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 15 毫秒
1.
Burkholderia cepacia R34 mineralizes 2,4-dinitrotoluene via an oxidative pathway. The initial steps in the degradative pathway lead to formation of 2,4,5-trihydroxytoluene, which serves as the substrate for the ring cleavage dioxygenase. The trihydroxylated substrate differs from the usual substituted catechols found in pathways for aromatic compound degradation. To determine whether the characteristics of the trihydroxytoluene oxygenase reflect the unusual ring cleavage substrate of the 2,4-dinitrotoluene pathway, the gene encoding trihydroxytoluene oxygenase (dntD) was cloned and sequenced, and ring cleavage activity determined from recombinant bacteria carrying the cloned gene. The findings were compared to the trihydroxytoluene oxygenase from Burkholderia sp. strain DNT and to other previously described ring cleavage dioxygenases. The comparison revealed that only 60% identity was shared by the two trihydroxytoluene oxygenases, but the amino acid residues involved in cofactor binding, catalysis, and protein folding were conserved in the DntD sequence. The enzyme catalyzed meta-fission of trihydroxytoluene as well as the substrate analogues 1,2,4-benzenetriol, catechol, 3-methylcatechol, 4-methylcatechol, 3-chlorocatechol, 4-chlorocatechol and 2,3-dihydroxybiphenyl. However, results from enzyme assays indicated a strong preference for trihydroxytoluene, implying that it was the native substrate for the enzyme. The apparent enzyme specificity, its similarity to the trihydroxytoluene oxygenase from Burkholderia sp. strain DNT, and the distant genetic relationship to other ring cleavage enzymes suggest that dntD evolved expressly to carry out trihydroxytoluene transformation.  相似文献   

2.
The gene (vgb) encoding the hemoglobin (VHb) of Vitreoscilla sp. was cloned into a broad host range vector and stably transformed into Burkholderia (formerly Pseudomonas) sp. strain DNT, which is able to degrade and metabolize 2,4-dinitrotoluene (DNT). Vgb was stably maintained and expressed in functional form in this recombinant strain (YV1). When growth of YV1, in both tryptic soy broth and minimal salts broth containing DNT and yeast extract, was compared with that of the untransformed strain, YV1 grew significantly better on a cell mass basis (A(600)) and reached slightly higher maximum viable cell numbers. YV1 also had roughly twice the respiration as strain DNT on a cell mass basis, and in DNT-containing medium, YV1 degraded DNT faster than the untransformed strain. YV1 cells pregrown in medium containing DNT plus succinate showed the fastest degradation: 100% of the initial 200 ppm DNT was removed from the medium within 3 days.  相似文献   

3.
Biodegradation of 2,4-dinitrotoluene by a Pseudomonas sp.   总被引:11,自引:0,他引:11  
Previous studies of the biodegradation of nonpolar nitroaromatic compounds have suggested that microorganisms can reduce the nitro groups but cannot cleave the aromatic ring. We report here the initial steps in a pathway for complete biodegradation of 2,4-dinitrotoluene (DNT) by a Pseudomonas sp. isolated from a four-member consortium enriched with DNT. The Pseudomonas sp. degraded DNT as the sole source of carbon and energy under aerobic conditions with stoichiometric release of nitrite. During induction of the enzymes required for growth on DNT, 4-methyl-5-nitrocatechol (MNC) accumulated transiently in the culture fluid when cells grown on acetate were transferred to medium containing DNT as the sole carbon and energy source. Conversion of DNT to MNC in the presence of 18O2 revealed the simultaneous incorporation of two atoms of molecular oxygen, which demonstrated that the reaction was catalyzed by a dioxygenase. Fully induced cells degraded MNC rapidly with stoichiometric release of nitrite. The results indicate an initial dioxygenase attack at the 4,5 position of DNT with the concomitant release of nitrite. Subsequent reactions lead to complete biodegradation and removal of the second nitro group as nitrite.  相似文献   

4.
Biodegradation of 2,4-dinitrotoluene by a Pseudomonas sp.   总被引:5,自引:4,他引:5       下载免费PDF全文
Previous studies of the biodegradation of nonpolar nitroaromatic compounds have suggested that microorganisms can reduce the nitro groups but cannot cleave the aromatic ring. We report here the initial steps in a pathway for complete biodegradation of 2,4-dinitrotoluene (DNT) by a Pseudomonas sp. isolated from a four-member consortium enriched with DNT. The Pseudomonas sp. degraded DNT as the sole source of carbon and energy under aerobic conditions with stoichiometric release of nitrite. During induction of the enzymes required for growth on DNT, 4-methyl-5-nitrocatechol (MNC) accumulated transiently in the culture fluid when cells grown on acetate were transferred to medium containing DNT as the sole carbon and energy source. Conversion of DNT to MNC in the presence of 18O2 revealed the simultaneous incorporation of two atoms of molecular oxygen, which demonstrated that the reaction was catalyzed by a dioxygenase. Fully induced cells degraded MNC rapidly with stoichiometric release of nitrite. The results indicate an initial dioxygenase attack at the 4,5 position of DNT with the concomitant release of nitrite. Subsequent reactions lead to complete biodegradation and removal of the second nitro group as nitrite.  相似文献   

5.
Continuous flow sand column bioreactor experiments were conducted to investigate the effect of 2,4-dinitrotoluene (DNT) concentration (i.e. DNT loading rate) and influent dissolved oxygen (DO) concentration on aerobic biodegradation of DNT by wild type (DNT) and recombinant (YV1) Burkholderia sp., the latter containing plasmid pSC160 which carries the gene (vgb) encoding the hemoglobin (VHb) from the bacterium Vitreoscilla. The experiments were conducted in two continuous flow packed bed sand column bioreactors, one growing the wild type strain and the other growing YV1. Under oxygen-rich feed conditions (6.8 mg DO/L in the feed) with an influent DNT concentration of 99.6 mg/L (DNT loading rate approximately = 9.2 mg/m2/day), the effluent DNT concentration from the wild type bioreactor reached 0.7 mg DNT/L in 40 days whereas it was less than 0.2 mg DNT/L for the YV1 bioreactor in about 25 days. When influent DNT concentration was increased to 214 mg/L (DNT loading rate approximately = 20.3 mg/m2/day) while maintaining the same influent DO level of 6.8 mg/L, the effluent DNT concentration increased to about 5 mg/L for the wild type bioreactor whereas it was maintained at less than 0.2 mg/L for the YV1 bioreactor. Additionally, when influent DO was reduced from 6.8 mg/L to 3.1 mg/L while the influent DNT concentration remained at 214 mg/L, the effluent DNT concentration increased to more than 20 mg/L for the wild type bioreactor but up to only 1.7 mg/L for the YV1 bioreactor. A subsequent increase of influent DO back to 6.6 mg/L reduced the effluent DNT concentration to about 5 mg/L for the wild type bioreactor and to 0.10-0.19 mg/L for the YV1 bioreactor. These results confirm the utility of vgb technology to enhance biodegradation of aromatic compounds under hypoxic conditions and also that this enhancement can be maintained over extended periods of time as evidenced by plasmid stability in Burkholderia YV1.  相似文献   

6.
The degradation of 2,4-dinitrotoluene (DNT) by Pseudomonas sp. strain DNT is initiated by a dioxygenase attack to yield 4-methyl-5-nitrocatechol (MNC) and nitrite. Subsequent oxidation of MNC by a monooxygenase results in the removal of the second molecule of nitrite, and further enzymatic reactions lead to ring fission. Initial studies on the molecular basis of DNT degradation in this strain revealed the presence of three plasmids. Mitomycin-derived mutants deficient in either DNT dioxygenase only or DNT dioxygenase and MNC monooxygenase were isolated. Plasmid profiles of mutant strains suggested that the mutations resulted from deletions in the largest plasmid. Total plasmid DNA partially digested by EcoRI was cloned into a broad-host-range cosmid vector, pCP13. Recombinant clones containing genes encoding DNT dioxygenase, MNC monooxygenase, and 2,4,5-trihydroxytoluene oxygenase were characterized by identification of reaction products and the ability to complement mutants. Subcloning analysis suggests that the DNT dioxygenase is a multicomponent enzyme system and that the genes for the DNT pathway are organized in at least three different operons.  相似文献   

7.
Saturation mutagenesis of the 2,4-dinitrotoluene dioxygenase (DDO) of Burkholderia cepacia R34 at position valine 350 of the DntAc alpha-subunit generated mutant V350F with significantly increased activity towards o-nitrophenol (47 times), m-nitrophenol (34 times), and o-methoxyphenol (174 times) as well as an expanded substrate range that now includes m-methoxyphenol, o-cresol, and m-cresol (wild-type DDO had no detectable activity for these substrates). Another mutant, V350M, also displays increased activity towards o-nitrophenol (20 times) and o-methoxyphenol (162 times) as well as novel activity towards o-cresol. Products were synthesized using whole Escherichia coli TG1 cells expressing the recombinant R34 dntA loci from pBS(Kan)R34, and the initial rates of product formation were determined at 1 mM substrate by reverse-phase high-pressure liquid chromatography. V350F produced both nitrohydroquinone at a rate of 0.75 +/- 0.15 nmol/min/mg of protein and 3-nitrocatechol at a rate of 0.069 +/- 0.001 nmol/min/mg of protein from o-nitrophenol, 4-nitrocatechol from m-nitrophenol at 0.29 +/- 0.02 nmol/min/mg of protein, methoxyhydroquinone from o-methoxyphenol at 2.5 +/- 0.6 nmol/min/mg of protein, methoxyhydroquinone from m-methoxyphenol at 0.55 +/- 0.02 nmol/min/mg of protein, both methylhydroquinone at 1.52 +/- 0.02 nmol/min/mg of protein and 2-hydroxybenzyl alcohol at 0.74 +/- 0.05 nmol/min/mg of protein from o-cresol, and methylhydroquinone at 0.43 +/- 0.1 nmol/min/mg of protein from m-cresol. V350M produced both nitrohydroquinone at a rate of 0.33 nmol/min/mg of protein and 3-nitrocatechol at 0.089 nmol/min/mg of protein from o-nitrophenol, methoxyhydroquinone from o-methoxyphenol at 2.4 nmol/min/mg of protein, methylhydroquinone at 1.97 nmol/min/mg of protein and 2-hydroxybenzyl alcohol at 0.11 nmol/min/mg of protein from o-cresol. The DDO variants V350F and V350M also exhibited 10-fold-enhanced activity towards naphthalene (8 +/- 2.6 nmol/min/mg of protein), forming (1R,2S)-cis-1,2-dihydro-1,2-dihydroxynaphthalene. Hence, mutagenesis of wild-type DDO through active-site engineering generated variants with relatively high rates toward a previously uncharacterized class of substituted phenols for the nitroarene dioxygenases; seven previously uncharacterized substrates were evaluated for wild-type DDO, and four novel monooxygenase-like products were found for the DDO variants V350F and V350M (methoxyhydroquinone, methylhydroquinone, 2-hydroxybenzyl alcohol, and 3-nitrocatechol).  相似文献   

8.
2,4-Dinitrotoluene (2,4-DNT) and 2,6-DNT are priority pollutants, and 2,4-DNT dioxygenase of Burkholderia sp. strain DNT (DDO) catalyzes the initial oxidation of 2,4-DNT to form 4-methyl-5-nitrocatechol and nitrite but has significantly less activity on other dinitrotoluenes and nitrotoluenes (NT). Hence, oxidation of 2,3-DNT, 2,4-DNT, 2,5-DNT, 2,6-DNT, 2NT, and 4NT were enhanced here by performing saturation mutagenesis on codon I204 of the alpha subunit (DntAc) of DDO and by using a membrane agar plate assay to detect catechol formation. Rates of degradation were quantified both by the formation of nitrite and by the formation of the intermediates with high performance liquid chromatography. The degradation of both 2,3-DNT and 2,5-DNT were achieved for the first time (no detectable activity with the wild-type enzyme) using whole Escherichia coli TG1 cells expressing DDO variants DntAc I204L and I204Y (0.70 +/- 0.03 and 0.22 +/- 0.02 nmol/min/mg protein for 2,5-DNT transformation, respectively). DDO DntAc variant I204L also transformed both 2,6-DNT and 2,4-DNT 2-fold faster than wild-type DDO (0.8 +/- 0.6 nmol/min/mg protein and 4.7 +/- 0.5 nmol/min/mg protein, respectively). Moreover, the activities of DDO for 2NT and 4NT were also enhanced 3.5-fold and 8-fold, respectively. Further, DntAc variant I204Y was also discovered with comparable rate enhancements for the substrates 2,4-DNT, 2,6-DNT, and 2NT but not 4NT. Sequencing information obtained during this study indicated that the 2,4-DNT dioxygenases of Burkholderia sp. strain DNT and B. cepacia R34 are more closely related than originally reported. This is the first report of engineering an enzyme for enhanced degradation of nitroaromatic compounds and the first report of degrading 2,5-DNT.  相似文献   

9.
10.
lux -marked biosensors for assessing the toxicity and bioremediation potential of polluted environments may complement traditional chemical techniques. lux CDABE genes were introduced into the chromosome of the 2,4-dichlorophenol (2,4-DCP)-mineralizing bacterium, Burkholderia sp. RASC c2, by biparental mating using the Tn 4431 system. Experiments revealed that light output was constitutive and related to cell biomass concentration during exponential growth. The transposon insertion was stable and did not interrupt 2,4-DCP-degradative genes, and expression of lux CDABE did not constitute a metabolic burden to the cell. A bioluminescence response was detectable at sublethal 2,4-DCP concentrations: at < 10.26 μg ml−1, bioluminescence was stimulated (e.g. 218% of control), but at concentrations > 60 μg ml−1 it declined to < 1%. Investigating the effect of [14C]-2,4-DCP concentration on the evolution of 14CO2 revealed that, for initial concentrations of 2.5–25 μg ml−1, ≈55% of the added 14C was mineralized after 24 h compared with < 1% at 50 and 100 μg ml−1. Inhibition of 2,4-DCP mineralization between 25 and 50 μg ml−1 corresponded well to the EC50 value (33.83 μg ml−1) obtained from bioluminescence inhibition studies. lux -marked RASC c2 may therefore be used as a functionally (i.e. 2,4-DCP degrader) and environmentally relevant biosensor of toxicity and biodegradation inhibition.  相似文献   

11.
The findings of previous studies indicate that the genes required for metabolism of the pesticide 2,4-dichlorophenoxyacetic acid (2,4-D) are typically encoded on broad-host-range plasmids. However, characterization of plasmid-cured strains of Burkholderia sp. strain RASC, as well as mutants obtained by transposon mutagenesis, suggested that the 2,4-D catabolic genes were located on the chromosome of this strain. Mutants of Burkholderia strain RASC unable to degrade 2,4-D (2,4-D- strains) were obtained by insertional inactivation with Tn5. One such mutant (d1) was shown to have Tn5 inserted in tfdARASC, which encodes 2,4-D/alpha-ketoglutarate dioxygenase. This is the first reported example of a chromosomally encoded tfdA. The tfdARASC gene was cloned from a library of wild-type Burkholderia strain RASC DNA and shown to express 2,4-D/alpha-ketoglutarate dioxygenase activity in Escherichia coli. The DNA sequence of the gene was determined and shown to be similar, although not identical, to those of isofunctional genes from other bacteria. Moreover, the gene product (TfdARASC) was purified and shown to be similar in molecular weight, amino-terminal sequence, and reaction mechanism to the canonical TfdA of Alcaligenes eutrophus JMP134. The data presented here indicate that tfdA genes can be found on the chromosome of some bacterial species and suggest that these catabolic genes are rather mobile and may be transferred by means other than conjugation.  相似文献   

12.
A bacterium capable of utilizing fenitrothion (O,O-dimethyl O-4-nitro-m-tolyl phosphorothioate) as a sole carbon source was isolated from fenitrothion-treated soil. This bacterium was characterized taxonomically as being a member of the genus Burkholderia and was designated strain NF100. NF100 first hydrolyzed an organophosphate bond of fenitrothion, forming 3-methyl-4-nitrophenol, which was further metabolized to methylhydroquinone. The ability to degrade fenitrothion was found to be encoded on two plasmids, pNF1 and pNF2.  相似文献   

13.
The genetic systems responsible for naphthalene and phenanthrene catabolism have been analyzed in the five strains of Burkholderia sp. isolated from soil samples (West Siberia) contaminated by heavy residual fuel and in the laboratory collection strain Burkholderia sp. BS3702 isolated from soil samples of the coke gas plant (Vidnoe, Moscow oblast). The results of this work demonstrate that naphthalene and phenanthrene degradation in the above strains is encoded by the sequences not homologous to the classical nah genes of pseudomonades. In the Burkholderia sp. BS3702 strain, the initial stages of phenanthrene degradation and the subsequent stages of salicylate degradation are controlled by the sequences of different evolutionary descent (phn and nag genes).  相似文献   

14.
Smith AR  Beadle CA 《Biodegradation》2008,19(5):669-681
Burkholderia cepacia 2a inducibly degraded 2,4-dichlorophenoxyacetate (2,4-D) sequentially via 2,4-dichlorophenol, 3,5-dichlorocatechol, 2,4-dichloromuconate, 2-chloromuconolactone and 2-chloromaleylacetate. Cells grown on nutrient agar or broth grew on 2,4-D-salts only if first passaged on 4-hydroxybenzoate- or succinate-salts agar. Buffered suspensions of 4-hydroxybenzoate-grown cells did not adapt to 2,4-D or 3,5-dichlorocatechol, but responded to 2,4-dichlorophenol at concentrations <0.4 mM. Uptake of 2,4-dichlorophenol by non-induced cells displayed a type S (cooperative uptake) uptake isotherm in which the accelerated uptake of the phenol began before the equivalent of a surface monolayer had been adsorbed, and growth inhibition corresponded with the acquisition of 2.2-fold excess of phenol required for the establishment of the monolayer. No evidence of saturation was seen even at 2 mM 2,4-dichlorophenol, possibly due to absorption by intracellular poly-beta-hydroxybutyrate inclusions. With increasing concentration, 2,4-dichlorophenol caused progressive cell membrane damage and, sequentially, leakage of intracellular K(+), P(i), ribose and material absorbing light at 260 nm (presumed nucleotide cofactors), until at 0.4 mM, protein synthesis and enzyme induction were forestalled. Growth of non-adapted cells was inhibited by 0.35 mM 2,4-dichlorophenol and 0.25 mM 3,5-dichlorocatechol; the corresponding minimum bacteriocidal concentrations were 0.45 and 0.35 mM. Strain 2a grew in chemostat culture on carbon-limited media containing 2,4-D, with an apparent growth yield coefficient of 0.23, and on 2,4-dichlorophenol. Growth on 3,5-dichlorocatechol did not occur without a supplement of succinate, probably due to accumulation of toxic quantities of quinonoid and polymerisation products. Cells grown on these compounds were active towards all three, but not when grown on other substrates. The enzymes of the pathway therefore appeared to be induced by 3,5-dichlorocatechol or some later metabolite. A possible reason is offered for the environmental persistence of 2,4,5-trichlorophenoxyacetic acid (2,4,5-T).  相似文献   

15.
A lactonase hydrolyzing (R)-5-oxo-2-tetrahydrofurancarboxylic acid to D-alpha-hydroxyglutaric acid was purified 170-fold with 2% recovery to near homogeneity from crude extracts of Burkholderia sp. R-711, which had been isolated as a bacterium able to assimilate (R)-5-oxo-2-tetrahydrofurancarboxylic acid. The molecular mass was estimated to be 33 kDa by gel filtration. The purified preparation migrated as a single band of molecular mass 38 kDa upon SDS-PAGE. The maximum activity was observed at pH 7.0-8.0 and 35-40 degrees C. The enzyme required no added cofactors or metal ions; the activity was inhibited to 60-100% by SH-blocking reagents, but was not affected by metal-chelating reagents. The enzyme showed lower activity and affinity toward (S)-5-oxo-2-tetrahydrofurancarboxylic acid, but did not act on other natural and synthetic lactones tested.  相似文献   

16.
Burkholderia sp. strain LB400 is one of the most potent aerobic polychlorobiphenyl (PCB)-degrading microorganisms that have been characterized. Its PCB-dioxygenating activity originates predominantly or exclusively from the biphenyl dioxygenase encoded by its bph gene cluster. Analysis of the dioxygenation products of several di- to pentachlorinated biphenyls formed by this enzyme revealed a complex dependence of the regiospecificity and the yield of dioxygenation on the substitution patterns of both the oxidized and the nonoxidized rings. No dioxygenolytic attack involving chlorinated meta or para carbons was observed. Therefore, the ability of the enzyme to hydroxylate chlorinated carbons appears to be limited to the ortho position. However, it is not limited to monochlorinated rings, as evidenced by dioxygenation of the 2, 4-disubstituted ring at carbons 2 and 3. This site of attack is strikingly different from that of the 2,5-dichlorinated ring, which has been shown to be dihydroxylated at positions 3 and 4 (J. D. Haddock, J. R. Horton, and D. T. Gibson, J. Bacteriol. 177:20-26, 1995). These results demonstrate that a second substituent of ortho-chlorinated rings crucially influences the site of dioxygenation at this ring and thereby determines whether or not the initial chlorobiphenyl oxidation product is further metabolized through the bph-encoded pathway. The 2,4-dichlorinated ring can alternatively be attacked at carbons 5 and 6. The preferred site crucially depends on the substitution pattern of the other ring. The formation of more than a single dioxygenation product was found predominantly with congeners that contain two chlorinated rings, both of which are similarly prone to dioxygenation or one is substituted only at carbon 3.  相似文献   

17.
The successive application of distinct pesticides, or mixtures of them, is a frequent practice that could adversely affect the microbial species inhabiting soil and aquatic ecosystems. The ability of soil or aquatic microbiota to degrade a pesticide could be affected by the presence of another. If the degradation rate of the first compound is inhibited, its dissipation half‐life in the environment could be hazardously enlarged. Few studies have been made to quantify the impact on the biodegradation rate of pesticides in soils or water by the presence of other pesticides. In this work, a method for assessing the effect of a pesticide on the biodegradation rate of another, measuring its effect on the biodegradation kinetics of a single bacterial strain is presented. The mathematical analysis is a powerful tool to study the stoichiometry and kinetics of microbial processes, which was used to evaluate independently, in detail, the effect of three pesticides (propanil, linuron, and dicamba) on the biodegradation kinetics of 2,4‐dichlorophenoxyacetic acid by a strain of Burkholderia sp. It was evidenced that linuron and dicamba caused a decay of more than 40% in the top instantaneous degradation rate of 2,4‐dichlorophenoxyacetic acid, while propanil showed a minimal effect.  相似文献   

18.
19.
Burkholderia sp. strain AK-5 utilized 4-aminophenol as the sole carbon, nitrogen, and energy source. A pathway for the metabolism of 4-aminophenol in strain AK-5 was proposed based on the identification of three key metabolites by gas chromatography-mass spectrometry analysis. Strain AK-5 converted 4-aminophenol to 1,2,4-trihydroxybenzene via 1,4-benzenediol. 1,2,4-Trihydroxybenzene 1,2-dioxygenase cleaved the benzene ring of 1,2,4-trihydroxybenzene to form maleylacetic acid. The enzyme showed a high dioxygenase activity only for 1,2,4-trihydroxybenzene, with K(m) and V(max) values of 9.6 micro M and 6.8 micro mol min(-1) mg of protein(-1), respectively.  相似文献   

20.
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