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.
Circularized regions of the chromosome containing the origin of replication, oriC, can be maintained as autonomous minichromosomes, oriC plasmids. We show that oriC plasmids containing precise, pre-determined segments of the chromosome can be generated by a simple in vivo recombineering technique. We generated two such plasmids carrying fluorescent markers. These were transferred to a recipient strain with a different fluorescent marker near the chromosomal copy of oriC. Thus the fates of the oriC plasmid and chromosomal origins could be followed independently in living cells by fluorescence microscopy. In contrast to a previous report, we show that there is a strong tendency of oriC plasmid copies to accumulate at the cell center as a single or double focus at the plane of cell division. This is not simply due to exclusion from the nucleoid space but rather appears to be a specific recognition and retention of the plasmid by some central-located cell site.  相似文献   

2.
Minichromosomes, i.e. plasmids that can replicate from an integrated oriC, have been puzzling because of their high copy numbers compared to that of the chromosomal oriC, their lack of incompatibility with the chromosome and their high loss frequencies. Using single cell resistance to tetracycline or ampicillin as an indicator of copy number we followed the development of minichromosome distributions in Escherichia coli cells transformed with minichromosomes and then allowed to grow towards the steady state. The final copy number distribution was not reached within 15 to 20 generations. If the minichromosome carried the sop (partitioning) genes from plasmid F, the development of the copy number distribution was further drastically delayed. We conclude that E. coli cells have no function that directly controls minichromosomal copy numbers, hence the absence of incompatibility in the sense of shared copy number control. We suggest that minichromosomes are subject to the same replication control as the chromosome but segregate randomly in the absence of integrated partitioning genes. This, combined with evidence that the lowest copy number classes are normally present despite high average copy numbers, can account for the high loss frequencies.  相似文献   

3.
We have estimated the statistical distribution of the number of plasmids taken up by individual Jurkat lymphoma cells during electroporation in the presence of two plasmids, one encoding for yellow (EYFP) the other for cyan (ECFP) fluorescent protein. The plasmid concentration at which most of the cells take up only one plasmid or several molecules was determined by statistical analysis. We found that cells behaved slightly heterogeneous in plasmid uptake and describe how the homogeneity of a cell population can be quantified by Poisson statistics in order to identify experimental conditions that yield homogeneously transfection-competent cell populations. The experimental procedure worked out with Jurkat cells was applied to assay the effectiveness of antisense RNA in knocking down gene expression in Physarum polycephalum. Double transfection of flagellates with vectors encoding EYFP and antisense-EYFP revealed for the first time that gene expression can be suppressed by co-expression of antisense RNA in Physarum. Quantitative analysis revealed that one copy of antisense expressing gene per EYFP gene was sufficient to completely suppress formation of the EYFP protein in Physarum.  相似文献   

4.
The random distribution of ColE1 plasmids between the daughter cells at cell division introduces large copy number variations. Statistic variation associated with limited copy number in single cells also causes fluctuations to emerge spontaneously during the cell cycle. Efficient replication control out of steady state is therefore important to tame such stochastic effects of small numbers. In the present model, the dynamic features of copy number control are divided into two parts: first, how sharply the replication frequency per plasmid responds to changes in the concentration of the plasmid-coded inhibitor, RNA I, and second, how tightly RNA I and plasmid concentrations are coupled. Single (hyperbolic)- and multiple (exponential)-step inhibition mechanisms are compared out of steady state and it is shown how the response in replication frequency depends on the mode of inhibition. For both mechanisms, sensitivity of inhibition is “bought” at the expense of a rapid turnover of a replication preprimer, RNA II. Conventional, single-step, inhibition kinetics gives a sloppy replication control even at high RNA II turnover rates, whereas multiple-step inhibition has the potential of working with unlimited precision. When plasmid concentration changes rapidly, RNA I must be degraded rapidly to be “up to date” with the change. Adjustment to steady state is drastically impaired when the turnover rate constants of RNA I decrease below certain thresholds, but is basically unaffected for a corresponding increase. Several features of copy number control that are shown to be crucial for the understanding of ColE1-type plasmids still remain to be experimentally characterized. It is shown how steady-state properties reflect dynamics at the heart of regulation and therefore can be used to discriminate between fundamentally different copy number control mechanisms. The experimental tests of the predictions made require carefully planned assays, and some suggestions for suitable experiments arise naturally from the present work. It is also discussed how the presence of the Rom protein may affect dynamic qualities of copy number control.  相似文献   

5.
The relationship between cell mass and cell number dynamics for bacteria such as Escherichia coli depends on the cell cycle parameters C and D. Effects of plasmid copy number on these cell cycle parameters have been studied for Escherichia coli HB101 containing pMB1 plasmids propagated at different copy numbers ranging from 12 to 122. Determination of cell cycle and cell size parameters was accomplished using flow cytometry data on single-cell light scattering and DNA content frequency functions in conjunction with a mathematical model of cell population statistics. Two independent methods for estimating C and D intervals based on flow cytometry were developed and applied with essentially identical results. The presence of plasmids decreases the C and D periods, mean cell sizes, and initiation masses for chromosome replication by 14, 24, 38, and 18%, respectively, relative to corresponding values for plasmid-free host cells. Plasmid copy number has a negligible influence on these parameters, suggesting that host-plasmid inter actions which determine these properties are centered on the single plasmid selected for replication according to the random selection model established for ColE1-type plasmids.  相似文献   

6.
Molecular clocks reduce plasmid loss rates: the R1 case   总被引:3,自引:0,他引:3  
Plasmids control their replication so that the replication frequency per plasmid copy responds to the number of plasmid copies per cell. High sensitivity amplification in replication response to copy number deviations generally reduces variation in copy numbers between different single cells, thereby reducing the plasmid loss rate in a cell population. However, experiments show that plasmid R1 has a gradual, insensitive replication control predicting considerable copy number variation between single cells. The critical step in R1 copy number control is regulation of synthesis of a rate-limiting cis-acting replication protein, RepA. De novo synthesis of a large number of RepA molecules is required for replication, suggesting that copy number control is exercised at multiple steps. In this theoretical kinetic study we analyse R1 multistep copy number control and show that it results in the insensitive replication response found experimentally but that it at the same time effectively prohibits the existence of only one plasmid copy in a dividing cell. In combination with the partition system of R1, this can lead to very high segregational stability. The R1 control mechanism is compared to the different multistep copy number control of plasmid ColE1 that is based on conventional sensitivity amplification. This implies that while copy number control for ColE1 efficiently corrects for fluctuations that have already occurred, R1 copy number control prevents their emergence in cells that by chance start their cycle with only one plasmid copy. We also discuss how regular, clock-like, behaviour of single plasmid copies becomes hidden in experiments probing collective properties of a population of plasmid copies because the individual copies are out of phase. The model is formulated using master equations, taking a stochastic approach to regulation, but the mathematical formalism is kept to a minimum and the model is simplified to its bare essence. This simplicity makes it possible to extend the analysis to other replicons with similar design principles.  相似文献   

7.
8.
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   

9.
Researchers often require customised variations of plasmids that are not commercially available. Here we demonstrate the applicability and versatility of standard synthetic biological parts (biobricks) to build custom plasmids. For this purpose we have built a collection of 52 parts that include multiple cloning sites (MCS) and common protein tags, protein reporters and selection markers, amongst others. Importantly, most of the parts are designed in a format to allow fusions that maintain the reading frame. We illustrate the collection by building several model contructs, including concatemers of protein binding-site motifs, and a variety of plasmids for eukaryotic stable cloning and chromosomal insertion. For example, in 3 biobrick iterations, we make a cerulean-reporter plasmid for cloning fluorescent protein fusions. Furthermore, we use the collection to implement a recombinase-mediated DNA insertion (RMDI), allowing chromosomal site-directed exchange of genes. By making one recipient stable cell line, many standardised cell lines can subsequently be generated, by fluorescent fusion-gene exchange. We propose that this biobrick collection may be distributed peer-to-peer as a stand-alone library, in addition to its distribution through the Registry of Standard Biological Parts (http://partsregistry.org/).  相似文献   

10.
S J Projan  R P Novick 《Plasmid》1984,12(1):52-60
An experimental analysis of the concept that incompatible plasmids occupy a common intracellular pool from which copies are drawn at random for replication and assortment is presented. Intrapool variations in an incompatible heteroplasmid strain are inevitable and it is shown that these variations can be exploited by differential selection to amplify one plasmid at the expense of the other. Constant overall copy number is demonstrated for isogenic wild-type replicons and also for isogenic copy mutants whose copy numbers are so great that segregational incompatibility cannot be measured. In the test system used, that of the Staphylococcus aureus plasmid pT181, the rate of replication is probably determined by the availability of a trans-active initiator protein, RepC. In heteroplasmid strains containing wild-type and dominant copy mutant plasmids, although intrapool variation occurs, the total copy number is not constant but varies as a consequence of selection for or against the mutant plasmid. This is because all of the RepC is synthesized from the mutant plasmid (the wild-type is hyper-repressed) and therefore the selection affects the supply of RepC at the same time that it affects the copy number of the plasmid. None of these effects are seen with single plasmids or with compatible pairs.  相似文献   

11.
The copy number per cell mass of plasmid pBR322 and a rom- derivative was measured as a function of generation time. In fast growing cells the copy number per cell mass was virtually identical for rom+ and rom- derivatives. However, the copy number of pBR322 only increased 3- to 4-fold from a 20- to 80-min generation time, whereas the copy number of the rom- derivative increased 7- to 10-fold. The copy number stayed constant for the rom+ and rom- plasmids at generation times longer than 80-100 min. Thus, the presence of the rom gene decreased the copy number of plasmid pBR322 in slowly growing cells at least 2-fold when compared with the rom- plasmid. To study the effect of the rom gene in trans we cloned the gene into the compatible P15A-derived rom- plasmid pACYC184. In cells carrying both pACYC184 rom+ and pBR322 rom- the presence of the rom gene in trans had little effect on the copy number of pBR322 rom- at fast growth, but it decreased its copy number at slow growth to the same level as found for pBR322, i.e., complemented the pBR322 rom- plasmid. The pACYC184 plasmid and its rom+ derivatives showed copy numbers similar to those of pBR322 rom- and pBR322 itself, respectively, at fast and slow growth. We conclude that the rom gene product-the Rom protein-is an important element in copy number control of ColE1-type plasmids especially in slowly growing cells.  相似文献   

12.
Efficient expression of a foreign protein product by the yeastSaccharomyces cerevisiaerequires a stable recombinant vector present at a high number of copies per cell. A conditional centromere yeast plasmid was constructed which can be amplified to high copy number by a process of unequal partitioning at cell division, followed by selection for increased copy number. However, in the absence of selection pressure for plasmid amplification, copy number rapidly drops from 25 plasmids/cell to 6 plasmids/cell in less than 10 generations of growth. Copy number subsequently decreases from 6 plasmids/cell to 2 plasmids/cell over a span of 50 generations. A combination of flow cytometric measurement of copy number distributions and segregated mathematical modeling were applied to test the predictions of a conceptual model of conditional centromereplasmid propagation. Measured distributions of plasmid content displayed a significant subpopulation of cells with a copy number of 4-6, evenin a population whose mean copy number was 13.5. This type of copy number distribution was reproduced by a mathematical model which assumes that amaximum of 4-6 centromere plasmids per cell can be stably partitionedat cell division. The model also reproduces the observed biphasic kinetics of plasmid number instability. The agreement between simulation and experimental results provides support for the proposed model and demonstrates the utility of the flow cytometry/segregated modeling approach for the study of multicopy recombinant vector propagation.  相似文献   

13.
Although plasmid copy number varies widely among different plasmid species, normally copy number is maintained within a narrow range for any given plasmid. Such copy number control has been shown to occur by regulation of the rate of plasmid DNA replication. Here we report a novel mechanism by which the pSC101 plasmid also can detect an imbalance between the cellular level of its replication protein, RepA, and plasmid-borne RepA binding sites to inhibit bacterial DNA replication and delay host cell division when RepA is in relative excess. We show that delayed cell division occurs by RepA-mediated induction of the SOS response and can be reversed by over-expression of the host DNA primase, DnaG. The effects of RepA excess are prevented by introducing a surfeit of RepA binding sites. The mechanism reported here may help to limit variation in plasmid copy number and allow repopulation of cells with plasmids when copy number falls--potentially pre-empting plasmid loss in cultures of dividing cells.  相似文献   

14.
Toxic effects of excess cloned centromeres.   总被引:24,自引:7,他引:17       下载免费PDF全文
Plasmids carrying a Saccharomyces cerevisiae centromere have a copy number of one or two, whereas other yeast plasmids have high copy numbers. The number of CEN plasmids per yeast cell was made artificially high by transforming cells simultaneously with several different CEN plasmids carrying different, independently selectable markers. Some host cells carried five different CEN plasmids and an average total of 13 extra copies of CEN3. Several effects were noted. The copy number of each plasmid was unexpectedly high. The plasmids were mutually unstable. Cultures contained many dead cells. The viable host cells grew more slowly than control cells, even in nonselective medium. There was a pause in the cell cycle at or just before mitosis. We conclude that an excess of centromeres is toxic and that the copy number of centromere plasmids is low partly because of selection against cells carrying multiple centromere plasmids. The toxicity may be caused by competition between the centromeres for some factor present in limiting quantities, e.g., centromere-binding proteins, microtubules, or space on the spindle pole body.  相似文献   

15.
Chen DQ  Zheng XC  Lu YJ 《Plasmid》2006,56(3):167-178
ColE1-type plasmids are commonly used in bacterial genetics research, and replication of these plasmids is regulated by interaction of RNA I and RNA II. Although these plasmids are narrow-host-range, they can be maintained in Legionella pneumophila under antibiotic selection, with low-copy number and instability. Here, we have described the isolation of two novel spontaneous mutants of pBC(gfp)Pmip, pBG307 and pBG309, which are able to mark the L. pneumophila with strong green fluorescence when exposed to visible light. One of the mutants, pBG307, has a single CG-->TA mutation in RNA II promoter located 2-bases upstream the - 10 region. Another one, pBG309, has the same mutation, as well as an additional CG-->AT mutation in the 76th nucleotide of RNA I, or in the 6th nucleotide of RNA II. A plasmid with the single mutation in RNA I, pBG308, was also constructed. Characterization of these plasmids carrying the enhanced green fluorescent protein (gfpmut2) gene revealed that the green fluorescence intensities of these plasmids were 2- to 30-fold higher than that of the wild type and both of the mutations contribute to increase the plasmid copy number and/or plasmid stability. The mutation located in RNA II promoter played a more dominant role in elevating the copy number, compared to the mutation in RNA I. We also tested the mutant plasmids for replication in Escherichia coli, and found that their copy number and stability were dramatically decreased, except pBG307. Our data suggest that these plasmids might be useful and convenient in genetic studies in L. pneumophila.  相似文献   

16.
Plasmid repopulation kinetics in Staphylococcus aureus   总被引:7,自引:0,他引:7  
We have analyzed the kinetic route by which the indirectly controlled Staphylococcus aureus plasmid, pT181, responds to and corrects fluctuations in copy number. The kinetics of copy number correction from low to steady-state levels (termed repopulation) were determined using two different methods of copy number reduction. Thermosensitive replication (Tsr) mutants of pT181 were grown at nonpermissive temperatures to lower copy number and then shifted to a permissive temperature to allow repopulation. After the downshift, both wild-type and copy mutant plasmids, with active inhibitors, exhibited a burst of exponential replication that resulted in a two- to threefold overshoot of normal steady-state copy numbers. This was followed by inhibition of replication and eventual reestablishment of the steady-state replication rate. Similar replication kinetics were observed when these plasmids were introduced into naive cells by high-frequency transduction. By contrast, a pT181 copy mutant with a nonfunctional inhibitor-target regulation did not overshoot its steady-state copy number, but instead repopulated asymptotically. These results suggest that at low copy numbers, pT181 and its derivatives replicate at near-maximal rates and overshoot prior to the establishment of an inhibitory concentration of repressor. The maximal replication rate is independent of the plasmid's cop genotype. As the copy number increases, inhibitor accumulates and eventually reduces the replication rate. In the absence of an active inhibitor, the steady-state copy number is established at a level that must be limited by some other invariant factor.  相似文献   

17.
18.
Plasmid-borne gene expression systems have found wide application in the emerging fields of systems biology and synthetic biology, where plasmids are used to implement simple network architectures, either to test systems biology hypotheses about issues such as gene expression noise or as a means of exerting artificial control over a cell's dynamics. In both these cases, fluorescent proteins are commonly applied as a means of monitoring the expression of genes in the living cell, and efforts have been made to quantify protein expression levels through fluorescence intensity calibration and by monitoring the partitioning of proteins among the two daughter cells after division; such quantification is important in formulating the predictive models desired in systems and synthetic biology research. A potential pitfall of using plasmid-based gene expression systems is that the high protein levels associated with expression from plasmids can lead to the formation of inclusion bodies, insoluble aggregates of misfolded, nonfunctional proteins that will not generate fluorescence output; proteins caught in these inclusion bodies are thus "dark" to fluorescence-based detection methods. If significant numbers of proteins are incorporated into inclusion bodies rather than becoming biologically active, quantitative results obtained by fluorescent measurements will be skewed; we investigate this phenomenon here. We have created two plasmid constructs with differing average copy numbers, both incorporating an unregulated promoter (P(LtetO-1) in the absence of TetR) expressing the GFP derivative enhanced green fluorescent protein (EGFP), and inserted them into Escherichia coli bacterial cells (a common model organism for work on the dynamics of prokaryotic gene expression). We extracted the inclusion bodies, denatured them, and refolded them to render them active, obtaining a measurement of the average number of EGFP per cell locked into these aggregates; at the same time, we used calibrated fluorescent intensity measurements to determine the average number of active EGFP present per cell. Both measurements were carried out as a function of cellular doubling time, over a range of 45-75 min. We found that the ratio of inclusion body EGFP to active EGFP varied strongly as a function of the cellular growth rate, and that the number of "dark" proteins in the aggregates could in fact be substantial, reaching ratios as high as approximately five proteins locked into inclusion bodies for every active protein (at the fastest growth rate), and dropping to ratios well below 1 (for the slowest growth rate). Our results suggest that efforts to compare computational models to protein numbers derived from fluorescence measurements should take inclusion body loss into account, especially when working with rapidly growing cells.  相似文献   

19.
During the past 2 decades, Bacillus megaterium has been systematically developed for the gram-per-liter scale production of recombinant proteins. The plasmid-based expression systems employed use a xylose-controlled promoter. Protein production analyses at the single-cell level using green fluorescent protein as a model product revealed cell culture heterogeneity characterized by a significant proportion of less productive bacteria. Due to the enormous size of B. megaterium, such bistable behavior seen in subpopulations was readily analyzed by time lapse microscopy and flow cytometry. Cell culture heterogeneity was not caused simply by plasmid loss: instead, an asymmetric distribution of plasmids during cell division was detected during the exponential-growth phase. Multicopy plasmids are generally randomly distributed between daughter cells. However, in vivo and in vitro experiments demonstrated that under conditions of strong protein production, plasmids are retained at one of the cell poles. Furthermore, it was found that cells with accumulated plasmids and high protein production ceased cell division. As a consequence, the overall protein production of the culture was achieved mainly by the subpopulation with a sufficient plasmid copy number. Based on our experimental data, we propose a model whereby the distribution of multicopy plasmids is controlled by polar fixation under protein production conditions. Thereby, cell lines with fluctuating plasmid abundance arise, which results in population heterogeneity. Our results provide initial insights into the mechanism of cellular heterogeneity during plasmid-based recombinant protein production in a Bacillus species.  相似文献   

20.
We characterized a number of widely used yeast-Escherichia coli shuttle vectors in the fission yeast Schizosaccharomyces pombe. The 2 micron vectors pDB248 and YEp13 showed high frequency of transformation, intermediate mitotic and low meiotic stability, and a low copy number in S. pombe, analogous to their behavior in [cir0] strains of Saccharomyces cerevisiae. The S. cerevisiae integration vectors pLEU2 and pURA3 transformed S. pombe at very low frequencies but, surprisingly, in a nonintegrative fashion. Instead, they replicated autonomously, and they showed very high copy numbers (up to 150 copies per plasmid-containing cell). This could reflect a lack of sequence specificity for replication of plasmid DNA in S. pombe. pFL20, an S. pombe ars vector, and a series of plasmids derived from it were studied to analyze the unusually high stability of this plasmid. Mitotic stability and partitioning of the plasmids was measured by pedigree analysis of transformed S. pombe cells. An S. pombe DNA fragment (stb) was identified that stabilizes pFL20 by improvement of plasmid partitioning in mitosis and meiosis.  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