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.
Pseudomonas sp. strain ADP contains the genes, atzA, -B, and -C, that encode three enzymes which metabolize atrazine to cyanuric acid. Atrazine-catabolizing pure cultures isolated from around the world contain genes homologous to atzA, -B, and -C. The present study was conducted to determine whether the same genes are present in an atrazine-catabolizing bacterial consortium and how the genes and metabolism are subdivided among member species. The consortium contained four or more bacterial species, but two members, Clavibacter michiganese ATZ1 and Pseudomonas sp. strain CN1, collectively mineralized atrazine. C. michiganese ATZ1 released chloride from atrazine, produced hydroxyatrazine, and contained a homolog to the atzA gene that encoded atrazine chlorohydrolase. C. michiganese ATZ1 stoichiometrically metabolized hydroxyatrazine to N-ethylammelide and contained genes homologous to atzB and atzC, suggesting that either a functional AtzB or -C catalyzed N-isopropylamine release from hydroxyatrazine. C. michiganese ATZ1 grew on isopropylamine as its sole carbon and nitrogen source, explaining the ability of the consortium to use atrazine as the sole carbon and nitrogen source. A second consortium member, Pseudomonas sp. strain CN1, metabolized the N-ethylammelide produced by C. michiganese ATZ1 to transiently form cyanuric acid, a reaction catalyzed by AtzC. A gene homologous to the atzC gene of Pseudomonas sp. strain ADP was present, as demonstrated by Southern hybridization and PCR. Pseudomonas sp. strain CN1, but not C. michiganese, metabolized cyanuric acid. The consortium metabolized atrazine faster than did C. michiganese individually. Additionally, the consortium metabolized a much broader set of triazine ring compounds than did previously described pure cultures in which the atzABC genes had been identified. These data begin to elucidate the genetic and metabolic bases of catabolism by multimember consortia.  相似文献   

2.
Cyanuric acid hydrolase (AtzD) from Pseudomonas sp. strain ADP was purified to homogeneity. Of 22 cyclic amides and triazine compounds tested, only cyanuric acid and N-methylisocyanuric acid were substrates. Other cyclic amidases were found not to hydrolyze cyanuric acid. Ten bacteria that use cyanuric acid as a sole nitrogen source for growth were found to contain either atzD or trzD, but not both genes.  相似文献   

3.
4.
5.
The strain Pseudomonas sp. strain ADP is able to degrade atrazine as a sole nitrogen source and therefore needs a single source for both carbon and energy for growth. In addition to the typical C source for Pseudomonas, Na2-succinate, the strain can also grow with phenol as a carbon source. Phenol is oxidized to catechol by a multicomponent phenol hydroxylase. Catechol is degraded via the ortho pathway using catechol 1,2-dioxygenase. It was possible to stimulate the strain in order to degrade very high concentrations of phenol (1,000 mg/liter) and atrazine (150 mg/liter) simultaneously. With cyanuric acid, the major intermediate of atrazine degradation, as an N source, both the growth rate and the phenol degradation rate were similar to those measured with ammonia as an N source. With atrazine as an N source, the growth rate and the phenol degradation rate were reduced to ~35% of those obtained for cyanuric acid. This presents clear evidence that although the first three enzymes of the atrazine degradation pathway are constitutively present, either these enzymes or the uptake of atrazine is the bottleneck that diminishes the growth rate of Pseudomonas sp. strain ADP with atrazine as an N source. Whereas atrazine and cyanuric acid showed no significant toxic effect on the cells, phenol reduces growth and activates or induces typical membrane-adaptive responses known for the genus Pseudomonas. Therefore Pseudomonas sp. strain ADP is an ideal bacterium for the investigation of the regulatory interactions among several catabolic genes and stress response mechanisms during the simultaneous degradation of toxic phenolic compounds and a xenobiotic N source such as atrazine.  相似文献   

6.
Bacterial atrazine catabolism is initiated by the enzyme atrazine chlorohydrolase (AtzA) in Pseudomonas sp. strain ADP. Other triazine herbicides are metabolized by bacteria, but the enzymological basis of this is unclear. Here we begin to address this by investigating the catalytic activity of AtzA by using substrate analogs. Purified AtzA from Pseudomonas sp. strain ADP catalyzed the hydrolysis of an atrazine analog that was substituted at the chlorine substituent by fluorine. AtzA did not catalyze the hydrolysis of atrazine analogs containing the pseudohalide azido, methoxy, and cyano groups or thiomethyl and amino groups. Atrazine analogs with a chlorine substituent at carbon 2 and N-alkyl groups, ranging in size from methyl to t-butyl, all underwent dechlorination by AtzA. AtzA catalyzed hydrolytic dechlorination when one nitrogen substituent was alkylated and the other was a free amino group. However, when both amino groups were unalkylated, no reaction occurred. Cell extracts were prepared from five strains capable of atrazine dechlorination and known to contain atzA or closely homologous gene sequences: Pseudomonas sp. strain ADP, Rhizobium strain PATR, Alcaligenes strain SG1, Agrobacterium radiobacter J14a, and Ralstonia picketti D. All showed identical substrate specificity to purified AtzA from Pseudomonas sp. strain ADP. Cell extracts from Clavibacter michiganensis ATZ1, which also contains a gene homologous to atzA, were able to transform atrazine analogs containing pseudohalide and thiomethyl groups, in addition to the substrates used by AtzA from Pseudomonas sp. strain ADP. This suggests that either (i) another enzyme(s) is present which confers the broader substrate range or (ii) the AtzA itself has a broader substrate range.  相似文献   

7.
Pseudomonas strain ADP metabolizes the herbicide atrazine via three enzymatic steps, encoded by the genes atzABC, to yield cyanuric acid, a nitrogen source for many bacteria. Here, we show that five geographically distinct atrazine-degrading bacteria contain genes homologous to atzA, -B, and -C. The sequence identities of the atz genes from different atrazine-degrading bacteria were greater than 99% in all pairwise comparisons. This differs from bacterial genes involved in the catabolism of other chlorinated compounds, for which the average sequence identity in pairwise comparisons of the known members of a class ranged from 25 to 56%. Our results indicate that globally distributed atrazine-catabolic genes are highly conserved in diverse genera of bacteria.Atrazine [2-chloro-4-(ethylamino)-6-(isopropylamino)- 1,3,5-triazine] is a herbicide used for controlling broad-leaf and grassy weeds and is relatively persistent in soils (51). Atrazine and other s-triazine compounds have been detected in ground and surface waters at levels exceeding the Environmental Protection Agency’s maximum contaminant level of 3 ppb (30).Microbial populations exposed to synthetic chlorinated compounds, such as atrazine, often respond by producing enzymes that degrade these molecules. Most of our current understanding of the genes and enzymes involved in atrazine degradation derives from studies using Pseudomonas strain ADP, in which the first three enzymatic steps in atrazine degradation have been defined (6, 14, 15, 48). The genes atz A, -B, and -C, which encode these enzymes, have been cloned and sequenced. Atrazine chlorohydrolase (AtzA), hydroxyatrazine ethylaminohydrolase (AtzB), and N-isopropylammelide isopropylaminohydrolase (AtzC) sequentially convert atrazine to cyanuric acid (6, 14, 15, 48) (Fig. (Fig.1).1). Cyanuric acid and related compounds are catabolized by many soil bacteria (10, 11, 17, 24, 26, 61), and by Pseudomonas sp. ADP, to carbon dioxide and ammonia (35). This provides the evolutionary pressure for the atzA, -B, and -C genes to permit bacterial growth on the more than one billion pounds of atrazine that have been applied to soils globally (20). Here we used a knowledge of the atzA, -B, and -C gene sequences to investigate the presence of homologous genes in other atrazine-degrading bacteria. In this study, we report that five atrazine-degrading microorganisms, which were recently isolated from geographically separated sites exposed to atrazine, contained nearly identical atzA, -B, and -C genes. Open in a separate windowFIG. 1Pathway for atrazine catabolism to cyanuric acid in Pseudomonas sp. strain ADP.

Atrazine-catabolizing bacteria used in this study.

Until recently, attempts at isolating bacteria (18) or fungi (27) that completely degrade atrazine to carbon dioxide, ammonia, and chloride were unsuccessful. While several microorganisms were shown to dealkylate atrazine, they were unable to displace the chlorine atom (41, 54). Since 1994, several research groups have independently isolated atrazine-degrading bacteria that displaced the chlorine atom and mineralized atrazine (3, 7, 13, 35, 39, 46). Six of these bacterial cultures, listed in Table Table1,1, were studied here, and the Clavibacter strain had been investigated previously (13).

TABLE 1

Recently isolated atrazine-catabolizing bacteria
GenusStrainLocation where isolatedYr reported (reference)
PseudomonasaADPAgricultural-chemical dealership site, Little Falls, Minn.1995 (35)
RalstoniaaM91-3Agricultural soil, Ohio1995 (46, 55)
Mixed cultureBasel, Switzerland1995 (57)
ClavibacterAgricultural soil, Riverside, Calif.1996 (13)
AgrobacteriumJ14aAgricultural soil, Nebraska1996 (39)
NDb38/38Atrazine-contaminated soil, Indiana1996 (3)
AlcaligenesaSG1Industrial settling pond, San Gabriel, La.1997 (7)
Open in a separate windowaIsolate identity based on 16S rRNA sequence analysis. bND, not determined. 

Detection of atzA, -B, and -C homologs in atrazine-degrading microorganisms by PCR analysis.

Recently isolated atrazine-degrading bacteria were screened for the presence of DNA homologous to the Pseudomonas strain ADP atzABC genes, which encode enzymes transforming atrazine to cyanuric acid (Fig. (Fig.1).1). Total genomic DNA was isolated from each of these bacteria as described elsewhere (49), and the PCR technique was used to amplify sequences internal to the atzA, -B, and -C genes as described elsewhere (13). Custom primers were designed specifically for atzA (5′CCATGTGAACCAGATCCT3′ and 5′TGAAGCGTCCACATTACC3′), atzB (5′TCACCGGGGATGTCGCGGGC3′ and 5′CTCTCCCGCATGGCATCGGG3′), and atzC (5′GCTCACATGCAGGTACTCCA3′ and 5′GTACCATATCACCGTTTGCCA3′) by using the Primer Designer package, version 2.01 (Scientific and Educational Software, State Line, Pa.), and were synthesized by Gibco BRL (Gaithersburg, Md.). PCR fragments were amplified by using Taq DNA polymerase (Gibco BRL) (22) and were separated from primers on a 1.0% agarose gel. The results of these studies (Fig. (Fig.2)2) indicated that PCR amplification consistently produced DNA fragments of 0.5 kb for all organisms when the atzA or -B primers were used and fragments of 0.6 kb when the atzC primers were used. Open in a separate windowFIG. 2PCR analysis with primers designed to amplify internal regions of atzA (lanes 1 to 5), atzB (lanes 6 to 10), and atzC (lanes 11 to 15). The atrazine-degrading bacteria analyzed were Pseudomonas strain ADP (35) (lanes 1, 6, and 11), Alcaligenes strain SGI (7) (lanes 2, 7, and 12), Ralstonia strain M91-3 (46) (lanes 3, 8, and 13), Agrobacterium strain J14a (39) (lanes 4, 9, and 14), and isolate 38/38 (3) (lanes 5, 10, and 15). Values to the right of the gel are sizes (in kilobase pairs).Southern hybridization analyses were performed on the PCR-amplified DNA as described elsewhere (49) to confirm the presence of homologous DNA. We used a 0.6-kb ApaI/PstI fragment from pMD4 (15), a 1.5-kb BglII fragment from pATZB-2 (6), and a 2.0-kb EcoRI/AvaI fragment from pTD2.5 (48) as probes for atzA, -B, and -C genes, respectively. DNA probes were labeled with [α-32P]dCTP by using the Rediprime Random Primer Labeling Kit (Amersham Life Science, Arlington Heights, Ill.) according to the manufacturer’s instructions. Southern hybridization analyses, performed under stringent conditions, confirmed that each strain contained DNA homologous to atzA, -B, and -C (data not shown). With strain M91-3 and isolate 38/38, however, in addition to the expected 0.5-kb atzB PCR product (Fig. (Fig.2,2, lanes 8 and 10), a 1.2-kb fragment was also obtained. However, no hybridization to this fragment was seen with the atzB probe. Similar investigations showed that a mixed culture obtained from Switzerland (Table (Table1),1), capable of degrading atrazine, also contained DNA homologous to all three atz genes (12).As a negative control, bacteria known not to degrade atrazine were analyzed. PCR analyses were carried out with genomic DNA from the following randomly chosen laboratory strains: Rhodococcus chlorophenolicus (1), Flavobacterium sp. (47), Streptomyces coelicolor M145 (21), Amycolatopsis mediterranei (19), Agrobacterium strain A136 and strain A348 (A136/pTiA6NC) (60), Arthrobacter globiformis MN1 (45), Bradyrhizobium japonicum (33), Rhizobium sp. strain NGR 234 (44), Pseudomonas NRRLB12228, and Klebsiella pneumoniae 99 (16). None of these strains contained DNA that was amplified by PCR using the primers designed to identify the atzA, -B, or -C gene (data not shown).

DNA sequences of atzA, -B, and -C homologs in atrazine-degrading microorganisms.

DNAs amplified from the five strains in Table Table11 with the atzA, -B, and -C primers were purified from gel slices by using the GeneClean II System (Bio 101, Inc., Vista, Calif.) and sequenced with a PRISM Ready Reaction DyeDeoxy Terminator Cycle Sequencing kit (Perkin-Elmer Corp., Norwalk, Conn.) and an ABI model 373A DNA sequencer (Applied Biosystems, Foster City, Calif.). The GCG sequence analysis software package (Genetics Computer Group, Inc., Madison, Wis.) was used for all DNA and protein sequence comparisons and alignments. Table Table22 summarizes these data. The PCR-amplified genes were ≥99% identical to the Pseudomonas strain ADP atzA, -B, and -C genes in all pairwise comparisons of DNA sequences. This remarkable sequence identity suggested that each atz gene in the different genera was derived from a common ancestor and that they have diverged evolutionarily only to a limited extent.

TABLE 2

Sequence identities of atzABC homologs from different atrazine-degrading bacteria
Strain% DNA sequence identitya
atzAatzBatzC
Pseudomonas ADP100100100
Alcaligenes SG199.2100100
Ralstonia M91-399.0100100
Agrobacterium J14a99.1100100
Isolate 38/3899.310099.8
Open in a separate windowaDNA sequences obtained from each strain by using the ataA, -B, and -C primers were compared with the atzABC gene sequences from Pseudomonas strain ADP. A review of the literature on other bacterial catabolic pathways indicated a much greater degree of divergence when genes encoding enzymes for the catabolism of other commercially relevant chlorinated compounds were compared (Table (Table3).3). As with atrazine, multiple bacterial strains that catabolize 1,2-dichloroethane, chloroacetic acid, 2,4-dichlorophenoxyacetate, dichloromethane, and 4-chlorobenzoate have been isolated. A comparison of the gene sequences encoding the initiating reactions in the catabolism of each of those compounds revealed that sequence divergence was comparatively high. In pairwise comparisons within each gene class, the average sequence identities ranged from 25 to 56% (divergence was 46 to 75%). With the atzABC genes, by contrast, there is at most a 1% sequence difference within the sequenced gene region (Table (Table2).2). Moreover, the atzB sequences were completely identical, and the atzC genes diverged by only 1 bp in one of the five strains tested. This suggests that the atz genes recently arose from a single origin and have become distributed globally. Similarly, identical parathion hydrolase genes were isolated from two bacteria representing different genera and global locations (40, 52, 53).

TABLE 3

Sequence comparisons of isofunctional bacterial enzymes that catabolize chlorinated compounds
GeneEnzymeAverage % protein sequence identitya (no. of pairwise comparisons)References
dhlA, dhaAHaloalkane dehalogenase25.0 (1)23, 31
dehC, hadL, dehH, dehH1, dehH2, dhlB, dehCI, dehCII2-Haloacid dehalogenase36.6 ± 3.9 (36)5, 25, 28, 29, 42, 43, 50, 59
tfdA2,4-Dichlorophenoxyacetate monooxygenase43.2 ± 4.6 (21)b34, 37, 38, 56, 58
dcmADichloromethane dehalogenase56.0 (1)4, 32
atzAAtrazine chlorohydrolase98.6 ± 0.12 (15)cThis study
atzBHydroxyatrazine ethylaminohydrolase100 (10)cThis study
atzCN-Isopropylammelide isopropylaminohydrolase99.0 ± 0.43 (10)cThis study
Open in a separate windowaAll possible pairwise alignments of translated gene sequences were made. The average percent identity is the mean of the percent identity values for all pairwise alignments ± standard error of the mean. bIncludes full protein sequences as well as partial protein sequences of ≥100 amino acids. cSequence identity within a 0.5-kb PCR product for atzA and -B and within a 0.6-kb PCR product for atzC. Six sequences were analyzed for atzA, and five were analyzed for atzB and -C. The data presented here provide further support for previous studies suggesting that hydroxyatrazine in the environment derives from biological processes (36), and not solely from abiotic reactions (2, 9). The present data, and a recent report by Bouquard et al. (8), indicate that the gene encoding atrazine chlorohydrolase is widespread in the United States and Europe.Our observations argue for a single, recent evolutionary origin of the atz genes and their subsequent global distribution. We have recently localized the atzA, -B, and -C genes to a large, self-transmissible plasmid in Pseudomonas strain ADP (12), and possible mechanisms of transfer of the atzABC genes are currently under investigation.  相似文献   

8.
Pseudomonas sp. strain ADP utilizes the human-made s-triazine herbicide atrazine as the sole nitrogen source. The results reported here demonstrate that atrazine and the atrazine degradation intermediates N-isopropylammelide and cyanuric acid are chemoattractants for strain ADP. In addition, the nonmetabolized s-triazine ametryn was also an attractant. The chemotactic response to these s-triazines was not specifically induced during growth with atrazine, and atrazine metabolism was not required for the chemotactic response. A cured variant of strain ADP (ADP M13-2) was attracted to s-triazines, indicating that the atrazine catabolic plasmid pADP-1 is not necessary for the chemotactic response and that atrazine degradation and chemotaxis are not genetically linked. These results indicate that atrazine and related s-triazines are detected by one or more chromosomally encoded chemoreceptors in Pseudomonas sp. strain ADP. We demonstrated that Escherichia coli is attracted to the s-triazine compounds N-isopropylammelide and cyanuric acid, and an E. coli mutant lacking Tap (the pyrimidine chemoreceptor) was unable to respond to s-triazines. These data indicate that pyrimidines and triazines are detected by the same chemoreceptor (Tap) in E. coli. We showed that Pseudomonas sp. strain ADP is attracted to pyrimidines, which are the naturally occurring structures closest to triazines, and propose that chemotaxis toward s-triazines may be due to fortuitous recognition by a pyrimidine chemoreceptor in Pseudomonas sp. strain ADP. In competition assays, the presence of atrazine inhibited chemotaxis of Pseudomonas sp. strain ADP to cytosine, and cytosine inhibited chemotaxis to atrazine, suggesting that pyrimidines and s-triazines are detected by the same chemoreceptor.Atrazine [2-chloro-4-(N-ethylamino)-6-(N-isopropylamino)-1,3,5-s-triazine] is a human-made herbicide that is used worldwide to control broadleaf and grassy weeds. As one of the most heavily used herbicides in the United States, atrazine can be present in parts per million in agricultural runoffs (3), which exceeds the U.S. Environmental Protection Agency''s maximum allowable contaminant level of 3 ppb in ground and surface waters (13). Atrazine is persistent in soil (34) and was once considered nontoxic to animals. However, recent studies have shown that atrazine causes sexual abnormalities in frogs (21, 22, 50), reduced testosterone production in rats (53), and elevated levels of prostate cancer in workers at an atrazine-manufacturing factory (45). These studies suggest that there is cause for concern about atrazine residues in soil, groundwater, and surface waters.Several bacterial strains capable of mineralizing atrazine have been isolated (4, 27, 41, 49, 51, 52, 58). The best-studied atrazine-degrading strain, Pseudomonas sp. strain ADP (atrazine degrading pseudomonad), was isolated from an atrazine spill site in Minnesota (27). Strain ADP utilizes atrazine as a sole nitrogen source and mineralizes it in the process (27). The pathway of atrazine degradation in strain ADP has been characterized in detail (Fig. (Fig.1),1), and the genes encoding the six enzymes required for atrazine degradation have been cloned and sequenced (5, 7, 9, 10, 29, 42). The six genes are located in four distant locations on the atrazine catabolic plasmid (pADP-1) present in strain ADP (10, 29). atzA, atzB, and atzC, which encode the first three enzymes of the pathway, are constitutively expressed and highly conserved in atrazine-degrading bacteria isolated from geographically distinct locations (8, 11). Products of the atzDEF gene cluster catalyze the last three steps of atrazine degradation. This operon is divergently transcribed from atzR, the product of which has high homology to LysR-type regulatory proteins (29). AtzR and the inducer cyanuric acid are required for the expression of the atzDEF operon (14), and the operon is also subject to nitrogen control (15).Open in a separate windowFIG. 1.Pathway of atrazine degradation in Pseudomonas sp. strain ADP (reviewed in reference 55).In a study investigating the bioavailability of atrazine, Park et al. provided evidence that two atrazine-degrading strains, Pseudomonas sp. strain ADP and Agrobacterium radiobacter J14a, were chemotactically attracted to atrazine (38). Chemotaxis, the ability of motile bacteria to detect and respond to specific chemicals, can help bacteria find an optimal niche for their survival and growth and may play a role in the efficient degradation of pollutants in the environment (33, 37). Chemotaxis has been shown to enhance naphthalene biodegradation in both a heterogeneous aqueous system (30) and a non-aqueous-phase liquid system (24). In addition, a chemotactic naphthalene-degrading strain caused a higher rate of naphthalene desorption than was observed with nonchemotactic and nonmotile strains (24). Pseudomonas sp. strain ADP and recombinant strains expressing atz genes have been used to remove atrazine from soil in laboratory and field scale experiments (32, 48). If chemotaxis can enhance bioavailability in environments where the chemicals are sorbed to particles, the use of a motile chemotactic strain for bioremediation would be advantageous. Aside from the practical implications of atrazine chemotaxis, we are interested in understanding the evolution of a chemotactic response to a human-made chemical that was initially synthesized just 50 years ago (23). The results reported here indicate that Pseudomonas sp. strain ADP is chemotactically attracted to atrazine, atrazine metabolites, and the nonmetabolizable structural analog ametryn. The chemotactic response is not induced during growth with atrazine in strain ADP and does not require atrazine metabolism. We demonstrated that a single chemoreceptor (Tap) mediates chemotaxis to s-triazines and structurally similar pyrimidines in Escherichia coli. Additionally, we found that Pseudomonas sp. strain ADP is attracted to pyrimidines. In competition assays, cytosine inhibited atrazine chemotaxis, and vice versa. We therefore concluded that pyrimidines and s-triazines are detected by a single chemoreceptor in Pseudomonas sp. strain ADP.  相似文献   

9.
Strain DNS10 was the only member that could utilize atrazine as the sole nitrogen source for growth in an atrazine-degrading consortium which was isolated from black soil previously in our laboratory. It belongs to the genus Arthrobacter according to the sequence of 16S rRNA gene and is designated as Arthrobacter sp. DNS10. 16S rRNA gene phylogenetic analysis showed that strain DNS10 was located in a different evolutionary branch comparing with other Arthrobacter sp. atrazine-degrading strains. The degrading genes such as trzN, atzB and atzC harbored in strain DNS10 revealed high sequence similarity with those in Arthrobacter aurescens TC1 and Pseudomonas sp. ADP. These genes enabled the strain DNS10 to decompose atrazine to cyanuric acid. This was further proved by the results that the strain DNS10 (108 CFU mL−1) could degrade the whole atrazine (100 mg L−1) in the medium within 24 h at 30 °C and there was 66.13 ± 2.11 mg L−1 cyanuric acid accumulated at 24 h. These results imply that the strain DNS10 seems to be an excellent atrazine-degrading strain. Furthermore, this paper helps us in the better understanding of the strain evolution by comparing the metabolic ability and gene characteristics of strain DNS10 with other geographically distinct atrazine-degrading strains.  相似文献   

10.
TrzF, the allophanate hydrolase from Enterobacter cloacae strain 99, was cloned, overexpressed in the presence of a chaperone protein, and purified to homogeneity. Native TrzF had a subunit molecular weight of 65,401 and a subunit stoichiometry of α2 and did not contain significant levels of metals. TrzF showed time-dependent inhibition by phenyl phosphorodiamidate and is a member of the amidase signature protein family. TrzF was highly active in the hydrolysis of allophanate but was not active with urea, despite having been previously considered a urea amidolyase. TrzF showed lower activity with malonamate, malonamide, and biuret. The allophanate hydrolase from Pseudomonas sp. strain ADP, AtzF, was also shown to hydrolyze biuret slowly. Since biuret and allophanate are consecutive metabolites in cyanuric acid metabolism, the low level of biuret hydrolase activity can have physiological significance. A recombinant Escherichia coli strain containing atzD, encoding cyanuric acid hydrolase that produces biuret, and atzF grew slowly on cyanuric acid as a source of nitrogen. The amount of growth produced was consistent with the liberation of 3 mol of ammonia from cyanuric acid. In vitro, TrzF was shown to hydrolyze biuret to liberate 3 mol of ammonia. The biuret hydrolyzing activity of TrzF might also be physiologically relevant in native strains. E. cloacae strain 99 grows on cyanuric acid with a significant accumulation of biuret.  相似文献   

11.
Pseudomonas sp. strain ADP metabolizes atrazine to cyanuric acid via three plasmid-encoded enzymes, AtzA, AtzB, and AtzC. The first enzyme, AtzA, catalyzes the hydrolytic dechlorination of atrazine, yielding hydroxyatrazine. The second enzyme, AtzB, catalyzes hydroxyatrazine deamidation, yielding N-isopropylammelide. In this study, the third gene in the atrazine catabolic pathway, atzC, was cloned from a Pseudomonas sp. strain ADP cosmid library as a 25-kb EcoRI DNA fragment in Escherichia coli. The atzC gene was further delimited by functional analysis following transposon Tn5 mutagenesis and subcloned as a 2.0-kb EcoRI-AvaI fragment. An E. coli strain containing this DNA fragment expressed N-isopropylammelide isopropylamino hydrolase activity, metabolizing N-isopropylammelide stoichiometrically to cyanuric acid and N-isopropylamine. The 2.0-kb DNA fragment was sequenced and found to contain a single open reading frame of 1,209 nucleotides, encoding a protein of 403 amino acids. AtzC showed modest sequence identity of 29 and 25%, respectively, to cytosine deaminase and dihydroorotase, both members of an amidohydrolase protein superfamily. The sequence of AtzC was compared to that of E. coli cytosine deaminase in the regions containing the five ligands to the catalytically important metal for the protein. Pairwise comparison of the 35 amino acids showed 61% sequence identity and 85% sequence similarity. AtzC is thus assigned to the amidohydrolase protein family that includes cytosine deaminase, urease, adenine deaminase, and phosphotriester hydrolase. Similar sequence comparisons of the most highly conserved regions indicated that the AtzA and AtzB proteins also belong to the same amidohydrolase family. Overall, the data suggest that AtzA, AtzB, and AtzC diverged from a common ancestor and, by random events, have been reconstituted onto an atrazine catabolic plasmid.  相似文献   

12.
The gene loci fcs, encoding feruloyl coenzyme A (feruloyl-CoA) synthetase, ech, encoding enoyl-CoA hydratase/aldolase, and aat, encoding β-ketothiolase, which are involved in the catabolism of ferulic acid and eugenol in Pseudomonas sp. strain HR199 (DSM7063), were localized on a DNA region covered by two EcoRI fragments (E230 and E94), which were recently cloned from a Pseudomonas sp. strain HR199 genomic library in the cosmid pVK100. The nucleotide sequences of parts of fragments E230 and E94 were determined, revealing the arrangement of the aforementioned genes. To confirm the function of the structural genes fcs and ech, they were cloned and expressed in Escherichia coli. Recombinant strains harboring both genes were able to transform ferulic acid to vanillin. The feruloyl-CoA synthetase and enoyl-CoA hydratase/aldolase activities of the fcs and ech gene products, respectively, were confirmed by photometric assays and by high-pressure liquid chromatography analysis. To prove the essential involvement of the fcs, ech, and aat genes in the catabolism of ferulic acid and eugenol in Pseudomonas sp. strain HR199, these genes were inactivated separately by the insertion of omega elements. The corresponding mutants Pseudomonas sp. strain HRfcsΩGm and Pseudomonas sp. strain HRechΩKm were not able to grow on ferulic acid or on eugenol, whereas the mutant Pseudomonas sp. strain HRaatΩKm exhibited a ferulic acid- and eugenol-positive phenotype like the wild type. In conclusion, the degradation pathway of eugenol via ferulic acid and the necessity of the activation of ferulic acid to the corresponding CoA ester was confirmed. The aat gene product was shown not to be involved in this catabolism, thus excluding a β-oxidation analogous degradation pathway for ferulic acid. Moreover, the function of the ech gene product as an enoyl-CoA hydratase/aldolase suggests that ferulic acid degradation in Pseudomonas sp. strain HR199 proceeds via a similar pathway to that recently described for Pseudomonas fluorescens AN103.  相似文献   

13.
Pseudomonas C12B is able to degrade alkyl sulfates, alkylbenzene sulfonates, and linear alkanes and alkenes. Mitomycin C curing experiments and conjugation experiments demonstrated that the ability to utilize n-alkanes (C9–C12) and n-alkenes (C10 and C12) of medium chain length was plasmid-encoded. The plasmid was designated pDEC. Its size was estimated at several hundreds kb according to mobility in agarose gels. The plasmid did not confer resistance to the antibiotics tested. Analysis of alkylsulfatases P1 and P2 in original and cured strains confirmed that both enzymes are encoded by the chromosome. The ability of Pseudomonas C12B to utilize alkylbenzene sulfonates also appears to be encoded by the chromosome. pDEC could be transferred only to cured derivatives of Pseudomonas C12B, but not to strains of P. aeruginosa, P. putida, or Acinetobacter sp. Cured derivatives of Pseudomonas C12B could not serve as hosts for the broad host range plasmid CAM–OCT. The enzyme system encoded by the putative dec genes present on plasmid pDEC differs from the system coded by the alk genes of plasmid OCT in the size range of hydrocarbons preferentially used.  相似文献   

14.
Di- and trichloroisocyanuric acids are widely used as water disinfection agents, but cyanuric acid accumulates with repeated additions and must be removed to maintain free hypochlorite for disinfection. This study describes the development of methods for using a cyanuric acid-degrading enzyme contained within nonliving cells that were encapsulated within a porous silica matrix. Initially, three different bacterial cyanuric acid hydrolases were compared: TrzD from Acidovorax citrulli strain 12227, AtzD from Pseudomonas sp. strain ADP, and CAH from Moorella thermoacetica ATCC 39073. Each enzyme was expressed recombinantly in Escherichia coli and tested for cyanuric acid hydrolase activity using freely suspended or encapsulated cell formats. Cyanuric acid hydrolase activities differed by only a 2-fold range when comparing across the different enzymes with a given format. A practical water filtration system is most likely to be used with nonviable cells, and all cells were rendered nonviable by heat treatment at 70°C for 1 h. Only the CAH enzyme from the thermophile M. thermoacetica retained significant activity under those conditions, and so it was tested in a flowthrough system simulating a bioreactive pool filter. Starting with a cyanuric acid concentration of 10,000 μM, more than 70% of the cyanuric acid was degraded in 24 h, it was completely removed in 72 h, and a respike of 10,000 μM cyanuric acid a week later showed identical biodegradation kinetics. An experiment conducted with water obtained from municipal swimming pools showed the efficacy of the process, although cyanuric acid degradation rates decreased by 50% in the presence of 4.5 ppm hypochlorite. In total, these experiments demonstrated significant robustness of cyanuric acid hydrolase and the silica bead materials in remediation.  相似文献   

15.
In the bacterial degradation of steroid compounds, the enzymes initiating the breakdown of the steroid rings are well known, while the reactions for degrading steroid side chains attached to C-17 are largely unknown. A recent in vitro analysis with Pseudomonas sp. strain Chol1 has shown that the degradation of the C5 acyl side chain of the C24 steroid compound cholate involves the C22 intermediate 7α,12α-dihydroxy-3-oxopregna-1,4-diene-20S-carbaldehyde (DHOPDCA) with a terminal aldehyde group. In the present study, candidate genes with plausible functions in the formation and degradation of this aldehyde were identified. All deletion mutants were defective in growth with cholate but could transform it into dead-end metabolites. A mutant with a deletion of the shy gene, encoding a putative enoyl coenzyme A (CoA) hydratase, accumulated the C24 steroid (22E)-7α,12α-dihydroxy-3-oxochola-1,4,22-triene-24-oate (DHOCTO). Deletion of the sal gene, formerly annotated as the steroid ketothiolase gene skt, resulted in the accumulation of 7α,12α,22-trihydroxy-3-oxochola-1,4-diene-24-oate (THOCDO). In cell extracts of strain Chol1, THOCDO was converted into DHOPDCA in a coenzyme A- and ATP-dependent reaction. A sad deletion mutant accumulated DHOPDCA, and expression in Escherichia coli revealed that sad encodes an aldehyde dehydrogenase for oxidizing DHOPDCA to the corresponding acid 7α,12α-dihydroxy-3-oxopregna-1,4-diene-20-carboxylate (DHOPDC) with NAD+ as the electron acceptor. These results clearly show that the degradation of the acyl side chain of cholate proceeds via an aldolytic cleavage of an acetyl residue; they exclude a thiolytic cleavage for this reaction step. Based on these results and on sequence alignments with predicted aldolases from other bacteria, we conclude that the enzyme encoded by sal catalyzes this aldolytic cleavage.  相似文献   

16.
Polyhydroxyalkanoic acids (PHA) are carbon and energy storage polymers that accumulate in inclusion bodies in many bacteria and archaea in response to environmental conditions. This work presents the results of a study of PHA inclusion body-associated proteins and an analysis of their coding region in Bacillus megaterium 11561. A 7,917-bp fragment of DNA was cloned and shown to carry a 4,104-bp cluster of 5 pha genes, phaP, -Q, -R, -B, and -C. The phaP and -Q genes were shown to be transcribed in one orientation, each from a separate promoter, while immediately upstream, phaR, -B, and -C were divergently transcribed as a tricistronic operon. Transfer of this gene cluster to Escherichia coli and to a PhaC mutant of Pseudomonas putida gave a Pha+ phenotype in both strains. Translational fusions to the green fluorescent protein localized PhaP and PhaC to the PHA inclusion bodies in living cells. The data presented are consistent with the hypothesis that the extremely hydrophilic protein PhaP is a storage protein and suggests that PHA inclusion bodies are not only a source of carbon, energy, and reducing equivalents but are also a source of amino acids.  相似文献   

17.
Mobilizable shuttle plasmids containing the origin-of-transfer (oriT) region of plasmids F (IncFI), ColIb-P9 (IncI1), and RP4/RP1 (IncPα) were constructed to test the ability of the cognate conjugation system to mediate gene transfer from Escherichia coli to Saccharomyces cerevisiae. Only the Pα system caused detectable mobilization to yeast, giving peak values of 5 × 10−5 transconjugants per recipient cell in 30 min. Transfer of the shuttle plasmid required carriage of oriT in cis and the provision in trans of the Pα Tra1 core and Tra2 core regions. Genes outside the Tra1 core did not increase the mobilization efficiency. All 10 Tra2 core genes (trbB, -C, -D, -E, -F, -G, -H, -I, -J, and -L) required for plasmid transfer to E. coli K-12 were needed for transfer to yeast. To assess whether the mating-pair formation (Mpf) system or DNA-processing apparatus of the Pα conjugation system is critical in transkingdom transfer, an assay using an IncQ-based shuttle plasmid specifying its own DNA-processing system was devised. RP1 but not ColIb mobilized the construct to yeast, indicating that the Mpf complex determined by the Tra2 core genes plus traF is primarily responsible for the remarkable fertility of the Pα system in mediating gene transfer from bacteria to eukaryotes.  相似文献   

18.
s-Triazine ring compounds are common industrial chemicals: pesticides, resin intermediates, dyes, and explosives. The fate of these compounds in the environment is directly correlated with the ability of microbes to metabolize them. Microbes metabolize melamine and the triazine herbicides such as atrazine via enzyme-catalyzed hydrolysis reactions. Hydrolytic removal of substituents on the s-triazine ring is catalyzed by enzymes from the amidohydrolase superfamily and yields cyanuric acid as an intermediate. Cyanuric acid is hydrolytically processed to yield 3 mol each of ammonia and carbon dioxide. In those cases studied, the genes underlying the hydrolytic reactions are localized to large catabolic plasmids. One such plasmid, pADP-1 from Pseudomonas sp. ADP, has been completely sequenced and contains the genes for atrazine catabolism. Insertion sequence elements play a role in constructing different atrazine catabolic plasmids in different bacteria. Atrazine chlorohydrolase has been purified to homogeneity from two sources. Recombinant Escherichia coli strains expressing atrazine chlorohydrolase have been constructed and chemically cross-linked to generate catalytic particles used for atrazine remediation in soil. The method was used for cleaning up a spill of 1,000 pounds of atrazine to attain a level of herbicide acceptable to regulatory agencies.  相似文献   

19.
The vaoA gene from Penicillium simplicissimum CBS 170.90, encoding vanillyl alcohol oxidase, which also catalyzes the conversion of eugenol to coniferyl alcohol, was expressed in Escherichia coli XL1-Blue under the control of the lac promoter, together with the genes calA and calB, encoding coniferyl alcohol dehydrogenase and coniferyl aldehyde dehydrogenase of Pseudomonas sp. strain HR199, respectively. Resting cells of the corresponding recombinant strain E. coli XL1-Blue(pSKvaomPcalAmcalB) converted eugenol to ferulic acid with a molar yield of 91% within 15 h on a 50-ml scale, reaching a ferulic acid concentration of 8.6 g liter−1. This biotransformation was scaled up to a 30-liter fermentation volume. The maximum production rate for ferulic acid at that scale was 14.4 mmol per h per liter of culture. The maximum concentration of ferulic acid obtained was 14.7 g liter−1 after a total fermentation time of 30 h, which corresponded to a molar yield of 93.3% with respect to the added amount of eugenol. In a two-step biotransformation, E. coli XL1-Blue(pSKvaomPcalAmcalB) was used to produce ferulic acid from eugenol and, subsequently, E. coli(pSKechE/Hfcs) was used to convert ferulic acid to vanillin (J. Overhage, H. Priefert, and A. Steinbüchel, Appl. Environ. Microbiol. 65:4837-4847, 1999). This process led to 0.3 g of vanillin liter−1, besides 0.1 g of vanillyl alcohol and 4.6 g of ferulic acid liter−1. The genes ehyAB, encoding eugenol hydroxylase of Pseudomonas sp. strain HR199, and azu, encoding the potential physiological electron acceptor of this enzyme, were shown to be unsuitable for establishing eugenol bioconversion in E. coli XL1-Blue.  相似文献   

20.
Pseudomonas pseudoalcaligenes POB310(pPOB) and Pseudomonas sp. strains B13-D5(pD30.9) and B13-ST1(pPOB) were introduced into soil microcosms containing 3-phenoxybenzoic acid (3-POB) in order to evaluate and compare bacterial survival, degradation of 3-POB, and transfer of plasmids to a recipient bacterium. Strain POB310 was isolated for its ability to use 3-POB as a growth substrate; degradation is initiated by POB-dioxygenase, an enzyme encoded on pPOB. Strain B13-D5 contains pD30.9, a cloning vector harboring the genes encoding POB-dioxygenase; strain B13-ST1 contains pPOB. Degradation of 3-POB in soil by strain POB310 was incomplete, and bacterial densities decreased even under the most favorable conditions (100 ppm of 3-POB, supplementation with P and N, and soil water-holding capacity of 90%). Strains B13-D5 and B13-ST1 degraded 3-POB (10 to 100 ppm) to concentrations of <50 ppb with concomitant increases in density from 106 to 108 CFU/g (dry weight) of soil. Thus, in contrast to strain POB310, the modified strains had the following two features that are important for in situ bioremediation: survival in soil and growth concurrent with removal of an environmental contaminant. Strains B13-D5 and B13-ST1 also completely degraded 3-POB when the inoculum was only 30 CFU/g (dry weight) of soil. This suggests that in situ bioremediation may be effected, in some cases, with low densities of introduced bacteria. In pure culture, transfer of pPOB from strains POB310 and B13-ST1 to Pseudomonas sp. strain B13 occurred at frequencies of 5 × 10−7 and 10−1 transconjugant per donor, respectively. Transfer of pPOB from strain B13-ST1 to strain B13 was observed in autoclaved soil but not in nonautoclaved soil; formation of transconjugant bacteria was more rapid in soil containing clay and organic matter than in sandy soil. Transfer of pPOB from strain POB310 to strain B13 in soil was never observed.  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