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 31 毫秒
1.
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   

2.
Purified 2,4-dichlorophenoxyacetic acid (2,4-D)/(alpha)-ketoglutarate dioxygenase (TfdA) was shown to use 4-nitrophenoxyacetic acid (K(infm) = 0.89 (plusmn) 0.04 mM, k(infcat) [catalytic constant] = 540 (plusmn) 10 min(sup-1)), producing intensely yellow 4-nitrophenol. This reagent was used to develop a rapid, continuous, colorimetric assay for the detection of TfdA and analogous activities in 2,4-D-degrading bacterial cells and extracts.  相似文献   

3.
 The first step in the degradation of 2,4-dichlorophenoxyacetic acid (2,4-D) by Ralstonia eutropha JMP134 is catalyzed by the α-ketoglutarate (α-KG)-dependent dioxygenase TfdA. Previously, EPR and ESEEM studies on inactive Cu(II)-substituted TfdA suggested a mixture of nitrogen/oxygen coordination with two imidazole-like ligands. Differences between the spectra for Cu TfdA and α-KG- and 2,4-D-treated samples were interpreted as a rearrangement of the g–tensor principal axis system. Herein, we report the use of X-ray absorption spectroscopy (XAS) to further characterize the metal coordination environment of Cu TfdA as well as that in the active, wild-type Fe(II) enzyme. The EXAFS data are interpreted in terms of four N/O ligands (two imidazole-like) in the Cu TfdA sample and six N/O ligands (one or two imidazole-like) in the Fe TfdA sample. Addition of α-KG results in no significant structural change in coordination for Cu or Fe TfdA. However, addition of 2,4-D results in a decrease in the number of imidazole ligands in both Cu and Fe TfdA. Since this change is seen both in the Fe and Cu EXAFS, loss of one histidine ligand upon 2,4-D addition best describes the phenomenon. These XAS data clearly demonstrate that changes occur in the atomic environment of the metallocenter upon substrate binding. Received: 3 July 1998 / Accepted: 13 October 1998  相似文献   

4.
Sphingomonas herbicidovorans MH was able to completely degrade both enantiomers of the chiral herbicide dichlorprop [(RS)-2-(2,4-dichlorophenoxy)propanoic acid], with preferential degradation of the (S) enantiomer over the (R) enantiomer. These results are in agreement with the recently reported enantioselective degradation of mecoprop [(RS)-2-(4-chloro-2-methylphenoxy)propanoic acid] by this bacterium (C. Zipper, K. Nickel, W. Angst, and H.-P. E. Kohler, Appl. Environ. Microbiol. 62:4318–4322, 1996). Uptake of (R)-dichlorprop, (S)-dichlorprop, and 2,4-D (2,4-dichlorophenoxyacetic acid) was inducible. Initial uptake rates of cells grown on the respective substrate showed substrate saturation kinetics with apparent affinity constants (Kt) of 108, 93, and 117 μM and maximal velocities (Vmax) of 19, 10, and 21 nmol min−1 mg of protein−1 for (R)-dichlorprop, (S)-dichlorprop, and 2,4-D, respectively. Transport of (R)-dichlorprop, (S)-dichlorprop, and 2,4-D was completely inhibited by various uncouplers and by nigericin but was only marginally inhibited by valinomycin and by the ATPase inhibitor N,N′-dicyclohexylcarbodiimine. Experiments on the substrate specificity of the putative transport systems revealed that (R)-dichlorprop uptake was inhibited by (R)-mecoprop but not by (S)-mecoprop, (S)-dichlorprop, or 2,4-D. On the other hand, the (S)-dichlorprop transport was inhibited by (S)-mecoprop but not by (R)-mecoprop, (R)-dichlorprop, or 2,4-D. These results provide evidence that the first step in the degradation of dichlorprop, mecoprop, and 2,4-D by S. herbicidovorans is active transport and that three inducible, proton gradient-driven uptake systems exist: one for (R)-dichlorprop and (R)-mecoprop, another for (S)-dichlorprop and (S)-mecoprop, and a third for 2,4-D.  相似文献   

5.
Two novel genes, rdpA and sdpA, encoding the enantiospecific α-ketoglutarate dependent dioxygenases catalyzing R,S-dichlorprop cleavage in Delftia acidovorans MC1 were identified. Significant similarities to other known genes were not detected, but their deduced amino acid sequences were similar to those of other α-ketoglutarate dioxygenases. RdpA showed 35% identity with TauD of Pseudomonas aeruginosa, and SdpA showed 37% identity with TfdA of Ralstonia eutropha JMP134. The functionally important amino acid sequence motif HX(D/E)X23-26(T/S)X114-183HX10-13R/K, which is highly conserved in group II α-ketoglutarate-dependent dioxygenases, was present in both dichlorprop-cleaving enzymes. Transposon mutagenesis of rdpA inactivated R-dichlorprop cleavage, indicating that it was a single-copy gene. Both rdpA and sdpA were located on the plasmid pMC1 that also carries the lower pathway genes. Sequencing of a 25.8-kb fragment showed that the dioxygenase genes were separated by a 13.6-kb region mainly comprising a Tn501-like transposon. Furthermore, two copies of a sequence similar to IS91-like elements were identified. Hybridization studies comparing the wild-type plasmid and that of the mutant unable to cleave dichlorprop showed that rdpA and sdpA were deleted, whereas the lower pathway genes were unaffected, and that deletion may be caused by genetic rearrangements of the IS91-like elements. Two other dichlorprop-degrading bacterial strains, Rhodoferax sp. strain P230 and Sphingobium herbicidovorans MH, were shown to carry rdpA genes of high similarity to rdpA from strain MC1, but sdpA was not detected. This suggested that rdpA gene products are involved in the degradation of R-dichlorprop in these strains.  相似文献   

6.
Twenty-two 2,4-dichlorophenoxyacetic acid (2,4-D)-degrading bacterial isolates were collected from agricultural soils at three sites in China. Sequence analysis of the 16S rRNA genes indicated that the isolates were phylogenetically grouped into four categories: Ochrobactrum anthropi, in the Alpha- class of the phylum Proteobacteria (3 out of 22 isolates), Cupriavidus sp., of the Betaproteobacteria (3 out of 22), Pseudomonas sp. and Stenotrophomonas sp., which are Gammaproteobacteria (7 out of 22), and Bacillus sp., of the phylum Firmicutes (9 out of 22). Primers were designed to amplify the conserved domain of tfdA, which is known to be involved in the degradation of 2,4-D. Results showed that the tfdA genes of all 22 strains were most similar to that of Cupriavidus necator JMP134, which belongs to the 2,4-D/α-ketoglutarate dioxygenase TfdA protein family, indicating that the JMP134-type tfdA gene is likely to be almost universal among the 2,4-D-degrading bacteria isolated from China. Degradation abilities of these 22 strains were investigated in assays using 2,4-D as the sole source of carbon and energy. Thirteen strains degraded >60 % of the available 2,4-D (500 mg l?1) over a 1-week incubation period, while a further nine Bacillus sp. strains degraded 50–81 % of the available 2,4-D. None of these nine strains degraded other selected herbicides, such as mecoprop, 2-methyl-4-chlorophenoxyacetic acid, quizalofop, and fluroxypyr. This is the first report of 2,4-D-degradation by Bacilli.  相似文献   

7.
Phenoxyalkanoic compounds are used worldwide as herbicides. Cupriavidus necator JMP134(pJP4) catabolizes 2,4-dichlorophenoxyacetate (2,4-D) and 4-chloro-2-methylphenoxyacetate (MCPA), using tfd functions carried on plasmid pJP4. TfdA cleaves the ether bonds of these herbicides to produce 2,4-dichlorophenol (2,4-DCP) and 4-chloro-2-methylphenol (MCP), respectively. These intermediates can be degraded by two chlorophenol hydroxylases encoded by the tfdBI and tfdBII genes to produce the respective chlorocatechols. We studied the specific contribution of each of the TfdB enzymes to the 2,4-D/MCPA degradation pathway. To accomplish this, the tfdBI and tfdBII genes were independently inactivated, and growth on each chlorophenoxyacetate and total chlorophenol hydroxylase activity were measured for the mutant strains. The phenotype of these mutants shows that both TfdB enzymes are used for growth on 2,4-D or MCPA but that TfdBI contributes to a significantly higher extent than TfdBII. Both enzymes showed similar specificity profiles, with 2,4-DCP, MCP, and 4-chlorophenol being the best substrates. An accumulation of chlorophenol was found to inhibit chlorophenoxyacetate degradation, and inactivation of the tfdB genes enhanced the toxic effect of 2,4-DCP on C. necator cells. Furthermore, increased chlorophenol production by overexpression of TfdA also had a negative effect on 2,4-D degradation by C. necator JMP134 and by a different host, Burkholderia xenovorans LB400, harboring plasmid pJP4. The results of this work indicate that codification and expression of the two tfdB genes in pJP4 are important to avoid toxic accumulations of chlorophenols during phenoxyacetic acid degradation and that a balance between chlorophenol-producing and chlorophenol-consuming reactions is necessary for growth on these compounds.  相似文献   

8.
A bioreporter was made containing a tfdRPDII-luxCDABE fusion in a modified mini-Tn5 construct. When it was introduced into the chromosome of Ralstonia eutropha JMP134, the resulting strain, JMP134-32, produced a sensitive bioluminescent response to 2,4-dichlorophenoxyacetic acid (2,4-D) at concentrations of 2.0 μM to 5.0 mM. This response was linear (R2 = 0.9825) in the range of 2.0 μM to 1.1 × 102 μM. Saturation occurred at higher concentrations, with maximal bioluminescence occurring in the presence of approximately 1.2 mM 2,4-D. A sensitive response was also recorded in the presence of 2,4-dichlorophenol at concentrations below 1.1 × 102 μM; however, only a limited bioluminescent response was recorded in the presence of 3-chlorobenzoic acid at concentrations below 1.0 mM. A significant bioluminescent response was also recorded when strain JMP134-32 was incubated with soils containing aged 2,4-D residues.  相似文献   

9.
α-Ketoglutarate-dependent (R)-dichlorprop dioxygenase (RdpA) and α-ketoglutarate-dependent (S)-dichlorprop dioxygenase (SdpA), which are involved in the degradation of phenoxyalkanoic acid herbicides in Sphingomonas herbicidovorans MH, were expressed and purified as His6-tagged fusion proteins from Escherichia coli BL21(DE3)(pLysS). RdpA and SdpA belong to subgroup II of the α-ketoglutarate-dependent dioxygenases and share the specific motif HXDX24TX131HX10R. Amino acids His-111, Asp-113, and His-270 and amino acids His-102, Asp-104, and His 257 comprise the 2-His-1-carboxylate facial triads and were predicted to be involved in iron binding in RdpA and SdpA, respectively. RdpA exclusively transformed the (R) enantiomers of mecoprop [2-(4-chloro-2-methylphenoxy)propanoic acid] and dichlorprop [2-(2,4-dichlorophenoxy)propanoic acid], whereas SdpA was specific for the (S) enantiomers. The apparent Km values were 99 μM for (R)-mecoprop, 164 μM for (R)-dichlorprop, and 3 μM for α-ketoglutarate for RdpA and 132 μM for (S)-mecoprop, 495 μM for (S)-dichlorprop, and 20 μM for α-ketoglutarate for SdpA. Both enzymes had high apparent Km values for oxygen; these values were 159 μM for SdpA and >230 μM for RdpA, whose activity was linearly dependent on oxygen at the concentration range measured. Both enzymes had narrow cosubstrate specificity; only 2-oxoadipate was able to replace α-ketoglutarate, and the rates were substantially diminished. Ferrous iron was necessary for activity of the enzymes, and other divalent cations could not replace it. Although the results of growth experiments suggest that strain MH harbors a specific 2,4-dichlorophenoxyacetic acid-converting enzyme, tfdA-, tfdAα-, or cadAB-like genes were not discovered in a screening analysis in which heterologous hybridization and PCR were used.  相似文献   

10.
Plasmid-mediated bioaugmentation was demonstrated using sequencing batch reactors (SBRs) for enhancing 2,4-dichlorophenoxyacetic acid (2,4-D) removal by introducing Cupriavidus necator JMP134 and Escherichia coli HB101 harboring 2,4-D-degrading plasmid pJP4. C. necator JMP134(pJP4) can mineralize and grow on 2,4-D, while E. coli HB101(pJP4) cannot assimilate 2,4-D because it lacks the chromosomal genes to degrade the intermediates. The SBR with C. necator JMP134(pJP4) showed 100 % removal against 200 mg/l of 2,4-D just after its introduction, after which 2,4-D removal dropped to 0 % on day 7 with the decline in viability of the introduced strain. The SBR with E. coli HB101(pJP4) showed low 2,4-D removal, i.e., below 10 %, until day 7. Transconjugant strains of Pseudomonas and Achromobacter isolated on day 7 could not grow on 2,4-D. Both SBRs started removing 2,4-D at 100 % after day 16 with the appearance of 2,4-D-degrading transconjugants belonging to Achromobacter, Burkholderia, Cupriavidus, and Pandoraea. After the influent 2,4-D concentration was increased to 500 mg/l on day 65, the SBR with E. coli HB101(pJP4) maintained stable 2,4-D removal of more than 95 %. Although the SBR with C. necator JMP134(pJP4) showed a temporal depression of 2,4-D removal of 65 % on day 76, almost 100 % removal was achieved thereafter. During this period, transconjugants isolated from both SBRs were mainly Achromobacter with high 2,4-D-degrading capability. In conclusion, plasmid-mediated bioaugmentation can enhance the degradation capability of activated sludge regardless of the survival of introduced strains and their 2,4-D degradation capacity.  相似文献   

11.
Cell extracts of Sphingomonas herbicidovorans MH grown on (R)-mecoprop contained an enzyme activity that selectively converted (R)-mecoprop to 4-chloro-2-methylphenol, whereas extracts of cells grown on (S)-mecoprop contained an enzyme activity selective for the S enantiomer. Both reactions were dependent on alpha-ketoglutarate and ferrous ions. Besides 4-chloro-2-methylphenol, pyruvate and succinate were detected as products of the reactions. Labeling experiments with (18)O2 revealed that both enzyme activities catalyzed a dioxygenation reaction. One of the oxygen atoms of pyruvate and one of the oxygen atoms of succinate were derived from molecular oxygen. Analysis of cell extracts obtained from cells grown on different substrates by sodium dodecyl sulfate-polyacrylamide gel electrophoresis showed that growth on (R)-mecoprop and (S)-mecoprop caused the appearance of prominent protein bands at 34 and 32 kDa, respectively. Both protein bands were present when cells grew on the racemic mixture. The results demonstrate that S. herbicidovorans initiated the degradation of each enantiomer of mecoprop by a specific alpha-ketoglutarate-dependent dioxygenase. By comparing conversion rates of various phenoxy herbicides, we confirmed that the two enzyme activities were distinct from that of TfdA, which catalyzes the first step in the degradation of 2,4-dichlorophenoxyacetic acid in Ralstonia eutropha JMP134.  相似文献   

12.
Combined cell suspensions of the 2,4,5-trichlorophenoxyacetic acid (2,4,5-T)-metabolizing organism Pseudomonas cepacia AC1100, and the 2,4-dichlorophenoxyacetic acid (2,4-D)-metabolizing organism Alcaligenes eutrophus JMP134 were shown to effectively degrade either of these compounds provided as single substrates. These combined cell suspensions, however, poorly degraded mixtures of the two compounds provided at the same concentrations. Growth and viability studies revealed that such mixtures of 2,4-D and 2,4,5-T were toxic to AC1100 alone and to combinations of AC1100 and JMP134. High-pressure liquid chromatography analyses of culture supernatants of AC1100 incubated with 2,4-D and 2,4,5-T revealed the accumulation of chlorohydroquinone as an apparent dead-end catabolite of 2,4-D and the subsequent accumulation of both 2,4-dichlorophenol and 2,4,5-trichlorophenol. JMP134 cells incubated in the same medium did not catabolize 2,4,5-T and were also inhibited in initiating 2,4-D catabolism. A new derivative of strain AC1100 was constructed by the transfer into this organism of the 2,4-D-degradative plasmid pJP4 from strain JMP134. This new strain, designated RHJ1, was shown to efficiently degrade mixtures of 2,4-D and 2,4,5-T through the simultaneous metabolism of these compounds.  相似文献   

13.
Abstract: Twenty-five 2,4-dichlorophenoxyacetic acid (2,4-D) degrading bacteria from geographically diverse locations and presenting various degrees of similarity or no similarity to the tfdA and tfdB genes from Alcaligenes eutrophus JMP134 were analysed by PCR-RFLP (restriction length fragment polymorphism). Primers for the 2,4-D etherase gene were derived by sequence alignment of the tfdA genes from A. eutrophus JMP134 and Burkholderia sp. RASC. Primers for the 2,4-dichlorophenolhydroxylase gene were based on the tfdB gene sequence from A. eutrophus JMP134 by taking codon degeneration and variations in amino acid residue sequences into consideration. PCR amplification using the tfdA primer set produced fragments of 0.3 kb from 17 strains which showed varying degrees of similarity to the tfdA gene probe from A. eutrophus JMP134. Significant variations in the gene sequences were confirmed by PCR-RFLP analysis. DNA amplification using the tfdB primer set produced a 1.1 kb fragment from 19 strains. Amongst them, two did not show any similarity to the tfdB gene probe. The size and restriction pattern of the products obtained from A. eutrophus JMP134 were in accordance with the expected size calculated from the A. eutrophus tfdA and tfdB gene sequence and their theoretical PCR-RFLP patterns. Some strains which did not amplify using the tfdA primer set did however amplify with the tfdB primer set. These results suggest the independent evolution of these two genes in the construction of the 2,4-D metabolic pathway. Our tfdA and tfdB primer sets could be used for the detection of similar sequences in bacteria and soils. Moreover, PCR-RFLP patterns could also be used to select subsets of strains for sequencing to study the phylogeny of the tfdA and tfdB genes.  相似文献   

14.
To isolate enantioselective α-hydroxyacid dehydrogenases (α-HADHs), a high-throughput screening method was established. 2,4-Dinitrophenylhydrazine solution forms a red-brown complex with ketoacid produced during the α-HADH-mediated oxidation of α-hydroxyacid. The complex can be easily quantified by spectrophotometric measurement at 458?nm. The enantioselectivity of α-HADH in each strain can be measured with this colorimetric method using (R)- and (S)-α-hydroxyacid concurrently as substrates to evaluate the apparent enantioselectivity (E app). The E app closely matches the value of true enantioselectivity (E true) determined by HPLC analysis. With this method, a total of 34 stains harboring enantioselective α-HADHs were selected from 526 potential α-HADH-producing microorganisms. Pseudomonas aeruginosa displayed the highest (S)-enantioselective α-HADH activity. This strain appears promising for potential application in industry to produce (R)-α-hydroxyacids. The method described herein represents a useful tool for the high-throughput isolation of enantioselective α-HADHs.  相似文献   

15.
We characterized the gene required to initiate the degradation of 2,4-dichlorophenoxyacetic acid (2,4-D) by the soil bacterium Burkholderia sp. strain TFD6, which hybridized to the tfdA gene of the canonical 2,4-D catabolic plasmid pJP4 under low-stringency conditions. Cleavage of the ether bond of 2,4-D by cell extracts of TFD6 proceeded by an (alpha)-ketoglutarate-dependent reaction, characteristic of TfdA (F. Fukumori and R. P. Hausinger, J. Bacteriol. 175:2083-2086, 1993). The TFD6 tfdA gene was identified in a recombinant plasmid which complemented a tfdA transposon mutant of TFD6 created by chromosomal insertion of Tn5. The plasmid also expressed TfdA activity in Escherichia coli DH5(alpha), as evidenced by enzyme assays with cell extracts. Sequence analysis of the tfdA gene and flanking regions from strain TFD6 showed 99.5% similarity to a tfdA gene cloned from the chromosome of a different Burkholderia species (strain RASC) isolated from a widely separated geographical area. This chromosomal gene has 77.2% sequence identity to tfdA from plasmid pJP4 (Y. Suwa, W. E. Holben, and L. J. Forney, abstr. Q-403, in Abstracts of the 94th General Meeting of the American Society for Microbiology 1994.). The tfdA homologs cloned from strains TFD6 and RASC are the first chromosomally encoded 2,4-D catabolic genes to be reported. The occurrence of highly similar tfdA genes in different bacterial species suggests that this chromosomal gene can be horizontally transferred.  相似文献   

16.
Phenoxyalkanoic compounds are used worldwide as herbicides. Cupriavidus necator JMP134(pJP4) catabolizes 2,4-dichlorophenoxyacetate (2,4-D) and 4-chloro-2-methylphenoxyacetate (MCPA), using tfd functions carried on plasmid pJP4. TfdA cleaves the ether bonds of these herbicides to produce 2,4-dichlorophenol (2,4-DCP) and 4-chloro-2-methylphenol (MCP), respectively. These intermediates can be degraded by two chlorophenol hydroxylases encoded by the tfdB(I) and tfdB(II) genes to produce the respective chlorocatechols. We studied the specific contribution of each of the TfdB enzymes to the 2,4-D/MCPA degradation pathway. To accomplish this, the tfdB(I) and tfdB(II) genes were independently inactivated, and growth on each chlorophenoxyacetate and total chlorophenol hydroxylase activity were measured for the mutant strains. The phenotype of these mutants shows that both TfdB enzymes are used for growth on 2,4-D or MCPA but that TfdB(I) contributes to a significantly higher extent than TfdB(II). Both enzymes showed similar specificity profiles, with 2,4-DCP, MCP, and 4-chlorophenol being the best substrates. An accumulation of chlorophenol was found to inhibit chlorophenoxyacetate degradation, and inactivation of the tfdB genes enhanced the toxic effect of 2,4-DCP on C. necator cells. Furthermore, increased chlorophenol production by overexpression of TfdA also had a negative effect on 2,4-D degradation by C. necator JMP134 and by a different host, Burkholderia xenovorans LB400, harboring plasmid pJP4. The results of this work indicate that codification and expression of the two tfdB genes in pJP4 are important to avoid toxic accumulations of chlorophenols during phenoxyacetic acid degradation and that a balance between chlorophenol-producing and chlorophenol-consuming reactions is necessary for growth on these compounds.  相似文献   

17.
Cleavage of the ether bond of chlorophenoxyalkanoate herbicides is catalyzed by an α‐ketoglutarate‐linked dioxygenase (TfdA). In this step, α‐ketoglutarate is decarboxylated to succinate and must be regenerated for continual substrate cleavage. Limitations in herbicide degradation are to be expected in the case of a shortage of α‐ketoglutarate. Such a situation was simulated and studied with Delftia (formerly Comamonas) acidovorans MC1 and Rhodoferax sp. P230, which constitutively express etherolytic dioxygenase activity by excreting 2,4‐dichlorophenol (DCP) as a dead‐end product. The results showed that 2,4‐dichlorophenoxyacetate (2,4‐D) could hardly be cleaved under these conditions which is attributed to the inability to regenerate α‐ketoglutarate from the cleavage products, i.e. succinate and glyoxylate [1 ]. With pyruvate, in contrast, liberated as the oxidized alkanoic acid from the cleavage of (RS )‐2‐(2,4‐dichlorophenoxy)propionate (2,4‐DP), the regeneration of α ‐ketoglutarate seems to be guaranteed from succinate as resulted from the utilization of 2,4‐DP to a considerable amount under these conditions. The extent was limited, however, which was apparently caused by the accumulation of DCP. Continual cleavage of 2,4‐DP could be demonstrated in the presence of Ochrobactrum sp. K2‐14, which functions as a DCP‐consuming strain. Addition of extra metabolites, i.e. α‐ketoglutarate or other readily metabolizable substrates, improved the cleavage of the herbicides. This was most pronounced with 2,4‐D that was found now to be also utilized to a considerable extent. Conversely, the cleavage of the herbicides (2,4‐DP) was reduced and ultimately ceased with cells depleted by starvation of the pool of metabolites. Again, this deficit could be restored by adding α‐ketoglutarate. The limitations in utilizing phenoxyalkanoate herbicides are discussed in terms of pseudo‐recalcitrance owing to deficits in metabolites (α‐ketoglutarate) rather than enzyme activity (TfdA).  相似文献   

18.
2,4-Dichlorophenoxyacetate/α-ketoglutarate dioxygenase (TfdA), the first enzyme in the catabolic pathway for the herbicide 2,4-dichlorophenoxyacetic acid (2,4-D), oxidizes α-ketoglutarate (α-kG) to CO2 and succinate while hydroxylating 2,4-D to yield an unstable hemiacetal that decomposes into 2,4-dichlorophenol and glyoxylate. In an effort to extend the potential biotechnological utility of this enzyme, a variety of non-phenoxyacetate compounds were examined as potential substrates. 2-Naphthoxyacetic acid was the best alternative substrate tested, followed by benzofuran-2-carboxylic acid, 2,4-dichlorocinnamic acid, 2-chlorocinnamic acid, 1-naphthoxyacetic acid, and 4-chlorocinnamic acid. TfdA appeared to oxidize the olefin bond of the cinnamic acids and benzofuran-2-carboxylate to form the corresponding epoxides. Whole cells were observed to also catalyze a TfdA-dependent oxidation of 2,4-dichlorocinnamic acid. Based on the ability of TfdA to metabolize chlorinated cinnamic acids, we speculate that tfdA-like sequences present in 2,4-D non-degrading natural isolates may function in metabolism of substituted cinnamic acids. These results support the use of TfdA and related enzymes in the specific oxidation of non-phenoxyacetate substrates.  相似文献   

19.
Cupriavidus necator (formerly Ralstonia eutropha) JMP134, harbouring the catabolic plasmid pJP4, is the best-studied 2,4-dichlorophenoxyacetic acid (2,4-D) herbicide degrading bacterium. A study of the survival and catabolic performance of strain JMP134 in agricultural soil microcosms exposed to high levels of 2,4-D was carried out. When C. necator JMP134 was introduced into soil microcosms, the rate of 2,4-D removal increased only slightly. This correlated with the poor survival of the strain, as judged by 16S rRNA gene terminal restriction fragment length polymorphism (T-RFLP) profiles, and the semi-quantitative detection of the pJP4-borne tfdA gene sequence, encoding the first step in 2,4-D degradation. After 3 days of incubation in irradiated soil microcosms, the survival of strain JMP134 dramatically improved and the herbicide was completely removed. The introduction of strain JMP134 into native soil microcosms did not produce detectable changes in the structure of the bacterial community, as judged by 16S rRNA gene T-RFLP profiles, but provoked a transient increase of signals putatively corresponding to protozoa, as indicated by 18S rRNA gene T-RFLP profiling. Accordingly, a ciliate able to feed on C.␣necator JMP134 could be isolated after soil enrichment. In␣native soil microcosms, C. necator JMP134 survived better than Escherichia coli DH5α (pJP4) and similarly to Pseudomonas putida KT2442 (pJP4), indicating that species specific factors control the survival of strains harbouring pJP4. The addition of cycloheximide to soil microcosms strongly improved survival of these three strains, indicating that the eukaryotic microbiota has a strong negative effect in bioaugmentation with catabolic bacteria.  相似文献   

20.
2,4-Dimethylphenoxyacetic acid and 2,4-dimethylphenol are not growth substrates for Alcaligenes eutrophus JMP 134 although being cooxidized by 2,4-dichlorophenoxyacetate grown cells. None of the relevant catabolic pathways were induced by the dimethylphenoxyacetate. 3,5-Dimethylcatechol is not subject to metacleavage. The alternative ortho-eleavage is also unproductive and gives rise to (+)-4-carboxymethyl-2,4-dimethylbut-2-en-4-olide as a dead-end metabolite. High yields of this metabolite were obtained with the mutant Alcaligenes eutrophys JMP 134-1 which constitutively expresses the genes of 2,4-dichlorophenoxyacetic acid metabolism.  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