from soil to silicone oil in the TLP slurry system containing 15% silicone oil. Rapid PAH transfer occurred during the first 8 h, probably resulting from the extraction of nonsolubilized and of poorly sorbed PAHs. This was followed by a period in which a slower but constant transfer occurred, suggesting extraction of more tightly bound PAHs. Second, a HMW PAH-degrading consortium was enriched in a TLP slurry system with a microbial population isolated from a creosote-contaminated soil. This consortium was then added to three other TLP slurry systems each containing 30% (w/v) sterilized soil that had been artificially contaminated with pyrene, chrysene, and benzo[a]pyrene, but different volumes of silicone oil (10%, 20%, and 30% [v/v]). The resulting TLP slurry bioreactors were much more efficient than the control slurry bioreactor containing the same contaminated soil but no oil phase. In the TLP slurry bioreactor containing 30% silicone oil, the rate of pyrene degradation was 19 mg L(-)(1) day(-)(1) and no pyrene was detected after 4 days. The degradation rates of chrysene and benzo[a]pyrene in the 30% TLP slurry bioreactor were, respectively, 3.5 and 0.94 mg L(-)(1) day(-)(1). Low degradation of pyrene and no significant degradation of chrysene and benzo[a]pyrene occurred in the slurry bioreactor. This is the first report in which a TLP system was combined with a slurry system to improve the biodegradation of PAHs in soil.  相似文献   

14.
A dual-label HPLC assay to measure femtomole quantities of ethyl acetate-extractable [3H]benzo[a]pyrene metabolites was developed. 14C-labeled metabolites of benzo[a]pyrene formed by rat liver 9000g supernatant were used as both internal standards and chromatographic markers. The percentage deviation between assays was determined to be between 11 and 13% for 9,10-dihydro-9,10-dihydroxybenzo[a]pyrene, 7,8-dihydro-7,8-dihydroxybenzo[a]pyrene, benzo[a]pyrene-3,6-quinone, benzo[a]pyrene-1,6-quinone, and 9-hydroxybenzo[a]pyrene, 22% for 4,5-dihydro-4,5-dihydroxybenzo[a]pyrene, and less than 5% for 3-hydroxybenzo[a]pyrene. The detection limit of this assay was between 3 and 10 fmol per metabolite. The application of this technique to the metabolism of [3H]benzo[a]pyrene by microsomes of hamster and human oral cavity tissue is described.  相似文献   

15.
This study investigated the biodegradation of high-molecular-weight polycyclic aromatic hydrocarbons (PAHs) in liquid media and soil by bacteria (Stenotrophomonas maltophilia VUN 10,010 and bacterial consortium VUN 10,009) and a fungus (Penicillium janthinellum VUO 10, 201) that were isolated from separate creosote- and manufactured-gas plant-contaminated soils. The bacteria could use pyrene as their sole carbon and energy source in a basal salts medium (BSM) and mineralized significant amounts of benzo[a]pyrene cometabolically when pyrene was also present in BSM. P. janthinellum VUO 10,201 could not utilize any high-molecular-weight PAH as sole carbon and energy source but could partially degrade these if cultured in a nutrient broth. Although small amounts of chrysene, benz[a]anthracene, benzo[a]pyrene, and dibenz[a,h]anthracene were degraded by axenic cultures of these isolates in BSM containing a single PAH, such conditions did not support significant microbial growth or PAH mineralization. However, significant degradation of, and microbial growth on, pyrene, chrysene, benz[a]anthracene, benzo[a]pyrene, and dibenz[a,h]anthracene, each as a single PAH in BSM, occurred when P. janthinellum VUO 10,201 and either bacterial consortium VUN 10,009 or S. maltophilia VUN 10,010 were combined in the one culture, i.e., fungal-bacterial cocultures: 25% of the benzo[a]pyrene was mineralized to CO(2) by these cocultures over 49 days, accompanied by transient accumulation and disappearance of intermediates detected by high-pressure liquid chromatography. Inoculation of fungal-bacterial cocultures into PAH-contaminated soil resulted in significantly improved degradation of high-molecular-weight PAHs, benzo[a]pyrene mineralization (53% of added [(14)C]benzo[a]pyrene was recovered as (14)CO(2) in 100 days), and reduction in the mutagenicity of organic soil extracts, compared with the indigenous microbes and soil amended with only axenic inocula.  相似文献   

16.
Our interests are in the development of immunoassay-based fast scanning methods for persistent organic pollutants. To develop the immunoassay method of polybrominated diphenyl ether (PBDE), a model compound of PBDE, 2,3′,4,5′,6-pentabromodiphenylether (BDE-121), has been chosen to develop its antibody and the competitive indirect enzyme-linked immunosorbent assay (ELISA) is developed. The hapten of BDE-121 containing reactive carboxylic acid was synthesized and conjugated to carrier proteins (bovine serum albumin [BSA] and ovalbumin [OVA]). Anti-BDE-121 polyclonal antibody was then developed in rabbits as a result of immunization with the BDE-121–BSA conjugate. The optimal amount of coating antigen BDE-121–OVA conjugate and the dilution of antiserum needed in the ELISA were determined with the checkerboard method, and the effects of the properties of PBST (phosphate-buffered saline and Tween 20) buffer (pH and salt concentration) and chemical solvent (types and concentrations) on the ELISA were investigated to achieve a rapid robust assay with high sensitivity. Under the optimized conditions, the developed indirect ELISA shows a linear detection range from 1.74 to 84.1 ng/ml, with an IC50 value of 8.07 ng/ml and a detection limit of 0.644 ng/ml. In total, 11 kinds of compounds were tested for calculating the cross-reactivity, which was less than 8% for nearly all of them. Real samples were analyzed by the proposed immunoassay and gas chromatography/mass spectrometry (GC/MS).  相似文献   

17.
The white-rot fungus Pleurotus ostreatus was able to degrade the polycyclic aromatic hydrocarbons (PAHs) benzo[a]anthracene, chrysene, benzo[b]fluoranthene, benzo[k]fluoranthene, benzo[a]pyrene, dibenzo[a,h]anthracene, and benzo[ghi]perylene in nonsterile soil both in the presence and in the absence of cadmium and mercury. During 15 weeks of incubation, recovery of individual compounds was 16 to 69% in soil without additional metal. While soil microflora contributed mostly to degradation of pyrene (82%) and benzo[a]anthracene (41%), the fungus enhanced the disappearance of less-soluble polycyclic aromatic compounds containing five or six aromatic rings. Although the heavy metals in the soil affected the activity of ligninolytic enzymes produced by the fungus (laccase and Mn-dependent peroxidase), no decrease in PAH degradation was found in soil containing Cd or Hg at 10 to 100 ppm. In the presence of cadmium at 500 ppm in soil, degradation of PAHs by soil microflora was not affected whereas the contribution of fungus was negligible, probably due to the absence of Mn-dependent peroxidase activity. In the presence of Hg at 50 to 100 ppm or Cd at 100 to 500 ppm, the extent of soil colonization by the fungus was limited.  相似文献   

18.
Polycyclic aromatic hydrocarbons (PAHs) are carcinogens with varying potencies. These compounds are metabolized to diol epoxides that react to form DNA adducts. Nucleotide excision repair is a critical cellular defense against these bulky DNA adducts which, if not repaired, can lead to mutations and the initiation of cancer. The structural features of the PAH-adducts play a role in differential repair of these adducts by the global genomic repair subpathway of nucleotide excision repair. DNA adducts derived from the PAHs containing bay-regions are repaired more rapidly than adducts derived from PAHs containing fjord-regions. We have employed the host cell reactivation assay to examine the rate of repair of these adducts in an actively transcribing gene. The pGL3 plasmid containing a luciferase gene was damaged with diol epoxides of benzo[a]pyrene (B[a]P-DE), dibenzo[a,l]pyrene (DB[a,l]P-DE), benzo[g]chrysene (B[g]Ch-DE), and benzo[c]phenanthrene (B[c]Ph-DE). The plasmids were transfected into B-lymphocytes with normal repair capacity as well as lymphocytes derived from patients with the XP-A, XP-C and CS-B syndromes. We found that XPA cells were able to transcribe slowly past B[g]Ch-adducts but not the other PAHs. Using the amount of luciferase produced as a measure of DNA repair, we found that the relative rates of repair in the actively transcribing luciferase gene was B[a]P-DE > DB[a,l]P-DE, B[g]Ch-DE, >B[c]Ph-DE in repair proficient and XP-C cells. These results indicate that the abilities to transcribe past and to repair the PAH adducts are dependent on different structural features of the DNA adducts.  相似文献   

19.
The potential of polycyclic aromatic hydrocarbons (PAHs) to modulate microsomal epoxide hydrolase activity, determined using benzo[a]pyrene 5-oxide as substrate, in human liver, was evaluated and compared to rat liver. Precision-cut liver slices prepared from fresh human liver were incubated with six structurally diverse PAHs, at a range of concentrations, for 24 h. Of the six PAHs studied, benzo[a]pyrene, dibenzo[a,h]anthracene and fluoranthene gave rise to a statistically significant increase in epoxide hydrolase activity, which was accompanied by a concomitant increase in epoxide hydrolase protein levels determined by immunoblotting. The other PAHs studied, namely dibenzo[a,l]pyrene, benzo[b]fluoranthene and 1-methylphenanthrene, influenced neither activity nor enzyme protein levels. When rat slices were incubated under identical conditions, only benzo[a]pyrene and dibenzo[a,h]anthracene elevated epoxide hydrolase activity, which was, once again accompanied by a rise in protein levels. At the mRNA level, however, all six PAHs caused an increase, albeit to different extent. In rat, epoxide hydroxylase activity in lung slices was much lower than in liver slices. In lung slices, epoxide hydrolase activity was elevated following exposure to benzo[a]pyrene and dibenzo[a,l]pyrene and, to a lesser extent, 1-methylphenanthrene; similar observations were made at the protein level. At both activity and protein levels extent of induction was far more pronounced in the lung compared with the liver. It is concluded that epoxide hydrolase activity is an inducible enzyme by PAHs, in both human and rat liver, but induction potential by individual PAHs varies enormously, depending on the nature of the compound involved. Marked tissue differences in the nature of PAHs stimulating activity in rat lung and liver were noted. Although in the rat basal lung epoxide hydrolase activity is much lower than liver, it is more markedly inducible by PAHs.  相似文献   

20.
[背景] 真菌和细菌被认为在多环芳烃污染土壤生物修复过程中发挥协同作用,目前在真实土壤体系中开展真菌-细菌协同降解研究较少。[目的] 研究真菌和细菌对不同种类多环芳烃降解的差异及对蒽和苯并[a]蒽的生物强化与协同作用。[方法] 选用多环芳烃降解真菌和细菌各一株,在液体纯培养体系下分析它们对不同种类多环芳烃降解的差异,在土壤体系中采用放射性同位素示踪技术研究2种微生物对蒽和苯并[a]蒽的生物强化与协同作用。[结果] 供试细菌鞘脂菌NS7能够很好地降解低环种类多环芳烃,以蒽作为唯一碳源时可以将其完全降解,在复合污染条件下对菲、蒽、荧蒽、芘等降解效果突出(>90%),对苯并[a]芘降解效果较差(9.76%)。相比而言,供试真菌糙皮侧耳菌对苯并[a]芘具有更好的降解效果(21.18%),对低环多环芳烃降解效果明显不如降解菌NS7。在自然土壤中,蒽和苯并[a]蒽具有明显不同的矿化效率,分别为18.61%和4.28%,在蒽污染土壤中加入鞘脂菌NS7并未显著提高蒽的矿化率(P>0.05),相比而言,苯并[a]蒽污染土壤中加入糙皮侧耳显著提高了污染物矿化效率(2.24倍),表明真菌和细菌在土壤环境中的定殖存活能力可能影响了生物强化效果。采用灭菌土壤排除土著微生物的竞争排斥作用,研究了真菌菌丝对生物强化降解的影响,发现在蒽污染土壤中,真菌菌丝的迁移作用显著提高了细菌鞘脂菌NS7对污染物的矿化率,从1.75%提高到5.91%;而在苯并[a]蒽灭菌污染土壤中,接种糙皮侧耳却没有发现苯并[a]蒽矿化率提高的现象,表明自然土壤中真菌强化降解苯并[a]蒽的作用可能是源于真菌菌丝促进污染物和土著降解菌的接触,而非直接来自真菌本身。[结论] 细菌能够很好地降解低环种类多环芳烃,而真菌对高环种类多环芳烃降解效果较好。真菌可能通过菌丝促进土著微生物在土壤中的迁移,增大多环芳烃和土著降解菌的接触,从而促进了多环芳烃降解。研究加深了对多环芳烃污染土壤生物强化修复的认识,对发展基于真菌-细菌协同作用的生物强化与调控技术提供理论指导。  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