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.
Atrazine chlorohydrolase, TrzN (triazine hydrolase or atrazine chlorohydrolase 2), initiates bacterial metabolism of the herbicide atrazine by hydrolytic displacement of a chlorine substituent from the s-triazine ring. The present study describes crystal structures and reactivity of wild-type and active site mutant TrzN enzymes. The homodimer native enzyme structure, solved to 1.40 Å resolution, is a (βα)8 barrel, characteristic of members of the amidohydrolase superfamily. TrzN uniquely positions threonine 325 in place of a conserved aspartate that ligates the metal in most mononuclear amidohydrolases superfamily members. The threonine side chain oxygen atom is 3.3 Å from the zinc atom and 2.6 Å from the oxygen atom of zinc-coordinated water. Mutation of the threonine to a serine resulted in a 12-fold decrease in kcat/Km, largely due to kcat, whereas the T325D and T325E mutants had immeasurable activity. The structure and kinetics of TrzN are reminiscent of carbonic anhydrase, which uses a threonine to assist in positioning water for reaction with carbon dioxide. An isosteric substitution in the active site glutamate, E241Q, showed a large diminution in activity with ametryn, no detectable activity with atratone, and a 10-fold decrease with atrazine, when compared with wild-type TrzN. Activity with the E241Q mutant was nearly constant from pH 6.0 to 10.0, consistent with the loss of a proton-donating group. Structures for TrzN-E241Q were solved with bound ametryn and atratone to 1.93 and 1.64 Å resolution, respectively. Both structure and kinetic determinations suggest that the Glu241 side chain provides a proton to N-1 of the s-triazine substrate to facilitate nucleophilic displacement at the adjacent C-2.  相似文献   

2.
Atrazine is one of the most environmentally prevalent s-triazine-ring herbicides. The widespread use of atrazine and its toxicity necessitates search for remediation technology. As atrazine is still used in India as a major herbicide, exploration of atrazine-degrading bacterial community is of immense importance. Considering lack of reports on well characterized atrazine-degrading bacterial cultures from India and wide diversity and density of microorganisms in rhizosphere, soil sample from rhizosphere of atrazine-resistant plant was studied. Arthrobacter sp. strain isolated in this investigation utilizes atrazine as the sole nitrogen source. In addition, the bacterium degrades other triazines such as ametryn, cyanizine, propazine and simazine. PCR analysis confirms the presence of atzBCD and triazine hydrolase (trzN) genes on chromosomal DNA. Sequencing of the trzN gene reveals high sequence similarity with trzN from Nocardioides sp. C190. An inducible and intracellular atrazine chlorohydrolase enzyme was isolated and partially purified from this isolate. This study confirms the presence of atrazine-degrading microbial population in Indian soils and could be used efficiently for remediation of contaminated soils. Presence of trzN gene indicates possible presence of bacterial community with more efficient and novel enzymatic capabilities. Comparison of enzyme and gene structure of this isolate with other geographically distinct atrazine-degrading strains will help us in the better understanding of gene transfer and evolution.  相似文献   

3.
Laboratory studies were conducted to evaluate effects of selected herbicides on hatching of free eggs of the soybean cyst nematode, Heterodera glycines. The herbicides used were Atrazine (atrazine), Basagran (bentazon), Bladex (cyanazine), Blazer (acifluorfen), Command (clomazone), Lasso (alachlor), Sonalan (ethalfluralin), and Treflan (trifluralin). Treatments comprised two concentrations of commercial herbicide formulations and deionized water and 3.14 mM zinc sulfate as negative and positive controls, respectively. Eggs were extracted from females and cysts, surface disinfested, and incubated in herbicide or control solutions at 25 ± 2 C in darkness. Hatched second-stage juveniles were counted every other day for 24 days. Hatching of H. glycines eggs in 50 and 500 μg/ml Blazer was 42 to 67% less than that in deionized water and 6l to 78% less than that in zinc sulfate solution. Zinc sulfate significantly increased hatching activity in 50 μg/ml but not 500 μg/ml Blazer. The other herbicides tested at various concentrations had no significant effect on egg hatching. The specific component of Blazer inhibiting egg hatching is unknown. Suppression of hatching by Blazer indicates that this postemergence soybean herbicide may have a potential role in managing H. glycines.  相似文献   

4.
Atrazine, a herbicide widely used in corn production, is a frequently detected groundwater contaminant. Fourteen bacterial strains able to use this herbicide as a sole source of nitrogen were isolated from soils obtained from two farms in Canada and two farms in France. These strains were indistinguishable from each other based on repetitive extragenic palindromic PCR genomic fingerprinting performed with primers ERIC1R, ERIC2, and BOXA1R. Based on 16S rRNA sequence analysis of one representative isolate, strain C147, the isolates belong to the genus Pseudaminobacter in the family Rhizobiaceae. Strain C147 did not form nodules on the legumes alfalfa (Medicago sativa L.), birdsfoot trefoil (Lotus corniculatus L.), red clover (Trifolium pratense L.), chickpea (Cicer arietinum L.), and soybean (Glycine max L.). A number of chloro-substituted s-triazine herbicides were degraded, but methylthio-substituted s-triazine herbicides were not degraded. Based on metabolite identification data, the fact that oxygen was not required, and hybridization of genomic DNA to the atzABC genes, atrazine degradation occurred via a series of hydrolytic reactions initiated by dechlorination and followed by dealkylation. Most strains could mineralize [ring-U-14C]atrazine, and those that could not mineralize atrazine lacked atzB or atzBC. The atzABC genes, which were plasmid borne in every atrazine-degrading isolate examined, were unstable and were not always clustered together on the same plasmid. Loss of atzB was accompanied by loss of a copy of IS1071. Our results indicate that an atrazine-degrading Pseudaminobacter sp. with remarkably little diversity is widely distributed in agricultural soils and that genes of the atrazine degradation pathway carried by independent isolates of this organism are not clustered, can be independently lost, and may be associated with a catabolic transposon. We propose that the widespread distribution of the atrazine-degrading Pseudaminobacter sp. in agricultural soils exposed to atrazine is due to the characteristic ability of this organism to utilize alkylamines, and therefore atrazine, as sole sources of carbon when the atzABC genes are acquired.  相似文献   

5.
The s-triazine herbicide atrazine was rapidly mineralized (i.e., about 60% of 14C-ring-labelled atrazine released as 14CO2 within 21 days) by an agricultural soil from the Nile Delta (Egypt) that had been cropped with corn and periodically treated with this herbicide. Seven strains able to degrade atrazine were isolated by enrichment cultures of this soil. DNA fingerprint and phylogenetic studies based on 16S rRNA analysis showed that the seven strains were identical and belonged to the phylogeny of the genus Arthrobacter (99% similarity with Arthrobacter sp. AD38, EU710554). One strain, designated Arthrobacter sp. strain TES6, degraded atrazine and mineralized the 14C-chain-labelled atrazine. However, it was unable to mineralize the 14C-ring-labelled atrazine. Atrazine biodegradation ended in a metabolite that co-eluted with cyanuric acid in HPLC. This was consistent with its atrazine-degrading genetic potential, shown to be dependent on the trzN, atzB, and atzC gene combination. Southern blot analysis revealed that the three genes were located on a large plasmid of about 175 kb and clustered on a 22-kb SmaI fragment. These results reveal for the first time the adaptation of a North African agricultural soil to atrazine mineralization and raise interesting questions about the pandemic dispersion of the trzN, atzBC genes among atrazine-degrading bacteria worldwide.  相似文献   

6.
Atrazine, a herbicide widely used in corn production, is a frequently detected groundwater contaminant. Nine gram-positive bacterial strains able to use this herbicide as a sole source of nitrogen were isolated from four farms in central Canada. The strains were divided into two groups based on repetitive extragenic palindromic (rep)-PCR genomic fingerprinting with ERIC and BOXA1R primers. Based on 16S ribosomal DNA sequence analysis, both groups were identified as Nocardioides sp. strains. None of the isolates mineralized [ring-U-(14)C]atrazine. There was no hybridization to genomic DNA from these strains using atzABC cloned from Pseudomonas sp. strain ADP or trzA cloned from Rhodococcus corallinus. S-Triazine degradation was studied in detail in Nocardioides sp. strain C190. Oxygen was not required for atrazine degradation by whole cells or cell extracts. Based on high-pressure liquid chromatography and mass spectrometric analyses of products formed from atrazine in incubations of whole cells with H(2)(18)O, sequential hydrolytic reactions converted atrazine to hydroxyatrazine and then to the end product N-ethylammelide. Isopropylamine, the putative product of the second hydrolytic reaction, supported growth as the sole carbon and nitrogen source. The triazine hydrolase from strain C190 was isolated and purified and found to have a K(m) for atrazine of 25 microM and a V(max) of 31 micromol/min/mg of protein. The subunit molecular mass of the protein was 52 kDa. Atrazine hydrolysis was not inhibited by 500 microM EDTA but was inhibited by 100 microM Mg, Cu, Co, or Zn. Whole cells and purified triazine hydrolase converted a range of chlorine or methylthio-substituted herbicides to the corresponding hydroxy derivatives. In summary, an atrazine-metabolizing Nocardioides sp. widely distributed in agricultural soils degrades a range of s-triazine herbicides by means of a novel s-triazine hydrolase.  相似文献   

7.
A novel s-triazine-mineralizing bacterium—Nocardioides sp. strain DN36—was isolated from paddy field soil treated with ring-U-14C-labeled simetryn ([14C]simetryn) in a model paddy ecosystem (microcosm). In a tenfold-diluted R2A medium, strain DN36 liberated 14CO2 from not only [14C]simetryn but also three ring-U-14C-labeled s-triazines: atrazine, simazine, and propazine. We found that DN36 mineralized ring-U-14C–cyanuric acid added as an initial substrate, indicating that the bacterium mineralized s-triazine herbicides via a common metabolite, namely, cyanuric acid. Strain DN36 harbored a set of genes encoding previously reported s-triazine-degrading enzymes (TrzN-AtzB-AtzC), and it also transformed ametryn, prometryn, dimethametryn, atraton, simeton, and prometon. The findings suggest that strain DN36 can mineralize a diverse range of s-triazine herbicides. To our knowledge, strain DN36 is the first Nocardioides strain that can individually mineralize s-triazine herbicides via the ring cleavage of cyanuric acid. Further, DN36 could not grow on cyanuric acid, and the degradation seemed to occur cometabolically.  相似文献   

8.
The effects of the photosystem II herbicides diuron (3-(3,4-dichlorophenyl)-1,1-dimethylurea) and atrazine (2-chloro-4-ethylamino-6-isopropylamino-s-triazine) on the photosynthetic membranes of a cyanobacterium, Aphanocapsa 6308, were compared to the effects on a higher plant, Spinacia oleracea. The inhibition of photosystem II electron transport by these herbicides was investigated by measuring the photoreduction of the dye 2,6-dichlorophenol-indophenol spectrophotometrically using isolated membranes. The concentration of herbicide that caused 50% inhibition of electron transport (I50 value) in Aphanocapsa membranes for diuron was 6.8 × 10−9 molar and the I50 value for atrazine was 8.8 × 10−8 molar. 14C-labeled diuron and atrazine were used to investigate herbicide binding with calculated binding constants (K) being 8.2 × 10−8 molar for atrazine and 1.7 × 10−7 molar for diuron. Competitive binding studies carried out on Aphanocapsa membranes using radiolabeled [14C]atrazine and unlabeled diuron revealed that diuron competed with atrazine for the herbicide-binding site. Experiments involving the photoaffinity label [14C]azidoatrazine (2-azido-4-ethylamino-6-isopropylamino-2-triazine) and autoradiography of polyacrylamide gels indicated that the herbicide atrazine binds to a 32-kilodalton protein in Aphanocapsa 6308 cell extracts.  相似文献   

9.
A multiple-strain algal biosensor was constructed for the detection of herbicides inhibiting photosynthesis. Nine different microalgal strains were immobilised on an array biochip using permeable membranes. The biosensor allowed on-line measurements of aqueous solutions passing through a flow cell using chlorophyll fluorescence as the biosensor response signal. The herbicides atrazine, simazine, diuron, isoproturon and paraquat were detectable within minutes at minimal LOEC (Lowest Observed Effect Concentration) ranging from 0.5 to 100μgL−1, depending on the herbicide and algal strain. The most sensitive strains in terms of EC50 values were Tetraselmis cordiformis and Scherffelia dubia. Less sensitive species were Chlorella vulgaris, Chlamydomonas sp. and Pseudokirchneriella subcapitata, but for most of the strains no general sensitivity or resistance was found. The different responses of algal strains to the five herbicides constituted a complex response pattern (RP), which was analysed for herbicide specificity within the linear dose-response relationship. Comparisons of herbicide-specific RP to reference RPs of the five herbicides always showed the lowest deviation of the herbicide-specific RP tested with the reference RP of the same herbicide for the triazine and phenylurea herbicides. We therefore conclude that, in principle, identification of a specific herbicide is possible employing the algal sensor chip.  相似文献   

10.
We examined the ability of a soil bacterium, Agrobacterium radiobacter J14a, to degrade the herbicide atrazine under a variety of cultural conditions, and we used this bacterium to increase the biodegradation of atrazine in soils from agricultural chemical distribution sites. J14a cells grown in nitrogen-free medium with citrate and sucrose as carbon sources mineralized 94% of 50 μg of [14C-U-ring]atrazine ml−1 in 72 h with a concurrent increase in the population size from 7.9 × 105 to 5.0 × 107 cells ml−1. Under these conditions cells mineralized the [ethyl-14C]atrazine and incorporated approximately 30% of the 14C into the J14a biomass. Cells grown in medium without additional carbon and nitrogen sources degraded atrazine, but the cell numbers did not increase. Metabolites produced by J14a during atrazine degradation include hydroxyatrazine, deethylatrazine, and deethyl-hydroxyatrazine. The addition of 105 J14a cells g−1 into soil with a low indigenous population of atrazine degraders treated with 50 and 200 μg of atrazine g−1 soil resulted in two to five times higher mineralization than in the noninoculated soil. Sucrose addition did not result in significantly faster mineralization rates or shorten degradation lag times. However, J14a introduction (105 cells g−1) into another soil with a larger indigenous atrazine-mineralizing population reduced the atrazine degradation lag times below those in noninoculated treatments but did not generally increase total atrazine mineralization.  相似文献   

11.
The widespread use and relative persistence of s-triazine compounds such as atrazine and simazine have led to increasing concern about environmental contamination by these compounds. Few microbial isolates capable of transforming substituted s-triazines have been identified. Rhodococcus corallinus NRRL B-15444 has previously been shown to possess a hydrolase activity that is responsible for the dechlorination of the triazine compounds deethylsimazine (6-chloro-N-ethyl-1,3,5-triazine-2,4-diamine) (CEAT) and deethylatrazine (6-chloro-N-isopropyl-1,3,5-triazine-2,4-diamine) (CIAT). The enzyme responsible for this activity was purified and shown to be composed of four identical subunits of 54,000 Da. Kinetic experiments revealed that the purified enzyme is also capable of deaminating the structurally related s-triazine compounds melamine (2,4,6-triamino-1,3,5-triazine) (AAAT) and CAAT (2-chloro-4,6-diamino-1,3,5-triazine), as well as the pyrimidine compounds 2,4,6-triaminopyrimidine (AAAP) and 4-chloro-2,6-diaminopyrimidine (CAAP). The triazine herbicides atrazine and simazine inhibit the hydrolytic activities of the enzyme but are not substrates. Induction experiments demonstrate that triazine hydrolytic activity is inducible and that this activity rises approximately 20-fold during induction.  相似文献   

12.
Evolution of atrazine-degrading capabilities in the environment   总被引:2,自引:0,他引:2  
Since their first introduction in the mid 1950s, man-made s-triazine herbicides such as atrazine have extensively been used in agriculture to control broadleaf weed growth in different crops, and thus contributed to improving crop yield and quality. Atrazine is the most widely used s-triazine herbicide for the control of weeds in crops such as corn and sorghum. Although atrazine was initially found to be slowly and partially biodegradable, predominantly by nonspecific P450 monoxygenases which do not sustain microbial growth, microorganisms gradually evolved as a result of repeated exposure, started using it as a growth substrate and eventually succeeded in mineralizing it. Within three decades, an entirely new hydrolase-dependent pathway for atrazine mineralization emerged and rapidly spread worldwide among genetically different bacteria. This review focuses on the enzymes involved in atrazine mineralization and their evolutionary histories, the genetic composition of microbial populations involved in atrazine degradation and the biotechnologies that have been developed, based on these systems, for the bioremediation of atrazine contamination in the environment.  相似文献   

13.
Initial denitration of hexahydro-1,3,5-trinitro-1,3,5-triazine (RDX) by Rhodococcus sp. strain DN22 produces CO2 and the dead-end product 4-nitro-2,4-diazabutanal (NDAB), OHCNHCH2NHNO2, in high yield. Here we describe experiments to determine the biodegradability of NDAB in liquid culture and soils containing Phanerochaete chrysosporium. A soil sample taken from an ammunition plant contained RDX (342 μmol kg−1), HMX (octahydro-1,3,5,7-tetranitro-1,3,5,7-tetrazocine; 3,057 μmol kg−1), MNX (hexahydro-1-nitroso-3,5-dinitro-1,3,5-triazine; 155 μmol kg−1), and traces of NDAB (3.8 μmol kg−1). The detection of the last in real soil provided the first experimental evidence for the occurrence of natural attenuation that involved ring cleavage of RDX. When we incubated the soil with strain DN22, both RDX and MNX (but not HMX) degraded and produced NDAB (388 ± 22 μmol kg−1) in 5 days. Subsequent incubation of the soil with the fungus led to the removal of NDAB, with the liberation of nitrous oxide (N2O). In cultures with the fungus alone NDAB degraded to give a stoichiometric amount of N2O. To determine C stoichiometry, we first generated [14C]NDAB in situ by incubating [14C]RDX with strain DN22, followed by incubation with the fungus. The production of 14CO2 increased from 30 (DN22 only) to 76% (fungus). Experiments with pure enzymes revealed that manganese-dependent peroxidase rather than lignin peroxidase was responsible for NDAB degradation. The detection of NDAB in contaminated soil and its effective mineralization by the fungus P. chrysosporium may constitute the basis for the development of bioremediation technologies.  相似文献   

14.
15.
The effects of exogenously supplied isoprene on chlorophyll fluorescence characteristics were examined in leaf discs of four isoprene-emitting plant species, kudzu (Pueraria lobata [Willd.] Ohwi.), velvet bean (Mucuna sp.), quaking aspen (Populus tremuloides Michx.), and pussy willow (Salix discolor Muhl). Isoprene, supplied to the leaves at either 18 μL L−1 in compressed air or 21 μL L−1 in N2, had no effect on the temperature at which minimal fluorescence exhibited an upward inflection during controlled increases in leaf-disc temperature. During exposure to 1008 μmol photons m−2 s−1 in an N2 atmosphere, 21 μL L−1 isoprene had no effect on the thermally induced inflection of steady-state fluorescence. The maximum quantum efficiency of photosystem II photochemistry decreased sharply as leaf-disc temperature was increased; however, this decrease was unaffected by exposure of leaf discs to 21 μL L−1 isoprene. Therefore, there were no discernible effects of isoprene on the occurrence of symptoms of high-temperature damage to thylakoid membranes. Our data do not support the hypothesis that isoprene enhances leaf thermotolerance.  相似文献   

16.
The effects of temperature, over a range of 5 to 30 C, on phytotoxicity of simazine, atrazine, propazine, prometryne, prometone, and ipazine to young Pinus resinosa seedlings were investigated in growth chambers. Herbicides were applied to the soil surface and then mixed into the soil before pine seeds were planted. Development of recently germinated seedlings was then studied for 7 weeks. High temperatures greatly accelerated herbicide toxicity, but the effects of temperature varied greatly among herbicides. Atrazine and simazine were more toxic than other herbicides tested at all temperatures. Toxicity of simazine and atrazine was apparent early, whereas effects of propazine, prometryne, prometone, and ipazine were somewhat delayed. After 7 weeks maximum dry-weight production of shoots under each herbicide treatment and control occurred at 20 C, with some decreases noted at lower temperatures and marked decreases at progressively higher ones. At 20 C final seedling dry weights following treatment with simazine or atrazine were only one-third as high as in control plants. Growth was also reduced in lesser amounts by propazine, prometryne, prometone, or ipazine. Variations in phytotoxicity of different triazine herbicides appeared to be related more to their structural differences than their solubilities. Under the constant environmental conditions of the experiments, toxicity symptoms in plants treated with triazine herbicides appeared more rapidly and decisively than in previous field experiments under fluctuating environments. The influence of high temperatures in enhancing triazine toxicity appeared to involve complex interactions of physiological activity of plants and temperature effects on herbicide uptake.  相似文献   

17.
trans-2′-Carboxybenzalpyruvate hydratase-aldolase was purified from a phenanthrene-degrading bacterium, Nocardioides sp. strain KP7, and characterized. The purified enzyme was found to have molecular masses of 38 kDa by sodium dodecyl sulfate-polyacrylamide gel electrophoresis and 113 kDa by gel filtration chromatography. Thus, the homotrimer of the 38-kDa subunit constituted an active enzyme. The Km and kcat values of this enzyme for trans-2′-carboxybenzalpyruvate were 50 μM and 13 s−1, respectively. trans-2′-Carboxybenzalpyruvate was transformed to 2-carboxybenzaldehyde and pyruvate by the action of this enzyme. The structural gene for this enzyme was cloned and sequenced; the length of this gene was 996 bp. The deduced amino acid sequence of this enzyme exhibited homology to those of trans-2′-hydroxybenzalpyruvate hydratase-aldolases from Pseudomonas putida PpG7 and Pseudomonas sp. strain C18.  相似文献   

18.
In soil, the way biotic parameters impact the relationship between bacterial diversity and function is still unknown. To understand these interactions better, we used RNA-based stable-isotope probing to study the diversity of active atrazine-degrading bacteria in relation to atrazine degradation and to explore the impact of earthworm-soil engineering with respect to this relationship. Bulk soil, burrow linings and earthworm casts were incubated with 13C-atrazine. The pollutant degradation was quantified by liquid chromatography–mass spectrometry for 8 days, whereas active atrazine degraders were identified at 2 and 8 days by sequencing the 16S ribosomal RNA in the 13C-RNA fractions from the three soil microsites. An original diversity of atrazine degraders was found. Earthworm soil engineering greatly modified the taxonomic composition of atrazine degraders with dominance of α-, β- and γ-proteobacteria in burrow linings and of Actinobacteria in casts. Earthworm soil bioturbation increased the γ-diversity of atrazine degraders over the soil microsites generated. Atrazine degradation was enhanced in burrow linings in which primary atrazine degraders, closely related to Pelomonas aquatica, were detected only 2 days after atrazine addition. Atrazine degradation efficiency was not linearly related to the species richness of degraders but likely relied on keystone species. By enhancing soil heterogeneity, earthworms sustained high phylogenetic bacterial diversity and exerted a biotic control on the bacterial diversity–function relationships. Our findings call for future investigations to assess the ecological significance of biotic controls on the relationships between diversity and function on ecosystem properties and services (for example, soil detoxification) at larger scales.  相似文献   

19.
The motility of the purple sulfur bacterium Marichromatium gracile was investigated under different light regimes in a gradient capillary setup with opposing oxygen and sulfide gradients. The gradients were quantified with microsensors, while the behavior of swimming cells was studied by video microscopy in combination with a computerized cell tracking system. M. gracile exhibited photokinesis, photophobic responses, and phobic responses toward oxygen and sulfide. The observed migration patterns could be explained solely by the various phobic responses. In the dark, M. gracile formed an ~500-μm-thick band at the oxic-anoxic interface, with a sharp border toward the oxic zone always positioned at ~10 μM O2. Flux calculations yielded a molar conversion ratio Stot/O2 of 2.03:1 (Stot = [H2S] + [HS] + [S2−]) for the sulfide oxidation within the band, indicating that in darkness the bacteria oxidized sulfide incompletely to sulfur stored in intracellular sulfur globules. In the light, M. gracile spread into the anoxic zone while still avoiding regions with >10 μM O2. The cells also preferred low sulfide concentrations if the oxygen was replaced by nitrogen. A light-dark transition experiment demonstrated a dynamic interaction between the chemical gradients and the cell's metabolism. In darkness and anoxia, M. gracile lost its motility after ca. 1 h. In contrast, at oxygen concentrations of >100 μM with no sulfide present the cells remained viable and motile for ca. 3 days both in light and darkness. Oxygen was respired also in the light, but respiration rates were lower than in the dark. Observed aggregation patterns are interpreted as effective protection strategies against high oxygen concentrations and might represent first stages of biofilm formation.  相似文献   

20.
Atrazine is a widely used herbicide with great environmental concern due to its high potential to contaminate soil and waters. An atrazine-degrading bacterial strain HB-6 was isolated from industrial wastewater and the 16S rRNA gene sequencing identified HB-6 as a Bacillus subtilis. PCR assays indicated that HB-6 contained atrazine-degrading genes trzN, atzB and atzC. The strain HB-6 was capable of utilizing atrazine and cyanuric acid as a sole nitrogen source for growth and even cleaved the s-triazine ring and mineralized atrazine. The strain demonstrated a very high efficiency of atrazine biodegradation with a broad optimum pH and temperature ranges and could be enhanced by cooperating with other bacteria, suggesting its huge potential for remediation of atrazine-contaminated sites. To our knowledge, there are few Bacillus subtilis strains reported that can mineralize atrazine, therefore, the present work might provide some new insights on atrazine remediation.  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