首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
Pseudomonas putida mt-2, harbouring the TOL plasmid pWW0, was grown in chemostat culture under succinate-, sulphate-, ammonium- or phosphate-limitation at different dilution rates. The fraction of mutant cells lacking the plasmid-encoded enzymes for the degradation of toluene and xylene (TOL- cells), was determined. Genetic analysis revealed that all TOL- cells isolated harboured partially deleted plasmids, lacking the TOL catabolic genes. The growth-rate advantage of the TOL- cells was quantified from the kinetics of their increase as a fraction of the total population. At a dilution rate of 0.1 h-1 no growth-rate advantage of TOL- cells was found when phosphate or ammonium were limiting. Under sulphate-limitation, ingrowth of TOL- cells was evident but did not follow a straightforward pattern. Under succinate-limitation the growth-rate advantage was the highest, particularly at low dilution rates (about 50% at D = 0.05 h-1). In phauxostat culture, at the maximal growth rate, the growth-rate advantage of TOL- cells was less than 1%. The specific activity in TOL+ cells of the plasmid-encoded enzyme catechol 2,3-dioxygenase was relatively high at a low growth rate.  相似文献   

2.
WR211 is a transconjugant resulting from transfer of the 117-kilobase (kb) TOL degradative plasmid pWW0 into Pseudomonas sp. strain B13. The plasmid of this strain, pWW01211, is 78 kb long, having suffered a deletion of 39 kb. We show that WR211 contains the 39 kb that is missing from its plasmid, together with at least an additional 17 kb of pWW0 DNA integrated in another part of the genome, probably the chromosome. The ability of WR211 to grow on the TOL-specific substrate m-toluate is the result of expression of the TOL genes in this alternative location, whereas its inability to grow on m-xylene is caused by insertional mutagenesis by 3 kb of DNA of unknown origin in the xylR gene of this DNA. The resident plasmid pWW01211 plays no part in the degradative phenotype of WR211 since it can be expelled by mating in incompatible IncP9 resistance plasmid R2 or pMG18 without loss of the phenotype. This alternatively located DNA can be rescued back into the R2 and pMG18 plasmids as R2::TOL and pMG18::TOL recombinants by mating out into plasmid-free recipients and selecting for Mtol+ transconjugants. In all cases examined, these plasmids contained the entire R plasmid into which is inserted 59 kb of DNA, made up of 56 kb of pWW0 DNA and the 3-kb xylR insertion. Selection for faster growth on benzoate can lead to precise excision of the 39 kb from the TOL region of an R2::TOL recombinant, leaving a residual and apparently cryptic 17-kb segment of pWW0 DNA in the R plasmid.  相似文献   

3.
Supercoiled DNA plasmids were exposed in the frozen state to high-energy electrons. Surviving supercoiled molecules were separated from their degradation products (e.g., open circle and linear forms) by agarose gel electrophoresis and subsequently quantified by staining and image analysis. Complex survival curves were analyzed using radiation target theory, yielding the radiation-sensitive mass of each form. One of the irradiated plasmids was transfected into cells, permitting radiation analysis of gene expression. Loss of this function was associated with a mass much smaller than the entire plasmid molecule, indicating a lack of energy transfer in amounts sufficient to cause structural damage along the DNA polynucleotide. The method of radiation target analysis can be applied to study both structure and function of DNA.  相似文献   

4.
Abstract Batch mating experiments were employed to study the kinetics of the conjugal transfer of a TOL plasmid, using the transconjugant strain Pseudomonas aeruginosa PAO 1162 (TOL) as the plasmid donor and Pseudomonas putida PB 2442 and Pseudomonas aeruginosa PAO 1162N as the plasmid recipients. Transfer rates from PAO 1162 (TOL) to PAO 1162N and PB 2442 measured for exponentially grown PAO 1162 (TOL) were 1.81 × 10−14 (standard error (S.E.) 1.25 × 10−15) ml·cell−1min−1 and 3.32 × 10−13 (S.E. 4.42 × 10−14) ml·cell−1min−1, respectively. The instability of the TOL plasmid in PAO 1162 (TOL) was evaluated under conditions that were non-selective for maintenance of the TOL catabolic functions. The measured rates of instability were 6.7 10−6 to 8.3 10−6 min−1, and the loss of the catabolic functions was mainly caused by structural instability of the plasmid.  相似文献   

5.
A psychrotrophic bacterium, originally isolated from a natural aquatic environment, was characterized and identified as Pseudomonas putida Q5 for use as a representative recipient for biodegradative genes from a mesophilic microorganism. The TOL plasmid pWWO of the mesophile P. putida PaW1 was successfully transferred by conjugation to the naturally isolated psychrotroph P. putida Q5, as shown by plasmid analysis by agarose gel electrophoresis. Expression of the genes encoded by the mesophilic TOL plasmid in the psychrotroph was shown by the fact that the transconjugant (designated P. putida Q5T) had the capacity to degrade and utilize toluate (1,000 mg/liter) as a sole source of carbon at temperatures as low as 0 degrees C. Comparison of growth rates over a wide temperature range (0 to 30 degrees C) indicated that the physiological activity of the transconjugant was not reduced and that the plasmid DNA from the mesophile and its encoded enzymes functioned effectively in the psychrotroph at temperatures well below those at which the mesophile could grow. The production and demonstrated functioning of P. putida Q5T illustrates the possibility of developing specific degradative capacities in bacteria which can readily function at low temperatures in chemically contaminated environments or in industrial wastewater treatment systems.  相似文献   

6.
A psychrotrophic bacterium, originally isolated from a natural aquatic environment, was characterized and identified as Pseudomonas putida Q5 for use as a representative recipient for biodegradative genes from a mesophilic microorganism. The TOL plasmid pWWO of the mesophile P. putida PaW1 was successfully transferred by conjugation to the naturally isolated psychrotroph P. putida Q5, as shown by plasmid analysis by agarose gel electrophoresis. Expression of the genes encoded by the mesophilic TOL plasmid in the psychrotroph was shown by the fact that the transconjugant (designated P. putida Q5T) had the capacity to degrade and utilize toluate (1,000 mg/liter) as a sole source of carbon at temperatures as low as 0 degrees C. Comparison of growth rates over a wide temperature range (0 to 30 degrees C) indicated that the physiological activity of the transconjugant was not reduced and that the plasmid DNA from the mesophile and its encoded enzymes functioned effectively in the psychrotroph at temperatures well below those at which the mesophile could grow. The production and demonstrated functioning of P. putida Q5T illustrates the possibility of developing specific degradative capacities in bacteria which can readily function at low temperatures in chemically contaminated environments or in industrial wastewater treatment systems.  相似文献   

7.
Pseudomonas putida mt-2, harbouring the TOL plasmid PWW0, was grown continuously on benzoate in a phauxostat at a non-limited rate. The gradual decrease in the population carrying the complete TOL plasmid was caused predominantly by a growth-rate advantage of spontaneous mutants carrying a partially deleted plasmid (TOL- cells). The growth-rate difference (v) was quantified both by measuring the increase in the dilution rate (from 0.68 to 0.79 h-1; v = 0.11 h-1) and by mathematical analysis of the ingrowth of TOL- cells (v = 0.12 h-1). The latter procedure also established that the segregation rate was of the order of magnitude 10(-5) h-1. Similar values for the growth-rate advantage and the segregation rate were found when both benzoate and succinate were present in non-limiting concentrations. It is suggested that the growth-rate disadvantage of the wild-type strain is caused by inhibitory effects of an intermediate in the degradation of benzoate via the plasmid-encoded meta-pathway.  相似文献   

8.
Some plasmid-free Tol- strains derived from Pseudomonas putida PAW1 (which carries the TOL plasmid pWW0) have a segment of TOL DNA located chromosomally. Of three independently isolated strains, PAW86 had an integrated TOL segment of 16 kilobases and PAW85 had two copies of this segment in different chromosomal locations, whereas the chromosomal DNA of PAW82 showed no homology with the TOL plasmid. In cultures of the parental strain, it appears that a 56-kilobase TOL DNA segment is located chromosomally in some cells.  相似文献   

9.
The TOL plasmid originally isolated in Pseudomonas putida (arvilla) mt-2 was transmissible to strains of the fluorescens group of Pseudomonas, i.e., P. putida, P. fluorescens, and P. aeruginosa, except for a strain of P. aeruginosa, strain PAO. The same strain, however, could accept the plasmid when its restriction and modification abilities were lost by mutations or by growing at high temperature. In addition, the transmissibility of the TOL plasmid from strain PAO to P. putida was low when the plasmid was modified by the donor. By using P. aeruginosa PAO carrying the TOL plasmid, the stability and genetic expression of the plasmid as well as its effect on the host cell growth were examined. Thus the self-maintenance of the plasmid was found to be thermosensitive. Furthermore, the TOL plasmid inhibited the growth of strain PAO at high temperature, accompanied by the formation of some filamentous cells. These thermosensitive properties of the TOL plasmid were host dependent and not exhibited in another strain of P. aeruginosa.  相似文献   

10.
Retrotransfer of DNA refers to the phenomenon by which a plasmid travels from a host strain to a recipient one and returns to the original host, bringing with it DNA from the recipient. The resultant host strain with DNA from the recipient is called a retrotransconjugant. The retrotransfer phenomenon mediated by the TOL plasmid pWW0 and other plasmids has been documented on plates under optimal laboratory culture conditions, but never under natural conditions. In this work, we show that retrotransfer mediated by the IncP9 TOL pWW0 plasmid occurs in the rhizosphere, a niche in which the continuous supply of nutrients via root exudates allows cells to reach a high density. This suggests that this unusual sexual fertilization may be of great importance in lateral gene transfer. We also show that retrotransfer of DNA seems to require co-integration of the plasmid and the host chromosome and subsequent resolution, because a TOL plasmid with a mutation in the tnpR gene, encoding the resolvase of the Tn 4653 of the TOL plasmid, was self-transferred between Pseudomonas strains, but unable to mobilize chromosome.  相似文献   

11.
The TOL plasmid is naturally derepressed for transfer   总被引:5,自引:0,他引:5  
Pseudomonas putida mt-2, formerly known as Pseudomonas arvilla mt-2, which carries the wild-type TOL plasmid, and P. putida strain AC37 carrying TOL, were completely lysed by the pilus-adsorbing plasmid-specific bacteriophages PR4 and PRD1. Pseudomonas putida strain PpS388, also harbouring the plasmid, was not lysed. In a P. putida mt-2 host, TOL transferred 18-fold better on a surface (2.5 X 10(-1) transconjugants per donor h-1) than in liquid; when P. putida PpS388 was the host, however, a frequency of only 2.3 X 10(-4) transconjugants per donor h-1 was obtained. Thus, TOL was derepressed for transfer in P. putida mt-2 and P. putida AC37, but not in P. putida PpS388. Electron microscopy revealed that TOL determined thick (8.5-10 nm diameter) flexible pili in large numbers, suggesting constitutive expression in its derepressed state.  相似文献   

12.
Chromosomal location of TOL plasmid DNA in Pseudomonas putida.   总被引:11,自引:6,他引:5       下载免费PDF全文
The soil isolate Pseudomonas putida MW1000 can grow on toluene and other hydrocarbons; in this respect it is similar to strains of Pseudomonas which carry the TOL plasmid. By conjugation experiments, the genes conferring these growth abilities have been shown to be located on the bacterial chromosome, linked to vil and catB. A 56-kilobase segment of the bacterial chromosome of MW strains carrying the TOL genes can transpose to the IncP-1 plasmid R18-18. Physical analysis of these TOL R18-18 hybrids has shown that the TOL segment is almost identical to the same region found in the TOL plasmid pWW0.  相似文献   

13.
The upper operon of the TOL plasmid pWW0 of Pseudomonas putida encodes a set of enzymes involved in the conversion of toluene and xylenes to their carboxylic acid derivatives. The last gene of the upper operon, xylN, encodes a 465-amino-acid polypeptide which exhibits significant sequence similarity to FadL, an outer membrane protein involved in fatty acid transport in Escherichia coli. To analyze the role of the xylN gene product, xylN on TOL plasmid pWW0 was disrupted by inserting a kanamycin resistance gene, and the phenotypes of P. putida harboring the wild-type and xylN mutant TOL plasmids were characterized. The growth of P. putida harboring the wild-type TOL plasmid was inhibited by a high concentration of m-xylene, while that of P. putida harboring the xylN mutant TOL plasmid was not. The apparent K(s) value for the oxidation of m-xylene in intact cells of the xylN mutant was fourfold higher than that of the wild-type strain, although the TOL catabolic enzyme activities in cell extracts from the two strains were almost identical. We therefore presume that the xylN gene product is a porin involved in the transport of m-xylene and its analogues across the outer membrane. Western blot analysis confirmed the localization of XylN in the outer membrane.  相似文献   

14.
Summary Growth in a chemostat of the 3-chlorobenzoatepositive Pseudomonas putida cells harboring the plasmid pAC25, in presence of cells harboring the TOL plasmid, allows emergence of cells that can also utilize 4-chlorobenzoate (4Cba). Isolation of plasmid DNA from such cells demonstrate the deletion of a 11kb (Kilobase pair) EcoR1 fragment from the pAC25 plasmid; a portion of the TOL plasmid (41.5 kb TOL*) is also found to be transposed onto the chromosome of such cells. Further enrichment of the 4-chlorobenzoate-positive cells with 3,5-dichlorobenzoate (3,5-Dcb) as a sole carbon source has produced cells that can also slowly utilize 3,5-dichlorobenzoate. Isolation of plasmid DNA from such cells demonstrates the appearance of a second plasmid (pAC29). Restriction hybridization of pAC29 EcoRI fragments with pAC25 and TOL demonstrates that pAC29 is derived primarily by duplication of a segment of the pAC27 plasmid and a fragment from TOL, with further mutational divergence. Southern hybridization of the EcoRI-digested chromosomal DNA with 32P-labeled TOL, pTS11 and pTS71 plasmid DNAs demonstrates the presence of the TOL* transposon containing xylD, G, E and F genes in both 4Cba+ (pAC27+) and 3,5-DCb+ (pAC27+, pAC29+) cells. Isolation of plasmid DNA from 3,5-Dcb+ faster growing variants, obtained from slow-growing pAC27+ pAC29+ cells, demonstrates the presence of a single type of plasmid, with identical size and EcoRI digestion profile as pAC27. The implications of gene duplications and subsequent homologous recombination with regard to the biochemical pathway of 3,5-dichlorobenzoate degradation have been discussed.  相似文献   

15.
Use of plasmid DNA (pDNA) in the emerging gene therapy requires pure DNA in large quantities requiring production of safe DNA on large scale. While a number of kit-based DNA purification techniques have become popular, large scale cost effective purification of DNA remains a technological challenge. Most traditional, as well as newly developed methods for DNA purification are expensive, tedious, use toxic reagents, and/or generally not amenable for scaled up production. Our attempts to develop a scalable adsorptive separation technology resulted in successful use of indigenously developed rigid cross-linked cellulose beads for single step purification of pDNA from alkaline cell lysates. This mode of purification employs a combination of intra-particle interactions that could give a product plasmid DNA free from chromosomal DNA, RNA and host proteins in a single scalable chromatographic step. The technology can be employed as a batch adsorption step on small scale, or on a large scale column chromatography. A high copy number 9.8 kb plasmid (from an Escherichia coli strain) was purified in yields of 77 and 52%, respectively in batch and column modes. The product obtained was homogeneous supercoiled plasmid with no RNA and protein contamination confirmed by quantitative analysis, agarose gel electrophoresis and SDS-PAGE.  相似文献   

16.
The TOL catabolic plasmid was shown to be compatible with the R91 drug resistance plasmid. However, the TOL plasmid was extremely unstable in mutant PA03 of P. aeruginosa. By selecting for stabilization of the TOL plasmid in PA03 harbouring R91, it was possible to isolate a strain in which markers from both R91 and TOL appeared to exist in a single recombinant plasmid. This plasmid, pND3, encoded resistance to carbenicillin, was able to transfer at the same frequency as the R91 plasmid and encoded the ability to grow on m-toluate, p-toluate, m-xylene, p-xylene and toluene. In addition, it was shown to be incompatible with the NAH catabolic plasmid and it could be transferred by transduction. The TOL plasmid could stabilize in PA03 harbouring R91 without fusion with R91, and could stabilize in PA03 in the absence of R91. PA03 harbouring either the recombinant plasmid or the stable TOL plasmid in the absence of R91 could promote bacterial chromosome transfer between mutant derivatives of P. aeruginosa strain PA0.  相似文献   

17.
Summary TOL plasmid pWW0 and plasmid NAH7 encode catabolic enzymes required for oxidative degradation of toluene and naphthalene, respectively. The gene order of the catabolic operon of NAH7 for salicylate oxidation was determined to be: promoter-nahG (the structural gene for salicylate hydroxylase)-nahH (catechol 2,3-dioxygenase)-nahI (hydroxymuconic semialdehyde dehydrogenase)-nahN (hydroxymuconic semialdehyde hydrolase)-nahL (2-oxopent-4-enoate hydratase). This order is identical to that of the isofunctional genes of TOL plasmid pWW0. The complete nucleotide sequence of nahH was determined and compared with that of xylE, the isofunctional gene of TOL plasmid pWW0. There were 20% and 16% differences in their nucleotide and amino acid sequences, respectively. The homology between the NAH7 and TOL pWW0 plasmids ends upstream of the Shine-Dalgarno sequences of nahH and xylE, but the homology continues downstream of these genes. This observation suggested that genes for the catechol oxidative enzymes of NAH7 and TOL pWW0 were derived from a common ancestral sequence which was transferred as a discrete segment of DNA between plasmids.  相似文献   

18.
Isolation of TOL and RP4 recombinants by integrative suppression.   总被引:4,自引:18,他引:4       下载免费PDF全文
We obtained genetic and molecular evidence of non-thermosensitive recombinants of RP4 (Kmr Tcr Cbr/Apr) and the thermosensitive TOL plasmid. As first isolated in Pseudomonas aeruginosa PAO, the recombinant plasmid pTN1 specified noninducible synthesis of TOL enzymes and was transmissible to Escherichia coli on selection for the transfer of kanamycin resistance. The phenotypic expression of TOL genes of pTN1 in E. coli was low and also noninducible. A spontaneous segregant, pTN2, appearing from pTN1, conferred inducible synthesis of TOL enzymes. These plasmids carry all of the TOL determinants as evidenced by the ability of Pseudomonas putida carrying recombinant plasmids to grow on toluene, xylene, and m-toluate. In E. coli the expression of TOL genes with normal regulation (pTN2) appears to be extremely low without induction, and the induced expression is comparable to that with defective regulation (pTN1). The measurement of the molecular weight of pTN2 by electron microscopy gave a value of about 74 X 10(6).  相似文献   

19.
The dimer catastrophe hypothesis has been proposed previously to explain instability of multicopy plasmids whose partitioning is random, contrary to low copy number plasmids which are stably maintained and actively partitioned. Until now, this hypothesis has been investigated using multicopy ColE1 plasmids. However, for more detailed testing of the dimer/multimer catastrophe hypothesis, one should use a plasmid which can be maintained at either low or high copy number and still possesses the same mechanism of replication regulation. Here we used a modified lambda plasmid, pTC lambda 1. The advantage of this plasmid is that it can be maintained at different copy numbers depending on the concentration of an inducer which stimulates the initiation of plasmid replication. Results obtained with this plasmid in recombination proficient and deficient cells generally support the dimer/multimer catastrophe hypothesis, but also suggest some modification in the model.  相似文献   

20.
Strains with greater ability to dissimilate m-toluate were obtained from the wild-type Pseudomonas putida (arvilla) mt-2 that harbors the TOL plasmid. Increased growth of a mutant strain on aromatic substrates was coupled with simultaneous increase in the activity of metapyrocatechase, an enzyme coded by the TOL plasmid, without changing its catalytic properties. In the mutant and the wild-type strains, the inducer specificity and the induction kinetics of metapyrocatechase synthesis were the same, and a half-maximal effect of m-toluate on the enzyme synthesis was observed at 0.25 mM. Thus, the increased utilizability seen in a mutant strain appeared to be due to an increased quantity of the enzymes coded by the TOL plasmid. The properties of the mutant strain were dependent upon the mutation on the TOL plasmid but not on the chromosome mutation. Transfer experiments with a strain carrying the mutant TOL (TOL-H) or the wild-type TOL plasmid revealed that the TOL-H transfer was 1,000 times greater than that of the wild type.  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