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 109 毫秒
1.
Pseudomonas putida strain DOT-T1E is highly tolerant to organic solvents, with a logP(ow) (the logarithm of the partition coefficient of a solvent in a two-phase water-octanol system of > or =2.5. Solvent tolerant microorganisms can be exploited to develop double-phase (organic solvent and water) biotransformation systems in which toxic substrates or products are kept in the organic phase. We tested P. putida DOT-T1E tolerance to different aliphatic alcohols with a logP(ow) value between 2 and 4, such as decanol, nonanol, and octanol, which are potentially useful in biotransformations in double-phase systems in which compounds with a logP(ow) around 1.5 are produced. P. putida DOT-T1E responds to aliphatic alcohols as the second phase through cis-to-trans isomerization of unsaturated cis fatty acids and through efflux of these aliphatic alcohols via a series of pumps that also extrude aromatic hydrocarbons. These defense mechanisms allow P. putida DOT-T1E to survive well in the presence of high concentrations of the aliphatic alcohols, and growth with nonanol or decanol occurred at a high rate, whereas in the presence of an octanol double-phase growth was compromised. Our results support that the logP(ow) of aliphatic alcohols correlates with their toxic effects, as octanol (logP(ow) = 2.9) has more negative effects in P. putida cells than 1-nonanol (logP(ow) = 3.4) or 1-decanol (logP(ow) = 4). A P. putida DOT-T1E derivative bearing plasmid pWW0-xylE::Km transforms m-xylene (logP(ow) = 3.2) into 3-methylcatechol (logP(ow) = 1.8). The amount of 3-methylcatechol produced in an aliphatic alcohol/water bioreactor was 10- to 20-fold higher than in an aqueous medium, demonstrating the usefulness of double-phase systems for this particular biotransformation.  相似文献   

2.
The basic mechanisms underlying solvent tolerance in Pseudomonas putida DOT-T1E are efflux pumps that remove the solvent from bacterial cell membranes. The solvent-tolerant P. putida DOT-T1E grows in the presence of high concentrations (e.g., 1% [vol/vol]) of toluene and octanol. Growth of P. putida DOT-T1E cells in LB in the presence of toluene supplied via the gas phase has a clear effect on cell survival: the sudden addition of 0.3% (vol/vol) toluene to P. putida DOT-T1E pregrown with toluene in the gas phase resulted in survival of almost 100% of the initial cell number, whereas only 0.01% of cells pregrown in the absence of toluene tolerated exposure to this aromatic hydrocarbon. One class of toluene-sensitive octanol-tolerant mutant was isolated after Tn5-′phoA mutagenesis of wild-type P. putida DOT-T1E cells. The mutant, called P. putida DOT-T1E-18, was extremely sensitive to 0.3% (vol/vol) toluene added when cells were pregrown in the absence of toluene, whereas pregrowth on toluene supplied via the gas phase resulted in survival of about 0.0001% of the initial number. Solvent exclusion was tested with 1,2,4-[14C]trichlorobenzene. The levels of radiochemical accumulated in wild-type cells grown in the absence and in the presence of toluene were not significantly different. In contrast, the mutant was unable to remove 1,2,4-[14C]trichlorobenzene from the cell membranes when grown on Luria-Bertani (LB) medium but was able to remove the aromatic compound when pregrown on LB medium with toluene supplied via the gas phase. The amount of 14C-labeled substrate in whole cells increased in competition assays in which toluene and xylenes were the unlabeled competitors, whereas this was not the case when benzene was the competitor. This finding suggests that the exclusion system works specifically with certain aromatic substrates. The mutation in P. putida DOT-T1E-18 was cloned, and the knockedout gene was sequenced and found to be homologous to the drug exclusion gene mexB, which belongs to the efflux pump family of the resistant nodulator division type.The sensitivity of microorganisms to toxic organic solvents is related to the logarithm of the partition coefficient of the solvent in a mixture of octanol and water (log Pow). Aromatic hydrocarbons with a log Pow of between 1.5 and 3.5 are extremely toxic to living organisms (47). These chemicals dissolve in the cytoplasmic membrane, disorganize it, and collapse the cell membrane potential; this, together with the induced loss of lipids and proteins, leads to irreversible damage resulting in the death of the cell (8, 47, 50).Independent laboratories have isolated Pseudomonas putida strains tolerant to different aromatic hydrocarbons such as toluene, styrene, and p-xylene (6, 15, 42, 48). All four isolated strains were able to grow in liquid culture medium to which a high concentration (1% [vol/vol]) of these aromatic hydrocarbons was added. Tolerance to organic solvents in these P. putida strains is achieved by a series of biochemical mechanisms that actively remove the organic solvent from cell membranes (16, 43) and by physical barriers that help the cell to become (to a certain degree) impermeable to the solvent (13, 37, 43, 48). The physical barriers involve the ordered organization of the cell surface lipopolysaccharides (37) together with modified phospholipids (4, 37, 43, 49). Modifications in phospholipids upon exposure to an organic solvent involve both a short-term response, in which the level of the trans isomers of unsaturated phospholipids increases, and a long-term response consisting of a modification of the polar head groups of phospholipids (4, 43, 49) and an increase in the total amount of phospholipids per dry weight (49). For P. putida DOT-T1, it was suggested that an energy-dependent exclusion system (such as an efflux pump) is critical for tolerance to solvents (43). This conclusion was based on the following findings: (i) P. putida DOT-T1 treated with the uncoupler carbonyl cyanide p-trifluoromethoxyphenyl hydrazone accumulated higher levels of 1,2,4-[14C]trichlorobenzene in cell membranes than did untreated cells, and (ii) P. putida DOT-T1 mutants which were sensitive to toluene, octanol, and other chemicals accumulated 5- to 20-fold-higher levels of 1,2,4-[14C]trichlorobenzene in cell membranes than did the wild-type strain. Similar observations have been reported for Pseudomonas sp. strain S12 (16).In this study, we report that P. putida DOT-T1 uses at least two efflux pumps for toluene exclusion, one that seems to be expressed constitutively and a second inducible one. A mini-Tn5′phoA-Kmr knocked out the constitutive efflux system of P. putida DOT-T1E. The mutant was shown to be hypersensitive to toluene but not to octanol. The Kmr marker of the mini-Tn5 and the 3′ adjacent chromosomal DNA were cloned, and the wild-type gene was rescued by colony screening hybridization and sequenced. Sequence analysis showed that the knocked-out gene in the mutant was a homolog of the mexB gene, which belongs to the efflux pump family of the resistant nodulator division type (3436, 3841).  相似文献   

3.
We assayed the tolerance to solvents of three toluene-degrading Pseudomonas putida strains and Pseudomonas mendocina KR1 in liquid and soil systems. P. putida DOT-T1 tolerated concentrations of heptane, propylbenzene, octanol, and toluene of at least 10% (vol/vol), while P. putida F1 and EEZ15 grew well in the presence of 1% (vol/vol) propylbenzene or 10% (vol/vol) heptane, but not in the presence of similar concentrations of octanol or toluene. P. mendocina KR1 grew only in the presence of heptane. All three P. putida strains were able to become established in a fluvisol soil from the Granada, Spain, area, whereas P. mendocina KR1 did not survive in this soil. The tolerance to organic solvents of all three P. putida strains was therefore assayed in soil. The addition to soil of 10% (vol/wt) heptane or 10% (vol/wt) propylbenzene did not affect the survival of the three P. putida strains. However, the addition of 10% (vol/wt) toluene led to an immediate decrease of several log units in the number of CFU per gram of soil for all of the strains, although P. putida F1 and DOT-T1 subsequently recovered. This recovery was influenced by the humidity of the soil and the incubation temperature. P. putida DOT-T1 recovered from the shock faster than P. putida F1; this allowed the former strain to become established at higher densities in polluted sites into which both strains had been introduced.  相似文献   

4.
The solvent-tolerant strain Pseudomonas putida DOT-T1E has been engineered for biotransformation of toluene into 4-hydroxybenzoate (4-HBA). P. putida DOT-T1E transforms toluene into 3-methylcatechol in a reaction catalyzed by toluene dioxygenase. The todC1C2 genes encode the α and β subunits of the multicomponent enzyme toluene dioxygenase, which catalyzes the first step in the Tod pathway of toluene catabolism. A DOT-T1EΔtodC mutant strain was constructed by homologous recombination and was shown to be unable to use toluene as a sole carbon source. The P. putida pobA gene, whose product is responsible for the hydroxylation of 4-HBA into 3,4-hydroxybenzoate, was cloned by complementation of a Pseudomonas mendocina pobA1 pobA2 double mutant. This pobA gene was knocked out in vitro and used to generate a double mutant, DOT-T1EΔtodCpobA, that was unable to use either toluene or 4-HBA as a carbon source. The tmo and pcu genes from P. mendocina KR1, which catalyze the transformation of toluene into 4-HBA through a combination of the toluene 4-monoxygenase pathway and oxidation of p-cresol into the hydroxylated carboxylic acid, were subcloned in mini-Tn5Tc and stably recruited in the chromosome of DOT-T1EΔtodCpobA. Expression of the tmo and pcu genes took place in a DOT-T1E background due to cross-activation of the tmo promoter by the two-component signal transduction system TodST. Several independent isolates that accumulated 4-HBA in the supernatant from toluene were analyzed. Differences were observed in these clones in the time required for detection of 4-HBA and in the amount of this compound accumulated in the supernatant. The fastest and most noticeable accumulation of 4-HBA (12 mM) was found with a clone designated DOT-T1E-24.  相似文献   

5.
The aim of the study was to investigate whether toxic fine chemical production can be improved using the solvent-tolerant Pseudomonas putida S12 in a two-liquid-phase system consisting of aqueous media and a water-immiscible octanol phase with production of 3-methylcatechol from toluene as the model conversion. For this purpose the genes involved in this conversion, todC1C2BAD from P. putida F1, were introduced into P. putida S12 with high stable expression. Production of 3-methylcatechol was monitored in batch incubations with different media using a single medium and a two-liquid medium–octanol system. The maximum concentration of 3-methylcatechol increased two-fold using the two-liquid medium–octanol system, irrespective of the selected medium. Received: 29 December 1999 / Received revision: 29 February 2000 / Accepted: 6 March 2000  相似文献   

6.
Self-subunit swapping is one of the post-translational maturation of the cobalt-containing nitrile hydratase (Co-NHase) family of enzymes. All of these NHases possess a gene organization of <β-subunit> <α-subunit> <activator protein>, which allows the activator protein to easily form a mediatory complex with the α-subunit of the NHase after translation. Here, we discovered that the incorporation of cobalt into another type of Co-NHase, with a gene organization of <α-subunit> <β-subunit> <activator protein>, was also dependent on self-subunit swapping. We successfully isolated a recombinant NHase activator protein (P14K) of Pseudomonas putida NRRL-18668 by adding a Strep-tag N-terminal to the P14K gene. P14K was found to form a complex [α(StrepP14K)2] with the α-subunit of the NHase. The incorporation of cobalt into the NHase of P. putida was confirmed to be dependent on the α-subunit substitution between the cobalt-containing α(StrepP14K)2 and the cobalt-free NHase. Cobalt was inserted into cobalt-free α(StrepP14K)2 but not into cobalt-free NHase, suggesting that P14K functions not only as a self-subunit swapping chaperone but also as a metallochaperone. In addition, NHase from P. putida was also expressed by a mutant gene that was designed with a <β-subunit> <α-subunit> <P14K> order. Our findings expand the general features of self-subunit swapping maturation.  相似文献   

7.
Pseudomonas putida DOT-T1E was isolated as a toluene-tolerant strain. We show that it is also able to grow on high concentrations (up to 17 g/liter [123 mM]) of p-hydroxybenzoate (4HBA). Tolerance to this aromatic carboxylic acid (up to 30 g/liter [217 mM]) is improved by preexposing the cells to low 4HBA concentrations; the adaptation process is caused by the substrate itself rather than by products resulting from its metabolism. The mechanisms of 4HBA tolerance seem to involve increased rigidity of the cell membrane as a result of a decrease in the cis/trans ratio of unsaturated fatty acids. In addition, energy-dependent efflux systems seem to operate in the exclusion of 4HBA from the cell membranes.  相似文献   

8.
Three groups of volatile compounds, i.e., alcohols, aldehydes and esters, were tested for their effects on arthrospore germination and mycelial growth ofGeotrichum candidum citrus race, the causal agent of citrus sour rot. Alcohols (heptanol, octanol, nonanol, decanol, geraniol, citronellol) at a concentration of 1.0 μl/ml showed 60% or more inhibitory effects on both germination and mycelial growth of this pathogen. Among aldehydes tested, only citral had an inhibitory effect of more than 50%, while esters had no effect. The chain length of straight-chain (C6–C12) alcohols correlated with inhibitory effect, and nonanol (C9) was most active. Treatment with alcohols or citral prior to inoculation reduced colonization and maceration of lemon peel by this pathogen by 70% or more. Results suggested that alcohols or citral can probably be used to prevent the development of citrus sour rot.  相似文献   

9.
In situ mating experiments were done in the River Taff, South Wales, United Kingdom, by using a natural mercury resistance plasmid (pQM1) isolated from a mixture of epilithic bacteria in vitro. The river temperature from March to November was found to influence transfer frequencies strongly (6.8 × 10−9 to 1.5 × 10−2 per recipient). A linear relationship existed between log10 transfer frequency and river temperature (6 to 21°C), a 2.6°C change in temperature giving a 10-fold change in transfer frequency. In vitro experiments showed that pQM1 transferred most efficiently between fluorescent pseudomonads and that one epilithic isolate (Pseudomonas fluorescens) was an efficient donor in situ. Experiments with a P. putida recipient showed that intact epilithic bacterial communities could transfer mercury resistance plasmids in situ at frequencies of up to 3.75 × 10−6 per recipient. Nineteen of the large (>250-kilobase) plasmids isolated by transfer into P. putida were studied in detail and grouped into seven types by restriction digests. Mercury resistance and UV resistance were found to be common linked phenotypes in 19 of the 23 plasmids tested.  相似文献   

10.

Introduction

An exceptionally interesting stress response of Pseudomonas putida strains to toxic substances is the induction of efflux pumps that remove toxic chemical substances from the bacterial cell out to the external environment. To exploit these microorganisms to their full potential a deeper understanding of the interactions between the bacteria and organic solvents is required. Thus, this study focuses on investigation of metabolic changes in P. putida upon exposure to toluene.

Objective

Investigate observable metabolic alterations during interactions of three strains of P. putida (DOT-T1E, and its mutants DOT-T1E-PS28 and DOT-T1E-18) with the aromatic hydrocarbon toluene.

Methods

The growth profiles were measured by taking optical density (OD) measurement at 660 nm (OD660) at various time points during incubation. For fingerprinting analysis, Fourier-transform infrared (FT-IR) spectroscopy was used to investigate any phenotypic changes resulting from exposure to toluene. Metabolic profiling analysis was performed using gas chromatography-mass spectrometry (GC–MS). Principal component—discriminant function analysis (PC-DFA) was applied to the FT-IR data while multiblock principal component analysis (MB-PCA) and N-way analysis of variance (N-way ANOVA) were applied to the GC–MS data.

Results

The growth profiles demonstrated the effect of toluene on bacterial cultures and the results suggest that the mutant P. putida DOT-T1E?18 was more sensitive (significantly affected) to toluene compared to the other two strains. PC-DFA on FT-IR data demonstrated the differentiation between different conditions of toluene on bacterial cells, which indicated phenotypic changes associated with the presence of the solvent within the cell. Fifteen metabolites associated with this phenotypic change, in P. putida due to exposure to solvent, were from central metabolic pathways. Investigation of MB-PCA loading plots and N-way ANOVA for condition | strain × time blocking (dosage of toluene) suggested ornithine as the most significant compound that increased upon solvent exposure.

Conclusion

The combination of metabolic fingerprinting and profiling with suitable multivariate analysis revealed some interesting leads for understanding the mechanism of Pseudomonas strains response to organic solvent exposure.
  相似文献   

11.
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   

12.
As a follow-up to earlier studies on the emission of long-chain alcohols from broth cultures of Gram-negative enteric bacteria, E. coli was examined for the production of 1-octanol, 1-decanol, and 1-dodecanol. Ten strains of E. coli cultured in tryptic soy broth were assayed for volatile metabolites using solid-phase microextraction. Long-chain alcohols were produced by all strains with 1-decanol predominating with production ranging from 23.6 ng mL−1 to 148 ng mL−1. The production of long-chain alcohols followed the onset of the exponential growth phase of the broth culture. Doubling the concentration of glucose (5 g L−1) in the broth had no effect on the concentration of long-chain alcohols produced. Addition of octanoic, decanoic, or dodecanoic acids (as K+ salts) to the broth (100 mg L−1) markedly increased the production of the corresponding alcohols by E. coli, ranging from a 13-fold increase for decanol to a 51-fold increase for dodecanol. However, decanol remained the predominant alcohol detected in all assays. These neutral volatile alcohols may have application as vapor-phase indicators for certain classes of bacteria, particularly, Gram-negative enteric bacteria.  相似文献   

13.
Ultrasound-mediated DNA transfer for bacteria   总被引:2,自引:0,他引:2  
In environmental microbiology, the most commonly used methods of bacterial DNA transfer are conjugation and electroporation. However, conjugation requires physical contact and cell–pilus–cell interactions; electroporation requires low-ionic strength medium and high voltage. These limitations have hampered broad applications of bacterial DNA delivery. We have employed a standard low frequency 40 kHz ultrasound bath to successfully transfer plasmid pBBR1MCS2 into Pseudomonas putida UWC1, Escherichia coli DH5α and Pseudomonas fluorescens SBW25 with high efficiency. Under optimal conditions: ultrasound exposure time of 10 s, 50 mM CaCl2, temperature of 22°C, plasmid concentration of 0.8 ng/µl, P. putida UWC1 cell concentration of 2.5 × 109 CFU (colony forming unit)/ml and reaction volume of 500 µl, the efficiency of ultrasound DNA delivery (UDD) was 9.8 ± 2.3 × 10−6 transformants per cell, which was nine times more efficient than conjugation, and even four times greater than electroporation. We have also transferred pBBR1MCS2 into E. coli DH5α and P. fluorescens SBW25 with efficiencies of 1.16 ± 0.13 × 10−6 and 4.33 ± 0.78 × 10−6 transformants per cell, respectively. Low frequency UDD can be readily scaled up, allowing for the application of UDD not only in laboratory conditions but also on an industrial scale.  相似文献   

14.
We recorded large data sets of swimming trajectories of the soil bacterium Pseudomonas putida. Like other prokaryotic swimmers, P. putida exhibits a motion pattern dominated by persistent runs that are interrupted by turning events. An in-depth analysis of their swimming trajectories revealed that the majority of the turning events is characterized by an angle of ϕ1 = 180° (reversals). To a lesser extent, turning angles of ϕ2 = 0° are also found. Remarkably, we observed that, upon a reversal, the swimming speed changes by a factor of two on average—a prominent feature of the motion pattern that, to our knowledge, has not been reported before. A theoretical model, based on the experimental values for the average run time and the rotational diffusion, recovers the mean-square displacement of P. putida if the two distinct swimming speeds are taken into account. Compared to a swimmer that moves with a constant intermediate speed, the mean-square displacement is strongly enhanced. We furthermore observed a negative dip in the directional autocorrelation at intermediate times, a feature that is only recovered in an extended model, where the nonexponential shape of the run-time distribution is taken into account.  相似文献   

15.
Suberin from the roots of carrots (Daucus carota), parsnip (Pastinaca sativa), rutabaga (Brassica napobrassica), turnip (Brassica rapa), red beet (Beta vulgaris), and sweet potato (Ipomoea batatas) was isolated by a combination of chemical and enzymatic techniques. Finely powdered suberin was depolymerized with 14% BF3 in methanol, and soluble monomers (20-50% of suberin) were fractionated into phenolic (<10%) and aliphatic (13-35%) fractions. The aliphatic fractions consisted mainly of ω-hydroxyacids (29-43%), dicarboxylic acids (16-27%), fatty acids (4-18%), and fatty alcohols (3-6%). Each fraction was subjected to combined gas-liquid chromatography and mass spectrometry. Among the fatty acids very long chain acids (>C20) were the dominant components in all six plants. In the alcohol fraction C18, C20, C22, and C24 saturated primary alcohols were the major components. C16 and C18 dicarboxylic acids were the major dicarboxylic acids of the suberin of all six plants and in all cases octadec-9-ene-1, 18-dioic acid was the major component except in rutabaga where hexadecane-1, 16-dioic acid was the major dicarboxylic acid. The composition of the ω-hydroxyacid fraction was quite similar to that of the dicarboxylic acids; 18-hydroxy-octadec-9-enoic acid was the major component in all plants except rutabaga, where equal quantities of 16-hydroxyhexadecanoic acid and 18-hydroxyoctadec-9-enoic acid (42% each) were found. Compounds which would be derived from 18-hydroxyoctadec-9-enoic acid and octadec-9-ene-1, 18-dioic acid by epoxidation, and epoxidation followed by hydration of the epoxide, were also detected in most of the suberin samples. The monomer composition of the six plants showed general similarities but quite clear taxonomic differences.  相似文献   

16.
Pseudomonas putida is a highly solvent-resistant microorganism and useful chassis for the production of value-added compounds from lignocellulosic residues, in particular aromatic compounds that are made from phenylalanine. The use of these agricultural residues requires a two-step treatment to release the components of the polysaccharides of cellulose and hemicellulose as monomeric sugars, the most abundant monomers being glucose and xylose. Pan-genomic studies have shown that Pseudomonas putida metabolizes glucose through three convergent pathways to yield 6-phosphogluconate and subsequently metabolizes it through the Entner–Doudoroff pathway, but the strains do not degrade xylose. The valorization of both sugars is critical from the point of view of economic viability of the process. For this reason, a P. putida strain was endowed with the ability to metabolize xylose via the xylose isomerase pathway, by incorporating heterologous catabolic genes that convert this C5 sugar into intermediates of the pentose phosphate cycle. In addition, the open reading frame T1E_2822, encoding glucose dehydrogenase, was knocked-out to avoid the production of the dead-end product xylonate. We generated a set of DOT-T1E-derived strains that metabolized glucose and xylose simultaneously in culture medium and that reached high cell density with generation times of around 100 min with glucose and around 300 min with xylose. The strains grew in 2G hydrolysates from diluted acid and steam explosion pretreated corn stover and sugarcane straw. During growth, the strains metabolized > 98% of glucose, > 96% xylose and > 85% acetic acid. In 2G hydrolysates P. putida 5PL, a DOT-T1E derivative strain that carries up to five independent mutations to avoid phenylalanine metabolism, accumulated this amino acid in the medium. We constructed P. putida 5PLΔgcd (xylABE) that produced up to 250 mg l−1 of phenylalanine when grown in 2G pretreated corn stover or sugarcane straw. These results support as a proof of concept the potential of P. putida as a chassis for 2G processes.  相似文献   

17.
The solvent-tolerant strain Pseudomonas putida DOT-T1E was grown in batch fermentations in a 5-liter bioreactor in the presence and absence of 10% (vol/vol) of the organic solvent 1-decanol. The growth behavior and cellular energetics, such as the cellular ATP content and the energy charge, as well as the cell surface hydrophobicity and charge, were measured in cells growing in the presence and absence of 1-decanol. Although the cells growing in the presence of 1-decanol showed an about 10% reduced growth rate and a 48% reduced growth yield, no significant differences were measured either in the ATP and potassium contents or in the energy charge, indicating that the cells adapted completely at the levels of membrane permeability and energetics. Although the bacteria needed additional energy for adaptation to the presence of the solvent, they were able to maintain or activate electron transport phosphorylation, allowing homeostasis of the ATP level and energy charge in the presence of the solvent, at the price of a reduced growth yield. On the other hand, significantly enhanced cell hydrophobicities and more negative cell surface charges were observed in cells grown in the presence of 1-decanol. Both reactions occurred within about 10 min after the addition of the solvent and were significantly different after killing of the cells with toxic concentrations of HgCl2. This adaptation of the surface properties of the bacterium to the presence of solvents seems to be very similar to previously observed reactions on the level of lipopolysaccharides, with which bacteria adapt to environmental stresses, such as heat shock, antibiotics, or low oxygen content. The results give clear physiological indications that the process with P. putida DOT-T1E as the biocatalyst and 1-decanol as the solvent is a stable system for two-phase biotransformations that will allow the production of fine chemicals in economically sound amounts.  相似文献   

18.
Lactic acid bacteria (LAB) have been shown to enhance performance of weaned piglets. However, few studies have reported the addition of LAB Enterococcus faecalis as alternatives to growth promoting antibiotics for weaned piglets. This study evaluated the effects of dietary E. faecalis LAB31 on the growth performance, diarrhea incidence, blood parameters, fecal bacterial and Lactobacillus communities in weaned piglets. A total of 360 piglets weaned at 26 ± 2 days of age were randomly allotted to 5 groups (20 pens, with 4 pens for each group) for a trial of 28 days: group N (negative control, without antibiotics or probiotics); group P (Neomycin sulfate, 100 mg/kg feed); groups L, M and H (supplemented with E. faecalis LAB31 0.5×109, 1.0×109, and 2.5×109 CFU/kg feed, respectively). Average daily gain and feed conversion efficiency were found to be higher in group H than in group N, and showed significant differences between group H and group P (P0 < 0.05). Furthermore, groups H and P had a lower diarrhea index than the other three groups (P0 < 0.05). Denaturing gradient gel electrophoresis (DGGE) showed that the application of probiotics to the diet changed the bacterial community, with a higher bacterial diversity in group M than in the other four groups. Real-time PCR revealed that the relative number of Lactobacillus increased by addition of probiotics, and was higher in group H than in group N (P0 < 0.05). However, group-specific PCR-DGGE showed no obvious difference among the five groups in Lactobacillus composition and diversity. Therefore, the dietary addition of E. faecalis LAB31 can improve growth performance, reduce diarrhea, and increase the relative number of Lactobacillus in feces of weaned piglets.  相似文献   

19.
The influence of trichloroethylene (TCE) on a mixed culture of four different toluene-degrading bacterial strains (Pseudomonas putida mt-2, P. putida F1, P. putida GJ31, and Burkholderia cepacia G4) was studied with a fed-batch culture. The strains were competing for toluene, which was added at a very low rate (31 nmol mg of cells [dry weight]−1 h−1). All four strains were maintained in the mixed culture at comparable numbers when TCE was absent. After the start of the addition of TCE, the viabilities of B. cepacia G4 and P. putida F1 and GJ31 decreased 50- to 1,000-fold in 1 month. These bacteria can degrade TCE, although at considerably different rates. P. putida mt-2, which did not degrade TCE, became the dominant organism. Kinetic analysis showed that the presence of TCE caused up to a ninefold reduction in the affinity for toluene of the three disappearing strains, indicating that inhibition of toluene degradation by TCE occurred. While P. putida mt-2 took over the culture, mutants of this strain which could no longer grow on p-xylene arose. Most of them had less or no meta-cleavage activity and were able to grow on toluene with a higher growth rate. The results indicate that cometabolic degradation of TCE has a negative effect on the maintenance and competitive behavior of toluene-utilizing organisms that transform TCE.  相似文献   

20.
Our objective was to evaluate methods for identifying cattle with high concentrations of Escherichia coli O157 in their feces. In two experiments, feces were collected from cattle orally inoculated with nalidixic acid (Nal)-resistant E. coli O157, and direct plating of diluted feces on sorbitol MacConkey agar with cefixime and potassium tellurite (CT-SMAC) containing Nal was considered the gold standard (GS) method. In experiment 1, methods evaluated were preenrichment direct streak, immunomagnetic separation with most probable number (MPN), and postenrichment direct streak with MPN, all using CT-SMAC. The mean concentration of Nal-resistant E. coli O157 in samples (n = 59) by use of the GS was 3.6 log10 CFU/g. The preenrichment streak detected >3.0 log10 CFU/g samples with a 74.4% sensitivity and 68.8% specificity. Postenrichment direct streak-MPN and immunomagnetic separation-MPN concentrations were correlated significantly with GS concentrations (r = 0.53 and r = 0.39, respectively). In experiment 2 (480 samples), pre- and postenrichment direct streaking performed in triplicate and spiral plating on CT-SMAC were evaluated. For preenrichment streaks, sensitivity was 79.7% and specificity was 96.7% for detecting >3.0 log10 CFU/g when the criterion was positive cultures on at least two plates. For spiral plating at that concentration, sensitivity and specificity were 83.9% and 56.3%, respectively. Postenrichment streaking performed relatively poorly. Triplicate preenrichment streaks of 1:10-diluted feces on CT-SMAC may be useful for identifying cattle shedding high concentrations of E. coli O157. Estimates of sensitivity and specificity enable appropriate application of methods and interpretation of results and may enhance applied research, surveillance, and risk assessments.  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