首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
2.
Meiotic recombination within the centromere of a yeast chromosome   总被引:6,自引:0,他引:6  
L S Symington  T D Petes 《Cell》1988,52(2):237-240
In order to examine the frequency of nonreciprocal recombination (gene conversion) within the centromere of the yeast chromosome, we constructed strains that contained heterozygous restriction sites in the conserved centromere sequences of chromosome III in addition to heterozygous markers flanking the centromere. One of these markers was the selectable URA3 gene, which was inserted less than one kb from the centromere. We found that meiotic conversion of the URA3 gene occurred at normal frequency (about 2% of unselected tetrads) and that more than one-third of these convertants coconverted the markers within the centromere. In addition, we observed tetrads in which conversion events extended through the centromere to include a marker on the opposite side from URA3. We conclude that meiotic conversion events occur within the centromere at rates similar to other genomic sequences.  相似文献   

3.
Mammalian cells are able to repair chromosomal double-strand breaks (DSBs) both by homologous recombination and by mechanisms that require little or no homology. Although spontaneous homologous recombination is rare, DSBs will stimulate recombination by 2 to 3 orders of magnitude when homology is provided either from exogenous DNA in gene-targeting experiments or from a repeated chromosomal sequence. Using a gene-targeting assay in mouse embryonic stem cells, we now investigate the effect of heterology on recombinational repair of DSBs. Cells were cotransfected with an endonuclease expression plasmid to induce chromosomal DSBs and with substrates containing up to 1.2% heterology from which to repair the DSBs. We find that heterology decreases the efficiency of recombinational repair, with 1.2% sequence divergence resulting in an approximately sixfold reduction in recombination. Gene conversion tract lengths were examined in 80 recombinants. Relatively short gene conversion tracts were observed, with 80% of the recombinants having tracts of 58 bp or less. These results suggest that chromosome ends in mammalian cells are generally protected from extensive degradation prior to recombination. Gene conversion tracts that were long (up to 511 bp) were continuous, i.e., they contained an uninterrupted incorporation of the silent mutations. This continuity suggests that these long tracts arose from extensive degradation of the ends or from formation of heteroduplex DNA which is corrected with a strong bias in the direction of the unbroken strand.  相似文献   

4.
Y Sakai  Y Tani 《Journal of bacteriology》1992,174(18):5988-5993
A model system for one-step gene disruption for an asporogenous methylotrophic yeast, Candida boidinii, is described. In this system, the 3-isopropylmalate dehydrogenase gene (C. boidinii LEU2) was selected as the target gene for disruption to derive new host strains for transformation. First, the C. boidinii LEU2 gene was cloned, and its complete nucleotide sequence was determined. Next, the LEU2 disruption vectors, which had the C. boidinii URA3 gene as the selectable marker, were constructed. Of the Ura+ transformants obtained with these plasmids, more than half showed a Leu- phenotype. Finally, the double-marker strains of C. boidinii were derived. When vectors with repeated flanking sequences of the C. boidinii URA3 gene were used for gene disruption, Leu- Ura+ transformants changed spontaneously to a Leu- Ura- phenotype ca. 100 times more frequently than they did when plasmids without the repeated sequences were used. Southern analysis showed that these events included a one-step gene disruption and a subsequent popping out of the C. boidinii URA3 sequence from the transformant chromosome.  相似文献   

5.
Chromosomal double-strand breaks (DSBs) stimulate homologous recombination by several orders of magnitude in mammalian cells, including murine embryonic stem (ES) cells, but the efficiency of recombination decreases as the heterology between the repair substrates increases (B. Elliott, C. Richardson, J. Winderbaum, J. A. Nickoloff, and M. Jasin, Mol. Cell. Biol. 18:93-101, 1998). We have now examined homologous recombination in mismatch repair (MMR)-defective ES cells to investigate both the frequency of recombination and the outcome of events. Using cells with a targeted mutation in the msh2 gene, we found that the barrier to recombination between diverged substrates is relaxed for both gene targeting and intrachromosomal recombination. Thus, substrates with 1.5% divergence are 10-fold more likely to undergo DSB-promoted recombination in Msh2(-/-) cells than in wild-type cells. Although mutant cells can repair DSBs efficiently, examination of gene conversion tracts in recombinants demonstrates that they cannot efficiently correct mismatched heteroduplex DNA (hDNA) that is formed adjacent to the DSB. As a result, >20-fold more of the recombinants derived from mutant cells have uncorrected tracts compared with recombinants from wild-type cells. The results indicate that gene conversion repair of DSBs in mammalian cells frequently involves mismatch correction of hDNA rather than double-strand gap formation. In cells with MMR defects, therefore, aberrant recombinational repair may be an additional mechanism that contributes to genomic instability and possibly tumorigenesis.  相似文献   

6.
S. R. Judd  T. D. Petes 《Genetics》1988,118(3):401-410
Physical lengths of gene conversion tracts for meiotic and mitotic conversions were examined, using the same diploid yeast strain in all experiments. This strain is heterozygous for a mutation in the URA3 gene as well as closely linked restriction site markers. In cells that had a gene conversion event at the URA3 locus, it was determined by Southern analysis which of the flanking heterozygous restriction sites had co-converted. It was found that mitotic conversion tracts were longer on the average than meiotic tracts. About half of the tracts generated by spontaneous mitotic gene conversion included heterozygous markers 4.2 kb apart; none of the meiotic conversions included these markers. Stimulation of mitotic gene conversion by ultraviolet light or methylmethanesulfonate had no obvious effect on the size or distribution of the tracts. Almost all conversion tracts were continuous.  相似文献   

7.
In this paper, we describe the movement of a genetically marked Saccharomyces cerevisiae transposon. Ty912(URA3), to new sites in the S. cerevisiae genome. Ty912 is an element present at the HIS4 locus in the his4-912 mutant. To detect movement of Ty912, this element has been genetically marked with the S. cerevisiae URA3 gene. Movement of Ty912(URA3) occurs by recombination between the marked element and homologous Ty elements elsewhere in the S. cerevisiae genome. Ty912(URA3) recombines most often with elements near the HIS4 locus on chromosome III, less often with Ty elements elsewhere on chromosome III, and least often with Ty elements on other chromosomes. These recombination events result in changes in the number of Ty elements present in the cell and in duplications and deletions of unique sequence DNA.  相似文献   

8.
One-step gene replacement in yeast by cotransformation   总被引:21,自引:0,他引:21  
H Rudolph  I Koenig-Rauseo  A Hinnen 《Gene》1985,36(1-2):87-95
A general method to replace chromosomal DNA sequences of Saccharomyces cerevisiae by any in vitro modified DNA sequence has been developed and was applied to the PHO5 locus on chromosome II. A recipient strain was constructed in which part of the chromosomal PHO5 sequence was substituted by the URA3 gene. Replacement of this pho5-URA3 substitution by pho5 mutant alleles was achieved in one step by cotransformation with a pho5 DNA fragment and the self-replicating plasmid YEp13, which contains the LEU2 gene as a selectable marker. Leu+ transformants were selected, and the replacement events at the PHO5 locus were detected by their Ura- phenotype (1-4% of the Leu+ were Ura-). In a similar way the PHO5 coding sequence was replaced by the sequence coding for human tissue-type plasminogen activator (t-PA).  相似文献   

9.
We examined the genetic control of the activity of HOT1, a cis-acting recombination-stimulatory sequence of Saccharomyces cerevisiae. Mutations in RAD1 and RAD52 decrease the ability of HOT1 to stimulate intrachromosomal recombination while mutations in RAD4 and RAD50 do not affect HOT1 activity. In rad1 delta strains, the stimulation of excisive recombination by HOT1 is decreased while the rate of gene replacement is not affected. In rad52-8 strains the ability of HOT1 to stimulate both excisive recombination and gene replacement is decreased. All of the recombinants in the rad52-8 strains that would be categorized as gene replacements based on their phenotype are diploids apparently derived by endomitosis and excisive recombination. Studies on rad1 delta rad52-8 strains show that these mutations interact synergistically in the presence or absence of HOT1, resulting in low levels of recombination. The rate of gene replacement but not excisive recombination is stimulated by HOT1 in rad1 delta rad52-8 strains. Taken together, the results show that HOT1 stimulates exchange using multiple recombination pathways. Some of the activity of HOT1 is RAD1-dependent, some is RAD52-dependent, and some requires either RAD1 or RAD52 as suggested by the synergistic interaction found in double mutant strains. There is also a component of HOT1 activity that is independent of both RAD1 and RAD52.  相似文献   

10.
Centromeres are critically important for chromosome stability and integrity. Most eukaryotes have regional centromeres that include long tracts of repetitive DNA packaged into pericentric heterochromatin. Neocentromeres, new sites of functional kinetochore assembly, can form at ectopic loci because no DNA sequence is strictly required for assembly of a functional kinetochore. In humans, neocentromeres often arise in cells with gross chromosome rearrangements that rescue an acentric chromosome. Here, we studied the properties of centromeres in Candida albicans, the most prevalent fungal pathogen of humans, which has small regional centromeres that lack pericentric heterochromatin. We functionally delimited centromere DNA on Chromosome 5 (CEN5) and then replaced the entire region with the counter-selectable URA3 gene or other marker genes. All of the resulting cen5Δ::URA3 transformants stably retained both copies of Chr5, indicating that a functional neocentromere had assembled efficiently on the homolog lacking CEN5 DNA. Strains selected to maintain only the cen5Δ::URA3 homolog and no wild-type Chr5 homolog also grew well, indicating that neocentromere function is independent of the presence of any wild-type CEN5 DNA. Two classes of neocentromere (neoCEN) strains were distinguishable: “proximal neoCEN” and “distal neoCEN” strains. Neocentromeres in the distal neoCEN strains formed at loci about 200–450 kb from cen5Δ::URA3 on either chromosome arm, as detected by massively parallel sequencing of DNA isolated by CENP-ACse4p chromatin immunoprecipitation (ChIP). In the proximal neoCEN strains, the neocentromeres formed directly adjacent to cen5Δ::URA3 and moved onto the URA3 DNA, resulting in silencing of its expression. Functional neocentromeres form efficiently at several possible loci that share properties of low gene density and flanking repeated DNA sequences. Subsequently, neocentromeres can move locally, which can be detected by silencing of an adjacent URA3 gene, or can relocate to entirely different regions of the chromosome. The ability to select for neocentromere formation and movement in C. albicans permits mechanistic analysis of the assembly and maintenance of a regional centromere.  相似文献   

11.
P. Zhao  E. Kafer 《Genetics》1992,130(4):717-728
Methyl methane-sulfonate (MMS)-sensitive, radiation-induced mutants of Aspergillus were shown to define nine new DNA repair genes, musK to musS. To test mus mutations for effects on mitotic recombination, intergenic crossing over was assayed between color markers and their centromeres, and intragenic recombination between two distinguishable adE alleles. Of eight mutants analyzed, four showed significant deviations from mus+ controls in both tests. Two mutations, musK and musL, reduced recombination, while musN and musQ caused increases. In contrast, musO diploids produced significantly higher levels only for intragenic recombination. Effects were relatively small, but averages between hypo- and hyperrec mus differed 15-20-fold. In musL diploids, most of the rare color segregants resulted from mitotic malsegregation rather than intergenic crossing over. This indicates that the musL gene product is required for recombination and that DNA lesions lead to chromosome loss when it is deficient. In addition, analysis of the genotypes of intragenic (ad+) recombinants showed that the musL mutation specifically reduced single allele conversion but increased complex conversion types (especially recombinants homozygous for ad+). Similar analysis revealed differences between the effects of two hyperrec mutations; musN apparently caused high levels solely of mitotic crossing over, while musQ increased various conversion types but not reciprocal crossovers. These results suggest that mitotic gene conversion and crossing over, while generally associated, are affected differentially in some of the mus strains of Aspergillus nidulans.  相似文献   

12.
Because Candida albicans is a diploid organism, two consecutive steps of gene disruption are required to generate a gene knock-out. The same marker (URA3) is often used for disruption of both copies of the gene. This is possible because, after the first round of disruption, homologous recombination between direct repeats flanking the URA3 marker and the subsequent counterselection allow for the efficient recovery of Ura- revertants. Unfortunately, the URA-blaster disruption cassette cannot be used in a PCR-based disruption approach. The hisG repeats flanking the URA3 gene in the disruption cassette anneal to one another during PCR and thereby prevent amplification of the complete cassette. We explored the use of transformation based on split-marker recombination to circumvent this problem. To avoid any cloning steps and to retain the advantage of long flanking regions for disruption, we combined this with a PCR- and ligation-mediated approach for generating marker cassettes. We used this approach to disrupt the C. albicans FAL1 (ATP-dependent RNA helicase) gene. Long 5' and 3' FAL1-specific regions were amplified by PCR and individually ligated to a URA-blaster cassette. The resulting ligation reactions were used separately as templates to generate two FAL1 disruption cassettes with overlapping URA3 marker regions. Simultaneous transformation with both overlapping disruption cassettes yielded efficient disruption of one FAL1 allele.  相似文献   

13.
Using a genetic system of haploid strains of Saccharomyces cerevisiae carrying a duplication of the his4 region on chromosome III, the pso3-1 mutation was shown to decrease the rate of spontaneous mitotic intrachromosomal recombination 2- to 13-fold. As previously found for the rad52-1 mutant, the pso3-1 mutant is specifically affected in mitotic gene conversion. Moreover, both mutations reduce the frequency of spontaneous recombination. However, the two mutations differ in the extent to which they affect recombination between either proximally or distally located markers on the two his4 heteroalleles. In addition, amplifications of the his4 region were detected in the pso3-1 mutant. We suggest that the appearance of these amplifications is a consequence of the inability of the pso3-1 mutant to perform mitotic gene conversion.  相似文献   

14.
Using a genetic system of haploid strains of Saccharomyces cerevisiae carrying a duplication of the his4 region on chromosome III, the pso3-1 mutation was shown to decrease the rate of spontaneous mitotic intrachromosomal recombination 2- to 13-fold. As previously found for the rad52-1 mutant, the pso3-1 mutant is specifically affected in mitotic gene conversion. Moreover, both mutations reduce the frequency of spontaneous recombination. However, the two mutations differ in the extent to which they affect recombination between either proximally or distally located markers on the two his4 heteroalleles. In addition, amplifications of the his4 region were detected in the pso3-1 mutant. We suggest that the appearance of these amplifications is a consequence of the inability of the pso3-1 mutant to perform mitotic gene conversion.  相似文献   

15.
C. R. Preston  W. R. Engels 《Genetics》1996,144(4):1611-1622
A P-element insertion flanked by 13 restriction fragment length polymorphism (RFLP) marker sites was used to examine male recombination and gene conversion at an autosomal site. The great majority of crossovers on chromosome arm 2R occurred within the 4-kb region containing the P element and RFLP sites. Of the 128 recombinants analyzed, approximately two-thirds carried duplications or deletions flanking the P element. These rearrangements are described in more detail in the accompanying report. In a parallel experiment, we examined 91 gene conversion tracts resulting from excision of the same autosomal P element. We found the average tract length was 1463 bp, which is essentially the same as found previously at the white locus. The distribution of conversion tract endpoints was indistinguishable from the distribution of crossover points among the nonrearranged male recombinants. Most recombination events can be explained by the ``hybrid element insertion' model, but, for those lacking a duplication or deletion, a second step involving double-strand gap repair must be postulated to explain the distribution of crossover points.  相似文献   

16.
Summary Psoralen photoreaction with DNA produces interstrand crosslinks, which require the activity of excision and recombinational pathways for repair. Yeast replicating plasmids, carrying the HIS3, TRP1, and URA3 genes, were photoreacted with psoralen in vitro and transfected into Saccharomyces cerevisiae cells. Repair was assayed as the relative transformation efficiency. A recombination-deficient rad52 strain was the least efficient in the repair of psoralen-damaged plasmids; excision repair-deficient rad1 and rad3 strains had repair efficiencies intermediate between those of rad52 and RAD cells. The level of repair also depended on the conditions of transformant selection; repair was more efficient in medium lacking tryptophan than in medium from which either histidine or uracil was omitted. The plasmid repair differential between these selective media was greatest in rad1 cells, and depended on RAD52. Plasmid-chromosome recombination was stimulated by psoralen damage, and required RAD52 function. Chromosome to plasmid gene conversion was seen most frequently at the HIS3 locus. In RAD and rad3 cells, the majority of the conversions were associated with plasmid integration, while in rad1 cells most were non-crossover events. Plasmid to chromosome gene conversion was observed most frequently at the TRP1 locus, and was accompanied by plasmid loss.  相似文献   

17.
To examine the mechanism of recombination and the role of DNA repair in this process, we transfected a plasmid carrying duplicated copies of the Herpes simplex virus I thymidine kinase (Htk) gene, each containing an 8 bp XhoI site inserted in a unique site and with the neo coding for geneticin resistance located between them, into tk-deficient human cell lines which differ in their ability to carry out nucleotide excision repair. One parental cell line has a normal level of repair activity; the second has an intermediate level, and the third has virtually no repair activity. Several geneticin-resistant transfectant cell strains from each parental line were isolated and assayed for the ability to undergo productive recombination giving rise to tk+ cells. Approximately 25% of them could do so. Southern blot analysis of these transfectants indicated that the majority contained a single copy, or at most, two copies of the plasmid integrated into the chromosome. Fluctuation analysis tests to determine the rate of spontaneous recombination (events per 10(6) cells per cell generation) in the various cell strains showed that the rates ranged from 0.15 to 4.1. The mean rate for the cell strains derived from the repair-deficient cell line was 3.6; for those derived from the cells with an intermediate rate, it was 0.8; and for those with a normal rate of excision repair, it was 0.9. Southern blot analysis of tk+ recombinants showed that in all cases, one of the Htk genes had become wild-type, i.e., XhoI-resistant. 90% of the recombinants retained the Htk gene duplication, consistent with non-reciprocal transfer of genetic information, i.e., gene conversion. The rest contained a single, wild-type Htk gene, consistent with a single reciprocal exchange within a chromatid or a single unequal exchange between sister chromatids. These cell strains will be useful for investigating the role of DNA damage and repair in homologous recombination.  相似文献   

18.
B. D. Bethke  J. Golin 《Genetics》1994,137(2):439-453
In Saccharomyces cerevisiae, spontaneous mitotic gene conversion at one site is statistically correlated with recombination at other loci. In general, coincident conversion frequencies are highest for tightly linked markers and decline as a function of intermarker distance. Paradoxically, a significant fraction of mitotic gene convertants exhibits concomitant nonreciprocal segregation for multiple and widely spaced markers. We have undertaken a detailed genetic analysis of this class of mitotic recombinants. Our results indicate that mitotic gene conversion in yeast is frequently associated with nonreciprocal segregation of markers centromere-distal to the selected site of conversion. In addition, distal markers are often found to be mosaic within the product colonies. These observations, and others described here, suggest that a percentage of gene conversion in vegetative yeast cells is coupled to a chromosome break and repair mechanism. This hypothesis was further tested using a strain trisomic for chromosome VII which was specially marked to detect homolog-dependent repair events. An association between mitotic gene conversion events and the production of broken chromosomes which are repaired by a homologous-pairing-copy mechanism was supported.  相似文献   

19.
In diploid eukaryotes, repair of double-stranded DNA breaks by homologous recombination often leads to loss of heterozygosity (LOH). Most previous studies of mitotic recombination in Saccharomyces cerevisiae have focused on a single chromosome or a single region of one chromosome at which LOH events can be selected. In this study, we used two techniques (single-nucleotide polymorphism microarrays and high-throughput DNA sequencing) to examine genome-wide LOH in a diploid yeast strain at a resolution averaging 1 kb. We examined both selected LOH events on chromosome V and unselected events throughout the genome in untreated cells and in cells treated with either γ-radiation or ultraviolet (UV) radiation. Our analysis shows the following: (1) spontaneous and damage-induced mitotic gene conversion tracts are more than three times larger than meiotic conversion tracts, and conversion tracts associated with crossovers are usually longer and more complex than those unassociated with crossovers; (2) most of the crossovers and conversions reflect the repair of two sister chromatids broken at the same position; and (3) both UV and γ-radiation efficiently induce LOH at doses of radiation that cause no significant loss of viability. Using high-throughput DNA sequencing, we also detected new mutations induced by γ-rays and UV. To our knowledge, our study represents the first high-resolution genome-wide analysis of DNA damage-induced LOH events performed in any eukaryote.  相似文献   

20.
The intracellular concentrations of the polypeptides encoded by the two enolase (ENO1 and ENO2) and three glyceraldehyde-3-phosphate dehydrogenase (TDH1, TDH2, and TDH3) genes were coordinately reduced more than 20-fold in a Saccharomyces cerevisiae strain carrying the gcr1-1 mutation. The steady-state concentration of glyceraldehyde-3-phosphate dehydrogenase mRNA was shown to be approximately 50-fold reduced in the mutant strain. Overexpression of enolase and glyceraldehyde-3-phosphate dehydrogenase in strains carrying multiple copies of either ENO1 or TDH3 was reduced more than 50-fold in strains carrying the gcr1-1 mutation. These results demonstrated that the GCR1 gene encodes a trans-acting factor which is required for efficient and coordinate expression of these glycolytic gene families. The GCR1 gene and the gcr1-1 mutant allele were cloned and sequenced. GCR1 encodes a predicted 844-amino-acid polypeptide; the gcr1-1 allele contains a 1-base-pair insertion mutation at codon 304. A null mutant carrying a deletion of 90% of the GCR1 coding sequence and a URA3 gene insertion was constructed by gene replacement. The phenotype of a strain carrying this null mutation was identical to that of the gcr1-1 mutant strain.  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