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.
We report the identification and cloning of an ntrA-like (glnF rpoN) gene of Rhizobium meliloti and show that the R. meliloti ntrA product (NtrA) is required for C4-dicarboxylate transport as well as for nitrate assimilation and symbiotic nitrogen fixation. DNA sequence analysis showed that R. meliloti NtrA is 38% homologous with Klebsiella pneumoniae NtrA. Subcloning and complementation analysis suggested that the R. meliloti ntrA promoter lies within 125 base pairs of the initiation codon and may be constitutively expressed.  相似文献   

2.
Summary An ntrA (rpoN) mutant of Pseudomonas putida, in which the gene was insertionally inactivated, was constructed. The mutant cells did not have flagella, thus accounting for their poor motility. The mutant phenotype was complemented by introduction of the intact ntrA gene.  相似文献   

3.
We cloned the rpoN (ntrA and glnF) gene encoding sigma(54) from the phytopathogen Pseudomonas syringae pv. maculicola strain ES4326. The P. syringae ES4326 rpoN gene complemented Pseudomonas aeruginosa, Escherichia coli, and Klebsiella aerogenes rpoN mutants for a variety of rpoN mutant phenotypes, including the inability to utilize nitrate as sole nitrogen source. DNA sequence analysis of the P. syringae ES4326 rpoN gene revealed that the deduced amino acid sequence was most similar (86% identity; 95% similarity) to the sigma(54) protein encoded by the Pseudomonas putida rpoN gene. A marker exchange protocol was used to construct an ES4326 rpoN insertional mutation, rpoN::Km(r). In contrast to wild-type ES4326, ES4326 rpoN::Km(r) was nonmotile and could not utilize nitrate, urea, C(4)-dicarboxylic acids, several amino acids, or concentrations of ammonia below 2 mM as nitrogen sources. rpoN was essential for production of the phytotoxin coronatine and for expression of the structural genes encoding coronamic acid. In addition, ES4326 rpoN::Km(r) did not multiply or elicit disease symptoms when infiltrated into Arabidopsis thaliana leaves, did not elicit the accumulation of several Arabidopsis defense-related mRNAs, and did not elicit a hypersensitive response (HR) when infiltrated into tobacco (Nicotiana tabacum) leaves. Furthermore, whereas P. syringae ES4326 carrying the avirulence gene avrRpt2 elicited an HR when infiltrated into Arabidopsis ecotype Columbia leaves, ES4326 rpoN::Km(r) carrying avrRpt2 elicited no response. Constitutive expression of ES4326 hrpL in ES4326 rpoN::Km(r) partially restored defense-related mRNA accumulation, showing a direct role for the hrp cluster in host defense gene induction in a compatible host-pathogen interaction. However, constitutive expression of hrpL in ES4326 rpoN::Km(r) did not restore coronatine production, showing that coronatine biosynthesis requires factors other than hrpL.  相似文献   

4.
Summary The nucleotide sequence of the Azotobacter vinelandii ntrA gene has been determined. It encodes a 56916 Dalton acidic polypeptide (AvNtrA) with substantial homology to NtrA from Klebsiella pneumoniae (KpNtrA) and Rhizobium meliloti (RmNtrA). NtrA has been shown to act as a novel RNA polymerase sigma factor but the predicted sequence of AvNtrA substantiates our previous analysis of KpNtrA in showing no substantial homology to other known sigma factors. Alignment of the predicted amino acid sequences of AvNtrA, KpNtrA and RmNtrA identified three regions; two showing>50% homology and an intervening sequence of <10% homology. The predicted protein contains a short sequence near the centre with homology to a conserved region in other sigma factors. The C-terminal region contains a region of homology to the subunit of RNA polymerase (RpoC) and two highly conserved regions one of which is significantly homologous to known DNA-binding motifs. In A. vinelandii, ntrA is followed by another open reading frame (ORF) which is highly homologous to a comparable ORF downstream of ntrA in K. pneumoniae and R. meliloti.  相似文献   

5.
6.
We describe swarming in Pseudomonas aeruginosa as a third mode of surface translocation in addition to the previously described swimming and twitching motilities. Swarming in P. aeruginosa is induced on semisolid surfaces (0.5 to 0.7% agar) under conditions of nitrogen limitation and in response to certain amino acids. Glutamate, aspartate, histidine, or proline, when provided as the sole source of nitrogen, induced swarming, while arginine, asparagine, and glutamine, among other amino acids, did not sustain swarming. Cells from the edge of the swarm were about twice as long as cells from the swarm center. In both instances, bacteria possessing two polar flagella were observed by light and electron microscopy. While a fliC mutant of P. aeruginosa displayed slightly diminished swarming, a pilR and a pilA mutant, both deficient in type IV pili, were unable to swarm. Furthermore, cells with mutations in the las cell-to-cell signaling system showed diminished swarming behavior, while rhl mutants were completely unable to swarm. Evidence is presented for rhamnolipids being the actual surfactant involved in swarming motility, which explains the involvement of the cell-to-cell signaling circuitry of P. aeruginosa in this type of surface motility.  相似文献   

7.
Pseudomonas putida Which Can Grow in the Presence of Toluene   总被引:7,自引:3,他引:4       下载免费PDF全文
A Pseudomonas putida strain able to grow in the presence of more than 50% toluene was isolated from soil. The strain was tolerant of other toxic solvents, including aliphatic hydrocarbons, alicyclic hydrocarbons, aromatic hydrocarbons, alcohols, and ethers. The stability of the solvent tolerance of strain IH-2000 was stimulated by addition of Mg2+ and Ca2+ to the medium containing toluene.  相似文献   

8.
9.
Dehydroalanine is present in the histidine ammonia-lyase (histidase) from Pseudomonas putida ATCC 12633 as shown by reaction of purified enzyme with K14CN or NaB3H4 and subsequent identification of [14C]aspartate or [3H]alanine, respectively, following acid hydrolysis of the labeled protein. When labeling with cyanide was conducted under denaturing conditions, 4 mol of [14C]cyanide was incorporated per mol of enzyme (Mr 220 000), equivalent to one dehydroalanine residue being modified per subunit in this protein composed of four essentially identical subunits. In native enzyme, inactivation of catalytic activity by cyanide was complete when 1 mol of [14C]cyanide had reacted per mol of histidase, suggesting that modification of any one of the four dehydroalanine residues in the tetrameric enzyme was sufficient to prevent catalysis at all sites. Loss of activity on treatment with cyanide could be blocked by the addition of the competitive inhibitor cysteine or substrate if Mn2+ was also present. Cross-linking of native enzyme with dimethyl suberimidate produced no species larger than tetramer, thereby eliminating the possibility that an aggregation phenomenon might explain why only one-fourth of the dehydroalanyl residues was modified by cyanide during inactivation. A labeled tryptic peptide was isolated from enzyme inactivated with [14C]cyanide. Its composition was different from that of a tryptic peptide previously isolated from other histidases and shown to contain a highly reactive and catalytically important cysteine residue. Such a finding indicates the dehydroalanine group is distinct from the active site cysteine. Treatment of crude extracts with [14C]cyanide and purification of the inactive enzyme yielded labeled protein that release [14C]aspartate on acid hydrolysis.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

10.
11.
The uptake of germanium by Pseudomonas putida ATCC 33015 was studied in the presence of catechol or acetate or both as representative substrates differing in their ability to form complexes with this element. The bacteria were taken from a batch culture grown on acetate as the sole carbon source. Cells introduced into a medium containing germanium and either catechol or a mixture of catechol and acetate accumulated germanium in a biphasic way. After a lower level of accumulation that corresponded to the value obtained in the presence of acetate was reached, a further increase in the germanium content up to a higher saturation level was observed. The appearance of the second step of accumulation, which corresponded to the linear degradation of catechol, proved that catechol facilitated the transport of germanium into the cells through the nonspecific uptake of the germanium-catechol complex by an inducible catechol transport system.  相似文献   

12.
13.
14.
15.
Adapted Pseudomonas putida strains grew in the presence of up to 6% (vol/vol) butanol, the highest reported butanol concentration tolerated by a microbe. P. putida might be an alternative host for biobutanol production, overcoming the primary limitation of currently used strains—insufficient product titers due to low butanol tolerance.The focus of biofuel production research has recently shifted from ethanol to bioenergy carriers that are more compatible with existing infrastructure (e.g., refineries, transport, and car engines). At the forefront is n-butanol (hereafter referred to as butanol) for which large-scale production processes have been implemented (16, 35). Existing fermentations, however, are limited in energetically attractive butanol titers, because butanol inhibits microbial growth at concentrations above 16 g/liter (2, 10). As reported for other organic solvents with low logarithm of the partition coefficient in a two-phase octanol/water system (log Pow), this toxicity is due primarily to accumulation of butanol (log Pow, 0.8) in the cell membrane and subsequent impairment (4, 17, 30, 33). With the maximum aqueous solubility of 0.97 M (8.8% [vol/vol]), the maximum membrane concentration of butanol was calculated to be 1.59 M (17), spotlighting its potential toxicity. The low achievable butanol titers have necessitated large reactor volumes, resulting in high purification costs (8, 15). Recent metabolic engineering strategies for improving biobutanol fermentation have focused on maximization of butanol production rates (10, 19), reducing the levels of by-products (20), finding alternative substrates (20), or finding alternative hosts (2, 12, 21, 31). However, recently engineered microbial strains (1, 14) have not overcome butanol toxicity.High organic solvent concentrations are tolerated by strains of the bacterial species Pseudomonas putida reported to grow in a second phase of octanol (25), toluene (13), or styrene (32). This suggests that solvent-tolerant P. putida strains withstand high butanol titers and therefore warrant exploitation as host for butanol production. Indeed, viable solvent-tolerant P. putida S12 cells were observed at butanol concentrations of up to 10% (vol/vol) by live-dead staining and fluorescence microscopy (5) (see supplemental material). We used growth as the parameter of interest, because growth in the presence of butanol directly indicates the potential of selected P. putida strains as hosts for recombinant butanol production.Three solvent-tolerant P. putida strains, DOT-T1E (23), S12 (32), and Pseudomonas sp. strain VLB120 (18), and the solvent-sensitive P. putida reference strain KT2440 (24) were examined for their ability to grow in the presence of butanol. Toxicity assays were performed in 96-well microtiter plates (System Duetz [7]) at 30°C and 300 rpm using glucose-supplemented LB and M9 media (with 10 and 5 g/liter glucose, respectively) (26). Higher glucose concentrations in LB medium did not increase butanol tolerance (data not shown). Butanol was added in all experiments to cells in the mid-exponential phase. Cell growth was monitored by changes in optical density, and substrate and butanol concentrations were analyzed by high-pressure liquid chromatography (Trentec 308R-Gel.H; VWR Hitachi). Comparable low butanol concentrations were withstood by all P. putida strains, with butanol tolerance highly dependent on the medium composition (Table (Table1).1). Growth was observed at butanol concentrations up to 3% (vol/vol), occurring in a culture of Pseudomonas sp. strain VLB120 using glucose-supplemented LB medium.

TABLE 1.

Tolerated butanol concentrations in different growth media
Pseudomonas strain and treatment or cell typeMaximum butanol concn [% (vol/vol)]a
M9 minimal medium with glucose (5 g/liter)LB mediumLB medium with glucose (10 g/liter)
P. putida DOT-T1E
    Untreated1.51.5-2.02.5
    Adapted1.0 (1.0)1.5 (2.0)6.0 (5.0)
P. putida KT2440
    Untreated1.01.52.0
    Treated1.0 (1.0)1.5 (1.0)1.5 (1.5)
P. putida S12
    Untreated1.52.02.5
    Adapted1.0 (1.0)1.5-2.0 (1.5)6.0 (5.0)
Pseudomonas sp. strain VLB120
    Untreated1.52.02.5-3.0
    Adapted1.0 (1.5)1.5-2.0 (1.5)6.0 (6.0)
Open in a separate windowaValues represent the maximum butanol concentration allowing growth (growth rate of ≥0.05 h−1). Data in parentheses were measured in experiments with cells that were stored at −80°C.Because reported adaptation approaches (3, 17, 18, 32) were not successful (see supplemental material), a modified adaptation protocol was developed. Cells were incubated at 30°C on LB agar plates in an airtight desiccator with a butanol saturated gas phase. Colonies were repeatedly transferred every 2 days to new plates for at least 15 times. Cells that underwent this procedure, referred to as treated cells, were harvested and either stored at −80°C prior to testing or assessed immediately for tolerance to butanol (Fig. (Fig.11).Open in a separate windowFIG. 1.Butanol tolerance of P. putida. Growth rates of untreated (A) and adapted (B) cells in LB medium with 10 g/liter glucose as an additional energy and carbon source. The concentration of butanol (cBuOH) is shown on the x axis. The growth rates are normalized to the growth rate in the respective control experiments without butanol. Lines are drawn for better visualization. Error bars present standard deviations of independent experiments (n = 3 to 6). Symbols: ▪, P. putida DOT-T1E; •, P. putida KT2440; ▴, P. putida S12; ▾, Pseudomonas sp. strain VLB120.The treated solvent-tolerant cells grew at rates above 0.05 h−1 (approximately 5% of the maximum growth rate without butanol) in the presence of up to 6% (vol/vol) butanol. Butanol concentrations in the medium decreased during the experiments due to evaporation (i.e., at a rate of 0.76 ± 0.03 mmol l−1 h−1) from an initial concentration of 5% (vol/vol) and, more significantly, due to consumption. Similar butanol uptake rates were observed for all four strains at 5% (vol/vol) initial butanol, ranging from 5.2 to 6.6 mmol l−1 h−1. Therefore, the butanol concentration decreased to only 3.5% (vol/vol) and 4% (vol/vol) after 9 h of cultivation in experiments at initial butanol concentrations of 5% (vol/vol) and 6% (vol/vol), respectively. This decrease resulted in an average butanol concentration of 4.5% (vol/vol) tolerated by the DOT-T1E, S12, and VLB120 cells. Notably, the time course of butanol concentration did not differ significantly with solvent-sensitive P. putida KT2440 that did not grow above 1.5% (vol/vol) butanol.To rationalize the metabolic responses of untreated and treated strains to butanol, we performed 13C-labeled tracer-based flux analysis (3, 18, 27, 34), using minimal medium with 20% U-13C-labeled and 80% naturally labeled glucose, as reported recently (3, 6, 9). During growth without butanol, the four Pseudomonas strains had similar intracellular carbon flux distributions, independent of any prior adaptation to butanol (data not shown). In the presence of butanol, all untreated cells revealed significantly higher specific glucose uptake rates while growth rates decreased (Fig. (Fig.2).2). The reduced biomass yield was not caused by by-product formation (data not shown) but by changes in intracellular flux distribution: the carbon flux was rerouted from biomass synthesis to the tricarboxylic acid (TCA) cycle, which was fueled by pyruvate via pyruvate dehydrogenase and citrate synthase activity. The anaplerotic and gluconeogenic reactions were unaffected. The overall redox cofactor regeneration rates (approximately fourfold higher) resulting from this rerouting suggest that larger amounts of energy are demanded for cell maintenance during butanol stress, similar to the response of P. putida during growth in the presence of other organic solvents with low log Pow (22, 23, 28).Open in a separate windowFIG. 2.Flux distributions in P. putida under butanol stress conditions. The flux distributions in the P. putida strains DOT-T1E, KT2440, and S12 and Pseudomonas sp. strain VLB120 (from top to bottom) were determined during growth in glucose-containing M9 medium supplemented with 1% (vol/vol) butanol using untreated and adapted cells. Butanol catabolism was traced by the fractional labeling of central carbon metabolites (see text for details). The errors for all fluxes were below 10% with the exception of highly active or negligibly fluxes including PEP carboxykinase, pentose-phosphate-pathway (PPP), and phosphoglycoisomerase. The upper bound of the NAD(P)H regeneration rate is presented. Glucose-6-P, glucose-6-phosphate; PGA, 3-phosphoglycerate; PEP, phosphoenolpyruvate.In contrast, physiology and flux distributions differed for adapted DOT-T1E, S12, and VLB120 cells, but not treated KT2440 cells. These strains, coping with high butanol concentrations, had low net glucose consumptions, resulting in comparably lower TCA cycle fluxes and consequently lower redox cofactor regeneration rates (Fig. (Fig.2).2). As indicated above (Fig. (Fig.1),1), P. putida KT2440 did not adapt to butanol, and no metabolic changes were observed compared with the untreated strain.Coconsumption of butanol was considered in calculating the absolute intracellular fluxes by correcting the fractional labeling [FL = n13C/(n12C + n13C)] of the affected amino acids—aspartate, glutamine, isoleucine, leucine, and threonine. The dilution of the fractional isotope label due to butanol coconsumption decreased from acetyl coenzyme A (acetyl-CoA) (FL = 8%) to 2-ketoglutarate (FL = 13%) and oxaloacetate (FL = 15%), suggesting that butanol is cometabolized via β-oxidation to acetyl-CoA, followed by oxidation in the TCA cycle.As calculated from the fractional label of the m-15 isotopomer of leucine (FL = 14%), approximately 60% of the acetyl-CoA originated from butanol. For example, in P. putida KT2440, butanol contributed to the synthesis of acetyl-CoA about 7.22 ± 0.23 mmol g−1 h−1, corresponding to the measured glucose uptake rate of 11.22 ± 0.74 mmol g−1 h−1 [(7.22/11.22) × 100 = 64%]. The untreated solvent-tolerant strains had slightly lower consumption rates of approximately 6.5 mmol g−1 h−1 for butanol and 10.2 mmol g−1 h−1 for glucose. Compared with the untreated strains, adapted DOT-T1E, S12, and VLB120 cells had lower uptake rates of 3.8 to 5.2 mmol g−1 h−1 for butanol and 4.9 to 6.6 mmol g−1 h−1 for glucose. Butanol did not contribute significantly to the synthesis of pyruvate (FL = 19%) and PEP (FL = 20%, or the contribution was below the FL detection limit of 0.5%), suggesting that malic enzyme and phosphoenolpyruvate (PEP) carboxykinase are marginally active under these conditions. This suggests that a synthetic pathway for butanol synthesis from glucose can be implemented in P. putida using native genes for butanol dehydrogenase and aldehyde dehydrogenase with a concomitant decrease of ß-oxidation activity.Butanol degradation of P. putida KT2440 was comparable with the rates of solvent-tolerant cells, but butanol tolerance was not induced, suggesting activity of additional mechanisms of adaptation or tolerance, such as solvent removal by efflux pumps and physiochemical changes of membrane lipids (11, 22). These mechanisms reduce cellular growth rates and biomass yields by imposing higher energy demands. Additionally, energy loss can be caused by swelling and alteration of the lipid layer due to increased proton permeability of the membrane (4) and by reduced efficiency of the electron transport chain (30). In butanol-tolerant cells, the observed reduction in TCA cycle use and energy production in the presence of butanol suggests cell membrane adaptation by lowering its energy demands for maintenance.The observed higher tolerance to butanol in LB medium compared with minimal medium can also be explained by decreased metabolic costs for sustaining biomass synthesis due to direct supply of biomass precursors like amino acids (29). Additional supplementation of LB medium with glucose enhanced butanol tolerance, most likely due to increased energy supplies. For P. putida S12, we calculated glucose uptake rates of 8.01 ± 0.21 mmol g−1 h−1 and 13.53 ± 0.34 mmol g−1 h−1 at initial butanol concentrations of 1% (vol/vol) and 3% (vol/vol), respectively, translating into an increased ATP regeneration rate at 3% (vol/vol) butanol of minimally 13.5 mmol g−1 h−1 (substrate phosphorylation via the Entner-Doudoroff pathway) and up to approximately 350 mmol g−1 h−1 (oxidative phosphorylation). The additional energy demand in the presence of butanol necessitates particular attention during strain and medium engineering.We report solvent-tolerant P. putida strains growing at butanol concentrations as high as 6% (vol/vol). Metabolic flux analysis suggests that this is not based on glucose-butanol coconsumption but rather effected by lowered cell maintenance costs.In conclusion, butanol-tolerant P. putida strains are promising candidates as production hosts, overcoming the principal limitation of biobutanol production—product inhibition at low concentrations.  相似文献   

16.
Previously established PCR amplification and Southern hybridization procedures were developed for the isolation of the 0.8-kb flagellin gene in Pseudomonas putida. The deduced protein sequence has significant homology to the N- and C-terminal sequences of other bacterial flagellins. We propose that P. putida flagellin genes can be divided at least into three size groups: type I (2.0 kb), type II (1.4 kb), and type III (0.8 kb). Type I and type II flagellin genes have been reported. The new 0.8-kb type III gene was expressed in E. coli, and the resulting protein was purified and used to raise polyclonal antibody to study whether this small gene encodes flagellin. The antiserum reacted with purified flagellin monomers from representatives of each flagellin type, as well as proteins of the same sizes in lysates of these organisms, on Western immunoblots. This antiserum was determined to be functional in a motility inhibition assay. Similar results were obtained from antiserum directed against purified type III flagellin, indicating that a new type of flagellin gene in P. putida has been found. Preliminary electron microscopic study revealed that P. putida isolate with the smaller flagellin gene type appeared to have a thinner flagellar filament.  相似文献   

17.
A nicotine-sensitive mutant was generated from the nicotine-degrading bacterium, Pseudomonas putida strain J5, by mini-Tn5 transposon mutagenesis. This mutant was unable to grow with nicotine as the sole carbon source but could grow with glucose. Sequence analysis showed that the Tn5 transposon inserted at the site of the ketopantoate hydroxymethyltransferase gene (panB), which had 54% identity to PanB in Escherichia coli K-12. In-frame deletion of the panB gene abolished the nicotine-degrading ability of strain J5, while complementation with panB from P. putida J5 and E. coli K-12 restored the degrading activity of the mutant to the wild-type level. These results suggest that ketopantoate hydroxymethyltransferase is a crucial enzyme in nicotine metabolism in P. putida J5.  相似文献   

18.
19.
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