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 46 毫秒
1.
多环芳烃荧蒽诱导拟南芥氧化胁迫   总被引:3,自引:0,他引:3  
选用模式植物拟南芥为材料,以四环的多环芳烃(PAHs)荧蒽为研究对象,从植物对非生物胁迫响应紧密相关的抗氧化酶及膜保护系统的变化入手,研究了植物对多环芳烃胁迫的生理响应.结果表明:荧蒽胁迫下拟南芥经历了氧化胁迫和膜脂过氧化过程.0.75 mmol·L-1的荧蒽使拟南芥光合作用过程受到抑制;1.00 mmol·L-1的荧蒽使拟南芥丙二醛(MDA)含量极显著增加, 抗坏血酸过氧化物酶(APX)活性极显著下降, 膜脂过氧化作用加剧,1.25 mmol·L-1的荧蒽使拟南芥过氧化物酶(POD)活性极显著下降,H2O2在细胞内累积,拟南芥明显受害.  相似文献   

2.
Relatively little is known about the mutagenicity of C24H14 PAH, a diverse group of five- and six-ring PAH, some of which are present at trace levels in the environment. To better understand the mutagenicity of this class of compounds, 11 C24H14 PAH, including benzo[a]perylene, benzo[b]perylene, dibenzo[a,e]fluoranthene, dibenzo[a,f]fluoranthene, dibenzo[j,l]fluoranthene, dibenzo[a,h]pyrene, dibenzo[a,i]pyrene, dibenzo[e,l]pyrene, naphtho[1,2-b]fluoranthene, naphtho[2,3-a]pyrene, and naphtho[2,3-e]pyrene, were tested in a mutagenicity assay based on human h1A1v2 cells. h1A1v2 cells are a line of human B-lymphoblastoid cells that have been engineered to express cytochrome P4501A1 (CYP1A1), an enzyme capable of metabolizing promutagenic PAH. Mutagenicity was measured at the thymidine kinase (tk) locus following a 72-h exposure period. Our results show that nine of the compounds were mutagenic. Benzo[a]perylene, dibenzo[a,e]fluoranthene, dibenzo[a,i]pyrene, and naphtho[2,3-a]pyrene were the most potent mutagens, having minimum mutagenic concentrations (MMC) (i.e., the dose at which the induced response was twice that of the negative controls) in the 1-5 ng/ml range. Benzo[b]perylene, dibenzo[a,h]pyrene, dibenzo[a,f]fluoranthene, and naphtho[2,3-e]pyrene were somewhat less potent mutagens, having MMC in the 10-30 ng/ml range. Dibenzo[e,l]pyrene, which had an MMC of 280 ng/ml, was the least potent mutagen. Dibenzo[j,l]fluoranthene and naphtho[1,2-b]fluoranthene were not mutagenic at the doses tested (1-3000 ng/ml). The most mutagenic compounds were also quite toxic. At the highest doses tested, benzo[a]perylene, dibenzo[a,e]fluoranthene, dibenzo[a,i]pyrene, dibenzo[a,h]pyrene, and dibenzo[a,f]fluoranthene induced > 60% killing, and naphtho[2,3-a]pyrene and naphtho[2,3-e]pyrene induced > 50% killing. Benzo[b]perylene, dibenzo[e,l]pyrene, dibenzo[j,l]fluoranthene, and naphtho[1,2-b]fluoranthene induced < 50% killing at the highest doses tested. Comparing these results to a previous study in which nine other C24H14 PAH were tested for mutagenicity in this same assay, it was found that dibenzo[a]pyrene isomers were generally more mutagenic than the other groups of C24H14 PAH tested. These observations are discussed with emphasis given to identifying C24H14 PAH that may be important environmental mutagens.  相似文献   

3.
The mouse skin cocarcinogens fluoranthene, pyrene, and undecane were used with the indirect-acting carcinogen, benzo(a)pyrene (BP), and the direct-acting alkylating carcinogen, ß-propiolactone (BPL), in an in vitro transformation assay. Dose response, cytotoxicity, and transformation studies with these compounds were performed with a subclone (A31-1-1) of the Balb/3T3 cell line. Transformation frequencies were found to increase with increasing concentrations of BP used up to 1.0 µg/ml or when BPL was used up to 4.0 µg/ml. A significant increase (P<0.05) in the transformation frequency over that seen with carcinogen alone was observed when cells were exposed to a combination of fluoranthene (4.0 µg/ml) and BP (0.063 µg/ml) or pyrene (5.0 µg/ml) and BP (0.063 µg/ml). Thus, the transformation frequency obtained with BP + fluoranthene was 3.8 × 10–4 compared to 1.2 × 10–4 when BP was tested alone. Similarly, the transformation frequency using BP + pyrene was 2.8 × 10–4 vs 1.2 × 10–4 when BP was tested alone. Undecane did not exert any cocarcinogenic effect with BP in the dose range tested. In this in vitro assay, no cocarcinogenic effect was observed when BPL was used with any of the above mouse skin cocarcinogens. All cells isolated from transformed foci showed characteristics of transformed cells including anchorage-independent growth.Abbreviations BP benzo(a)pyrene - BPL ß-propiolactone - CE cloning efficiency - CE50 median CE - RCE relative CE Department of Cell Biology, New York University Medical CenterInstitute for Cancer Research, Fox Chase Cancer Center, Philadelphia, PA.Contribution No. L217 from the Laboratory of Organic Chemistry and Carcinogenesis.  相似文献   

4.
A soil sample collected underneath a sewage pipe of the west side of Yangpu refining factory in Haikou city, Hainan Province, China was inoculated in minimum medium supplemented with fluoranthene. After 8 enrichment cycles, a bacterial consortium (Y12) was obtained through water-silicone oil dual system in the laboratory. The consortium Y12 could degrade a mixture of polycyclic aromatic hydrocarbons (PAHs) including phenanthrene, anthracene, fluoranthene, pyrene and benzo[a]pyrene. The consortium Y12 was repeatedly cultured for more than 40 circles, from which a bacterial strain FB3 was isolated. This strain was identified as a Sphingobium sp. through the 16S rDNA sequence alignment. Strain FB3 could degrade 99 ± 0.4%, 67 ± 2%, 97 ± 3%, 72 ± 8%, and 6 ± 2% (uncorrected degradation percentages) of phenanthrene, anthracene, fluoranthene and pyrene each at level of 100 mg L−1 and benzo[a]pyrene at 10 mg L−1, respectively, in 10 days, which the five PAHs were the sole carbon source as a mixture in minimum medium. The degradation percentages of phenanthrene, anthracene, fluoranthene, pyrene (each at level of 100 mg L−1) and benzo[a]pyrene (10 mg L−1) by consortium Y12 were 99 ± 0.1%, 65 ± 3%, 99 ± 0.3%, 79 ± 1% and 7 ± 6%, respectively, in 10 days. Strain FB3 could degrade those PAHs under a range of pH 5–9, being optimum at pH 7.  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.
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   

7.
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   

8.
Effects of glucose, ammonium ions and phosphate on avilamycin biosynthesis in Streptomyces viridochromogenes AS4.126 were investigated. Twenty grams per liter of glucose, 10 mmol/L ammonium ions, and 10 mmol/L phosphate in the basal medium stimulated avilamycin biosynthesis. When the concentrations of glucose, ammonium ions, and phosphate in the basal medium exceeded 20 g/L, 10 mmol/L, and 10 mmol/L, respectively, avilamycin biosynthesis greatly decreased. When 20 g/L glucose was added at 32 h, avilamycin yield decreased by 70.2%. Avilamycin biosynthesis hardly continued when 2-deoxy-glucose was added into the basal medium at 32 h. There was little influence on avilamycin biosynthesis with the addition of the 3-methyl-glucose (20 g/L) at 32 h. In the presence of excess (NH4)2SO4 (20 mmol/L), the activities of valine dehydrogenase and glucose-6-phosphate dehydrogenase were depressed 47.7 and 58.3%, respectively, of that of the control at 48 h. The activity of succinate dehydrogenase increased 49.5% compared to the control at 48 h. The intracellular adenosine triphosphate level and 6-phosphate glucose content of S. viridochromogenes were 128 and 129%, respectively, of that of the control at 48 h, with the addition of the 40 mmol/L of KH2PO4. As a result, high concentrations of glucose, ammonium ions, and inorganic phosphate all led to the absence of the precursors for avilamycin biosynthesis and affected antibiotic synthesis.  相似文献   

9.
Aerobic, mesophilic bacteria from coal tar–contaminated soil were analyzed for pyrene utilization capacity and identified by 16S ribosomal DNA sequencing as members of three genera: Bacillus spp., Pseudomonas sp., and Rhodococcus sp. The soil contained nine different hazardous polyaromatic hydrocarbons (PAHs): benzo[g, h, i]perylene, dibenzo[a, h]anthracene, indeno[1,2,3-c,d]pyrene, pyrene, acenaphthylene, fluorene, phenanthrene, benzo[k]fluoranthene, and benzo[b]fluoranthene. Bacillus spp. (PK-6) MTCC 1005 showed 56.4% utilization of pyrene (C16H10) (50 μg ml?1) in 4 days, with growth associated biosurfactant activity and resulted in the formation of five new intermediates: phenanthrene (C14H10), 9,10-diphenylphenanthrene (C26H18), 9-methoxyphenanthrene (C15H12O), 5,6,7,8-tetrahydro-1-naphthoic acid (C11H12O2), and 1,6,7-trimethylnaphthalene (C13H14). The results suggested that Bacillus spp. could be found suitable for practical field application for effective in situ PAH bioremediation.  相似文献   

10.
A sensitive capillary electrophoresis–electrochemiluminescence (CE–ECL) assay with an ionic liquid (IL) was developed for the determination of arecoline in areca nut. The IL, 1‐butyl‐3‐methylimidazolium tetrafluoroborate (BMImBF4), was an effective additive improved not only the separation selectivity but also the detection sensitivity of the analyte. BMImBF4 in the separation electrolyte made the resistance of the separation buffer much lower than that of the sample solution, which resulted in an enhanced field amplified electrokinetic injection CE. ECL intensity of arecoline is about two times higher than that of the analyte with phosphate–IL buffer system. Resolution between arecoline and other unknown compounds in real samples was improved. Under the optimized conditions (ECL detection at 1.2 V, 16 kV separation voltage, 20 mmol/L phosphate with 10 mmol/L BMImBF4 buffer at pH 7.50, 5 mmol/L Ru(bpy)32+ and 50 mmol/L phosphate buffer in the detection reservoir), a detection limit of 5 × 10–9 mol/L for arecoline was obtained. Relative standard deviations of the ECL intensity and the migration time were 4.51% and 0.72% for arecoline. This method was successfully applied to determination of the amount of arecoline in areca nut within 450 s. Copyright © 2012 John Wiley & Sons, Ltd.  相似文献   

11.
Pyrene and fluoranthene, when supplied as the sole carbon source, were not degraded by Burkholderia sp. VUN10013. However, when added in a mixture with phenanthrene, both pyrene and fluoranthene were degraded in liquid broth and soil. The amounts of pyrene and fluoranthene in liquid media (initial concentrations of 50 mg l−1 each) decreased to 42.1% and 41.1%, respectively, after 21 days. The amounts of pyrene and fluoranthene in soil (initial concentrations of 75 mg kg−1 dry soil each) decreased to 25.8% and 12.1%, respectively, after 60 days. None of the high molecular weight (HMW) polycylic aromatic hydrocarbons (PAHs) tested adversely affected phenanthrene degradation by this bacterial strain and the amount of phenanthrene decreased rapidly within 3 and 15 days of incubation in liquid broth and soil, respectively. Anthracene also stimulated the degradation of pyrene or fluoranthene by Burkholderia sp. VUN10013, but to a lesser extent than phenanthrene. The extent of anthracene degradation decreased in the presence of these HMW PAHs.  相似文献   

12.
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   

13.
Degradation of pyrene by Mycobacterium flavescens   总被引:1,自引:0,他引:1  
 A strain of Mycobacterium flavescens was isolated from polluted sediments. It was capable of utilizing pyrene as a sole source of carbon and energy. When pyrene was supplied as a suspension at 50 μg/ml, the generation time was 9.6 h and the rate of pyrene utilization was 0.56 μg ml-1 day-1. In addition to pyrene, the strain could mineralize phenanthrene (17.7%) and fluoranthene (17.9%), but failed to mineralize naphthalene, chrysene, anthracene, fluorene, acenaphthene and benzo[a]pyrene, as determined by recovery of radiolabeled CO2 in incubations conducted for 2 weeks under growth conditions. Metabolites produced during growth on pyrene were detected and characterized by HPLC and GC-MS. The product of initial ring oxidation, 4,5-dihydroxy-4,5-dihydropyrene was identified, as well as ring-fission products including 4-phenanthroic acid, phthalic acid, and 4,5-phenanthrenedioic acid. Received: 3 October 1995/Received last revision: 1 April 1996/Accepted: 15 April 1996  相似文献   

14.
Gelatinous zooplankton play an important role in marine food webs both as major consumers of metazooplankton and as prey of apex predators (e.g., tuna, sunfish, sea turtles). However, little is known about the effects of crude oil spills on these important components of planktonic communities. We determined the effects of Louisiana light sweet crude oil exposure on survival and bioaccumulation of polycyclic aromatic hydrocarbons (PAHs) in adult stages of the scyphozoans Pelagia noctiluca and Aurelia aurita and the ctenophore Mnemiopsis leidyi, and on survival of ephyra larvae of A. aurita and cydippid larvae of M. leidyi, in the laboratory. Adult P. noctiluca showed 100% mortality at oil concentration ≥20 µL L−1 after 16 h. In contrast, low or non-lethal effects were observed on adult stages of A. aurita and M. leidyi exposed at oil concentration ≤25 µL L−1 after 6 days. Survival of ephyra and cydippid larva decreased with increasing crude oil concentration and exposition time. The median lethal concentration (LC50) for ephyra larvae ranged from 14.41 to 0.15 µL L−1 after 1 and 3 days, respectively. LC50 for cydippid larvae ranged from 14.52 to 8.94 µL L−1 after 3 and 6 days, respectively. We observed selective bioaccumulation of chrysene, phenanthrene and pyrene in A. aurita and chrysene, pyrene, benzo[a]pyrene, benzo[b]fluoranthene, benzo[k]fluoranthene, and benzo[a]anthracene in M. leidyi. Overall, our results indicate that (1) A. aurita and M. leidyi adults had a high tolerance to crude oil exposure compared to other zooplankton, whereas P. noctiluca was highly sensitive to crude oil, (2) larval stages of gelatinous zooplankton were more sensitive to crude oil than adult stages, and (3) some of the most toxic PAHs of crude oil can be bioaccumulated in gelatinous zooplankton and potentially be transferred up the food web and contaminate apex predators.  相似文献   

15.
Microbiological analysis of soils from a polycyclic aromatic hydrocarbon (PAH)-contaminated site resulted in the enrichment of five microbial communities capable of utilizing pyrene as a sole carbon and energy source. Communities 4 and 5 rapidly degraded a number of different PAH compounds. Three pure cultures were isolated from community 5 using a spray plate method with pyrene as the sole carbon source. The cultures were identified as strains of Burkholderia ( Pseudomonas ) cepacia on the basis of biochemical and growth tests. The pure cultures (VUN 10 001, VUN 10 002 and VUN 10 003) were capable of degrading fluorene, phenanthrene and pyrene (100 mg l−1) to undetectable levels within 7–10 d in standard serum bottle cultures. Pyrene degradation was observed at concentrations up to 1000 mg l−1. The three isolates were also able to degrade other PAHs including fluoranthene, benz[ a ]anthracene and dibenz[ a , h ]anthracene as sole carbon and energy sources. Stimulation of dibenz[ a , h ]anthracene and benzo[ a ]pyrene degradation was achieved by the addition of small quantities of phenanthrene to cultures containing these compounds. Substrate utilization tests revealed that these micro-organisms could also grow on n -alkanes, chlorinated- and nitro-aromatic compounds.  相似文献   

16.
Cryopreservation of testicular sperm in the African clawed frog, Xenopus laevis, was tested using three penetrating cryoprotectants (DMSO, methanol, and glycerol) and three semen diluents (300 mmol/L glucose, 300 mmol/L sucrose, and a motility inhibiting saline [MIS] solution [150 mmol/L NaCl, 3 mmol/L KCL, 1 mmol/L Mg2SO4, 1 mmol/L CaCl2, and 20 mmol/L Tris, pH 8.0]). Three freezing rates and four thawing rates were also tested, and the best freezing/thawing conditions have been determined. The responses of sperm motility, viability, and fertility were assessed. Incubation of the sperm macerates with penetrating cryoprotectants showed that DMSO was the least toxic and methanol the most toxic. Semen in cryodiluents frozen 10 cm above the surface of liquid nitrogen (freezing rate of 20 to 25 °C/min) and thawed at room temperature for 40 sec had significantly higher percentages of motile and viable sperm than that of semen frozen 5 cm or 8 cm above the surface of liquid nitrogen and thawed at 5, 25, or 30 °C for 10, 15, or 60 sec, respectively. Sperm frozen in MIS containing 5% DMSO had a higher hatching rate than that of sperm frozen in sucrose and glucose diluents containing 5% or 10% DMSO and in MIS containing 10% DMSO. Addition of 73 mmol/L sucrose to the sperm extender MIS + 5% DMSO could improve the postthaw sperm motility and fertility. In conclusion, dilution of collected sperm in MIS solution (to have a final concentration of 6.5 × 106 to 8 × 106/mL) containing 5% DMSO and 73 mmol/L sucrose, freezing in a vapor of liquid nitrogen at 10 cm above the surface, and thawing at room temperature for 40 sec was the best cryopreservation protocol. This protocol gave 70% hatching rate, 80% motility rate, and 75% viability rate of fresh hormonally induced sperm.  相似文献   

17.
The microsomal oxidation of 12 frequently occurring environmental polycyclic aromatic hydrocarbons after incubation with rat-liver microsomes has been studied and their metabolites characterized by means of gas-liquid chromatography/mass spectrometry. The method enables the detection and characterisation of phenols, diols, triols, and tetrols as trimethylsilyl ethers beside the original hydrocarbons. Moreover, the induction properties of some carcinogenic and non-carcinogenic hydrocarbons (benz[a]anthracene, pyrene, chrysene, benzo[a]-pyrene, benzo[e]pyrene, benzo[b]fluoranthene, benzo[j]fluoranthene, benzo[k]fluoranthene) have been studied. Except pyrene and benzo[e]pyrene, all compounds investigated significant but different induction factors. The relevance of the induction for an estimation of the biological effect of environmental polycyclic aromatic hydrocarbons is discussed.  相似文献   

18.
《Chirality》2017,29(8):422-429
A ligand‐exchange micellar electrokinetic capillary electrophoresis system with copper(II)‐L‐isoleucine complexes as the chiral selector incorporated in micelles of sodium dodecyl sulfate (SDS) was developed for the enantioseparation of ofloxacin and its four related substances (impurities A, C, E, and F). The effects of important parameters affecting separation such as buffer pH, SDS concentration, chiral selector concentration, and organic additive were investigated in detail. Under optimum experimental conditions, enantioseparation of ofloxacin, impurities A, C, E, and F enantiomers was accomplished with resolutions of 4.28, 2.83, 3.40, 3.58, and 2.46, respectively. Further, simultaneous separation of impurities A, C, E, and F enantiomers was achieved using 10 mmol/L NH4OAc as the running buffer containing 4 mmol/L copper sulfate,20 mmol/L L‐isoleucine, 20 mmol/L SDS, and 5% methanol at pH 8.5. To the best of our knowledge, the simultaneous enantioseparation of four impurities of ofloxacin has not been reported previously.  相似文献   

19.
为了研究CaCl2对NaCl胁迫下酸枣幼苗根、茎、叶的氮代谢影响,探索钙缓解幼苗NaCl胁迫的作用途径。该研究以酸枣幼苗为试验材料,检测不同浓度CaCl2(0、5、10、20 mmol/L)对NaCl(150 mmol/L)胁迫下幼苗叶片H2O2、O-·2含量,根、茎、叶中硝酸还原酶(NR)、谷氨酰胺合成酶(GS)、谷氨酸合酶(GOGAT)活性及游离氨基酸、可溶性蛋白、硝态氮含量的影响,并采用主成分分析法筛选出评价CaCl2缓解NaCl胁迫效应的生理指标。结果表明:与NaCl胁迫相比,盐胁迫幼苗叶片的H2O2、O-·2积累量在5、10 mmol/L CaCl2处理下显著减少;GOGAT活性在5、10 mmol/L CaCl2处理下的植株根和茎内以及各浓度 CaCl2处理的叶内均显著升高, GS、NR活性在10、20 mmol/L CaCl2处理的根内和10 mmol/L CaCl2处理的茎内以及5、10、20 mmol/L CaCl2处理的叶内均显著升高;可溶性蛋白含量在5、10、20 mmol/L CaCl2处理的根、茎、叶内均显著升高,游离氨基酸含量在10、20 mmol/L CaCl2处理的根和茎内以及10 mmol/L CaCl2处理的叶内均显著升高,硝态氮含量在10 mmol/L CaCl2处理的根和茎内以及5、10、20 mmol/L CaCl2处理的叶内均显著升高。研究发现,150 mmol/L NaCl胁迫对酸枣幼苗造成明显过氧化伤害,抑制了体内氮代谢;外源CaCl2可通过促进幼苗根和茎内GS/GOGAT循环对NH4+的同化作用,提高叶片NR活性,加快硝态氮的转化速率,从而增强幼苗对NaCl胁迫的适应性,并以10 mmol/L CaCl2处理缓解效果最佳;游离氨基酸、GOGAT、NR可以作为CaCl2缓解幼苗NaCl胁迫伤害的评价指标。  相似文献   

20.
The effects of structure and concentration of surfactants on the biodegradation of fluoranthene, a three rings polycyclic aromatic hydrocarbon in the aqueous phase, as well as their effects on the biodegradation and enzyme activity were investigated. The toxicity ranking of studied surfactants is: non-ionic Tween 80 <anionic sodium dodecyl sulfate <cationic Tetradecyltrimethylammonium bromide. The maximum growth of Armillaria sp. F022 (>4,500 mg/L) was showed by Tween 80 (10 mg/L) culture, manifesting that the non-ionic surfactant present in the culture were beneficial to the fungal growth. Laccase showed the highest enzymes activity in all surfactants culture. Non-ionic Tween 80 showed a significant result for laccase activity (1,902 U/L) in the Armillaria sp. F022 culture. The increased enzymes cumulative activity may stem directly from the rising fluoranthene biodegradability as addition of appropriate surfactants. The biotransformation of fluoranthene was greatly improved by Tween 80, and totally fluoranthene degradation was obtained as Tween 80 was 10 mg/L. Two fluoranthene metabolites were isolated from the culture medium and analyzed by a thin layer chromatography, UV visible spectrometer and gas chromatography–mass spectrometry (GC–MS). The oxidation of fluoranthene is initiated by oxygenation at the C-2,3 positions resulting 9-fluorenone. At the end of experiment, one metabolite was detected in the culture extract and identified as phthalic acid. Evidently, Armillaria sp. F022 seems efficient, high effective and deserves further application on the enhanced bioremediation technologies for the treatment of fluoranthene-contaminated soil.  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