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 9 毫秒
1.
L. W. Yuan  R. L. Keil 《Genetics》1990,124(2):263-273
Many genetic studies have shown that the frequency of homologous recombination depends largely on the distance in which recombination can occur. We have studied the effect of varying the length of duplicated sequences on the frequency of mitotic intrachromosomal recombination in Saccharomyces cerevisiae. We find that the frequency of recombination resulting in the loss of one of the repeats and the intervening sequences reaches a plateau when the repeats are short. In addition, the frequency of recombination to correct a point mutation contained in one of these repeats is not proportional to the size of the duplication but rather depends dramatically on the location of the mutation within the repeated sequences. However, the frequency of mitotic interchromosomal reciprocal recombination is dependent on the distance separating the markers. The difference in the response of intrachromosomal and interchromosomal mitotic recombination to increasing lengths of homology may indicate there are different rate-limiting steps for recombination in these two cases. These findings have important implications for the maintenance and evolution of duplicated sequences.  相似文献   

2.
E. J. Louis  J. E. Haber 《Genetics》1990,124(3):547-559
Y's are a dispersed family of repeats that vary in copy number, location and restriction fragment lengths between strains but exhibit within-strain homogeneity. We have studied mitotic recombination between members of the subtelomeric Y' repeated sequence family of Saccharomyces cerevisiae. Individual copies of Y's were marked with SUP11 and URA3 which allowed for the selection of duplications and losses of the marked Y's. Duplications occurred by ectopic recombinational interactions between Y's at different chromosome ends as well as by unequal sister chromatid exchange. Several of the ectopic duplications resulted in an originally Y'-less chromosome end acquiring a marked Y'. Among losses, most resulted from ectopic exchange or conversion in which only the marker sequence was lost. In some losses, the chromosome end became Y'-less. Although the two subsets of Y's, Y'-longs (6.7 kb) and Y'-shorts (5.2 kb), share extensive sequence homology, a marked Y' recombines highly preferentially within its own subset. These mitotic interactions can in part explain the maintenance of Y's and their subsets, the homogeneity among Y's within a strain, as well as diversity between strains.  相似文献   

3.
M. Lichten  J. E. Haber 《Genetics》1989,123(2):261-268
We have examined the role that genomic location plays in mitotic intragenic recombination. Mutant alleles of the LEU2 gene were inserted at five locations in the yeast genome. Diploid and haploid strains containing various combinations of these inserts were used to examine both allelic recombination (between sequences at the same position on parental homologs) and ectopic recombination (between sequences at nonallelic locations). Chromosomal location had little effect on mitotic allelic recombination. The rate of recombination to LEU2 at five different loci varied less than threefold. This finding contrasts with previous observations of strong position effects in meiosis; frequencies of meiotic recombination at the same five loci differ by about a factor of forty. Mitotic recombination between dispersed copies of leu2 displayed strong position effects. Copies of leu2 located approximately 20 kb apart on the same chromosome recombined at rates 6-13-fold higher than those observed for allelic copies of leu2. leu2 sequences located on nonhomologous chromosomes or at distant loci on the same chromosome recombined at rates similar to those observed for allelic copies. We suggest that, during mitosis, parental homologs interact with each other no more frequently than do nonhomologous chromosomes.  相似文献   

4.
E. Alani  RAG. Reenan    R. D. Kolodner 《Genetics》1994,137(1):19-39
The yeast Saccharomyces cerevisiae encodes a set of genes that show strong amino acid sequence similarity to MutS and MutL, proteins required for mismatch repair in Escherichia coli. We examined the role of MSH2 and PMS1, yeast homologs of mutS and mutL, respectively, in the repair of base pair mismatches formed during meiotic recombination. By using specifically marked HIS4 and ARG4 alleles, we showed that msh2 mutants displayed a severe defect in the repair of all base pair mismatches as well as 1-, 2- and 4-bp insertion/deletion mispairs. The msh2 and pms1 phenotypes were indistinguishable, suggesting that the wild-type gene products act in the same repair pathway. A comparison of gene conversion events in wild-type and msh2 mutants indicated that mismatch repair plays an important role in genetic recombination. (1) Tetrad analysis at five different loci revealed that, in msh2 mutants, the majority of aberrant segregants displayed a sectored phenotype, consistent with a failure to repair mismatches created during heteroduplex formation. In wild type, base pair mismatches were almost exclusively repaired toward conversion rather than restoration. (2) In msh2 strains 10-19% of the aberrant tetrads were Ab4:4. (3) Polarity gradients at HIS4 and ARG4 were nearly abolished in msh2 mutants. The frequency of gene conversion at the 3' end of these genes was increased and was nearly the frequency observed at the 5' end. (4) Co-conversion studies were consistent with mismatch repair acting to regulate heteroduplex DNA tract length. We favor a model proposing that recombination events occur through the formation and resolution of heteroduplex intermediates and that mismatch repair proteins specifically interact with recombination enzymes to regulate the length of symmetric heteroduplex DNA.  相似文献   

5.
N. Rudin  E. Sugarman    J. E. Haber 《Genetics》1989,122(3):519-534
We have investigated HO endonuclease-induced double-strand break (DSB) recombination and repair in a LACZ duplication plasmid in yeast. A 117-bp MATa fragment, embedded in one copy of LACZ, served as a site for initiation of a DSB when HO endonuclease was expressed. The DSB could be repaired using wild-type sequences located on a second, promoterless, copy of LACZ on the same plasmid. In contrast to normal mating-type switching, crossing-over associated with gene conversion occurred at least 50% of the time. The proportion of conversion events accompanied by exchange was greater when the two copies of LACZ were in direct orientation (80%), than when inverted (50%). In addition, the fraction of plasmids lost was significantly greater in the inverted orientation. The kinetics of appearance of intermediates and final products were also monitored. The repair of the DSB is slow, requiring at least an hour from the detection of the HO-cut fragments to completion of repair. Surprisingly, the appearance of the two reciprocal products of crossing over did not occur with the same kinetics. For example, when the two LACZ sequences were in the direct orientation, the HO-induced formation of a large circular deletion product was not accompanied by the appearance of a small circular reciprocal product. We suggest that these differences may reflect two kinetically separable processes, one involving only one cut end and the other resulting from the concerted participation of both ends of the DSB.  相似文献   

6.
H. L. Klein 《Genetics》1997,147(4):1533-1543
Most mitotic recombination and repair genes of Saccharomyces cerevisiae show no specificity of action for the genome ploidy. We describe here a novel repair and recombination gene that is specific for recombination and repair between homologous chromosomes. The RDH54 gene is homologous to the RAD54 gene, but rdh54 mutants do not show sensitivity to methyl methanesulfonate at concentrations that sensitize a rad54 mutant. However, the rdh54 null mutation enhances the methyl methanesulfonate sensitivity of a rad54 mutant and single rdh54 mutants are sensitive to prolonged exposure at high concentrations of methyl methanesulfonate. The RDH54 gene is required for recombination, but only in a diploid. We present evidence showing that the RDH54 gene is required for interhomologue gene conversion but not intrachromosomal gene conversion. The rdh54 mutation confers diploid-specific lethalities and reduced growth in various mutant backgrounds. These phenotypes are due to attempted recombination. The RDH54 gene is also required for meiosis as homozygous mutant diploids show very poor sporulation and reduced spore viability. The role of the RDH54 gene in mitotic repair and in meiosis and the pathway in which it acts are discussed.  相似文献   

7.
M. Saparbaev  L. Prakash    S. Prakash 《Genetics》1996,142(3):727-736
The RAD1 and RAD10 genes of Saccharomyces cerevisiae are required for nucleotide excision repair and they also act in mitotic recombination. The Rad1-Rad10 complex has a single-stranded DNA endonuclease activity. Here, we show that the mismatch repair genes MSH2 and MSH3 function in mitotic recombination. For both his3 and his4 duplications, and for homologous integration of a linear DNA fragment into the genome, the msh3Δ mutation has an effect on recombination similar to that of the rad1Δ and rad10Δ mutations. The msh2Δ mutation also reduces the rate of recombination of the his3 duplication and lowers the incidence of homologous integration of a linear DNA fragment. Epistasis analyses indicate that MSH2 and MSH3 function in the RAD1-RAD10 recombination pathway, and studies presented here suggest an involvement of the RAD1-RAD10 pathway in reciprocal recombination. The possible roles of Msh2, Msh3, Rad1, and Rad10 proteins in genetic recombination are discussed. Coupling of mismatch binding proteins with the recombinational machinery could be important for ensuring genetic fidelity in the recombination process.  相似文献   

8.
F. Spencer  S. L. Gerring  C. Connelly    P. Hieter 《Genetics》1990,124(2):237-249
We have isolated 136 independent mutations in haploid yeast strains that exhibit decreased chromosome transmission fidelity in mitosis. Eighty-five percent of the mutations are recessive and 15% are partially dominant. Complementation analysis between MATa and MAT alpha isolates identifies 11 chromosome transmission fidelity (CTF) complementation groups, the largest of which is identical to CHL1. For 49 independent mutations, no corresponding allele has been recovered in the opposite mating type. The initial screen monitored the stability of a centromere-linked color marker on a nonessential yeast chromosome fragment; the mitotic inheritance of natural yeast chromosome III is also affected by the ctf mutations. Of the 136 isolates identified, seven were inviable at 37 degrees and five were inviable at 11 degrees. In all cases tested, these temperature conditional lethalities cosegregated with the chromosome instability phenotype. Five additional complementation groups (ctf12 through ctf16) have been defined by complementation analysis of the mutations causing inviability at 37 degrees. Twenty-three of the 136 isolates exhibited growth defects at concentrations of benomyl permissive for the parent strain, and nine appeared to be tolerant of inhibitory levels of benomyl. All of the mutant strains showed normal sensitivity to ultraviolet and gamma-irradiation. Further characterization of these mutant strains will describe the functions of gene products crucial to the successful execution of processes required for aspects of the chromosome cycle that are important for chromosome transmission fidelity in mitosis.  相似文献   

9.
We tested predictions of the double-strand break repair (DSBR) model for meiotic recombination by examining the segregation patterns of small palindromic insertions, which frequently escape mismatch repair when in heteroduplex DNA. The palindromes flanked a well characterized DSB site at the ARG4 locus. The ``canonical'''' DSBR model, in which only 5'' ends are degraded and resolution of the four-stranded intermediate is by Holliday junction resolvase, predicts that hDNA will frequently occur on both participating chromatids in a single event. Tetrads reflecting this configuration of hDNA were rare. In addition, a class of tetrads not predicted by the canonical DSBR model was identified. This class represented events that produced hDNA in a ``trans'''' configuration, on opposite strands of the same duplex on the two sides of the DSB site. Whereas most classes of convertant tetrads had typical frequencies of associated crossovers, tetrads with trans hDNA were parental for flanking markers. Modified versions of the DSBR model, including one that uses a topoisomerase to resolve the canonical DSBR intermediate, are supported by these data.  相似文献   

10.
The genes in the RAD52 epistasis group of Saccharomyces cerevisiae are necessary for most mitotic and meiotic recombination events. Using an intrachromosomal inverted-repeat assay, we previously demonstrated that mitotic recombination of this substrate is dependent upon the RAD52 gene. In the present study the requirement for other genes in this epistasis group for recombination of inverted repeats has been analyzed, and double and triple mutant strains were examined for their epistatic relationships. The majority of recombination events are mediated by a RAD51-dependent pathway, where the RAD54, RAD55 and RAD57 genes function downstream of RAD51. Cells mutated in RAD55 or RAD57 as well as double mutants are cold-sensitive for inverted-repeat recombination, whereas a rad51 rad55 rad57 triple mutant is not. The RAD1 gene is not required for inverted-repeat recombination but is able to process spontaneous DNA lesions to produce recombinant products in the absence of RAD51. Furthermore, there is still considerably more recombination in rad1 rad51 mutants than in rad52 mutants, indicating the presence of another, as yet unidentified, recombination pathway.  相似文献   

11.
Recombination is an essential part of meiosis; in almost all organisms, including Saccharomyces cerevisiae, proper chromosome segregation and the viability of meiotic products is dependent upon normal levels of recombination. In this article we examine the kinetics of the meiotic divisions in four mutants defective in the initiation of recombination. We find that mutations in any of three Early Exchange genes (REC104, REC114 or REC102) confer a phenotype in which the reductional division occurs earlier than in an isogenic wild-type diploid. We also present data confirming previous reports that strains with a mutation in the Early Exchange gene MEI4 undergo the first division at about the same time as wild-type cells. The rec104 mutation is epistatic to the mei4 mutation for the timing of the first division. These observations suggest a possible relationship between the initiation of recombination and the timing of the reductional division. These data also allow these four Early Exchange genes examined to be distinguished in terms of their role in coordinating recombination with the reductional division.  相似文献   

12.
A. M. Bailis  R. Rothstein 《Genetics》1990,126(3):535-547
Null mutations in three recombination and DNA repair genes were studied to determine their effects on mitotic recombination between the duplicate AdoMet (S-adenosylmethionine) synthetase genes (SAM1 and SAM2) in Saccharomyces cerevisiae. SAM1 and SAM2, located on chromosomes XII and IV, respectively, encode functionally equivalent although differentially regulated AdoMet synthetases. These similar but not identical (homeologous) genes are 83% homologous at the nucleotide level and this identity is limited solely to the coding regions of the genes. Single frameshift mutations were introduced into the 5' end of SAM1 and the 3' end of SAM2 by restriction site ablation. The sequences surrounding these mutations differ significantly in their degree of homology to the corresponding area of the other gene. Mitotic ectopic recombination between the mutant sam genes occurs at a rate of 8.4 x 10(-9) in a wild-type genetic background. Gene conversion of the marker within the region of greater sequence homology occurs 20-fold more frequently than conversion of the marker within the region of relative sequence diversity. The relative orientation of the two genes prevents the recovery of translocations. Mitotic recombination between the sam genes is completely dependent on the DNA repair and recombination gene RAD52. A mutation in PMS1, a mismatch repair gene, causes a 4.5-fold increase in the rate of ectopic recombination. RAD1, an excision repair gene, is required to observe this increased rate of ectopic conversion. In addition, RAD1 is involved in modulating the pattern of coconversion during recombination between the homeologous sam genes. These results suggest that interactions between mismatch repair, excision repair and recombinational repair functions are involved in determining the ectopic gene conversion frequency between the sam genes.  相似文献   

13.
E. L. Ivanov  V. G. Korolev    F. Fabre 《Genetics》1992,132(3):651-664
The XRS2 gene of Saccharomyces cerevisiae has been previously identified as a DNA repair gene. In this communication, we show that XRS2 also encodes an essential meiotic function. Spore inviability of xrs2 strains is rescued by a spo13 mutation, but meiotic recombination (both gene conversion and crossing over) is highly depressed in spo13 xrs2 diploids. The xrs2 mutation suppresses spore inviability of a spo13 rad52 strain suggesting that XRS2 acts prior to RAD52 in the meiotic recombination pathway. In agreement with the genetic data, meiosis-specific double-strand breaks at the ARG4 meiotic recombination hotspot are not detected in xrs2 strains. Despite its effects on meiotic recombination, the xrs2 mutation does not prevent mitotic recombination events, including homologous integration of linear DNA, mating-type switching and radiation-induced gene conversion. Moreover, xrs2 strains display a mitotic hyper-rec phenotype. Haploid xrs2 cells fail to carry out G2-repair of gamma-induced lesions, whereas xrs2 diploids are able to perform some diploid-specific repair of these lesions. Meiotic and mitotic phenotypes of xrs2 cells are very similar to those of rad50 cells suggesting that XRS2 is involved in homologous recombination in a way analogous to that of RAD50.  相似文献   

14.
In the yeast, Saccharomyces cerevisiae, several genes appear to act early in meiotic recombination. HOP1 and RED1 have been classified as such early genes. The data in this paper demonstrate that neither a red1 nor a hop1 mutation can rescue the inviable spores produced by a rad52 spo13 strain; this phenotype helps to distinguish these two genes from other early meiotic recombination genes such as SPO11, REC104, or MEI4. In contrast, either a red1 or a hop1 mutation can rescue a rad50S spo13 strain; this phenotype is similar to that conferred by mutations in the other early recombination genes (e.g., REC104). These two different results can be explained because the data presented here indicate that a rad50S mutation does not diminish meiotic intrachromosomal recombination, similar to the mutant phenotypes conferred by red1 or hop1. Of course, RED1 and HOP1 do act in the normal meiotic interchromosomal recombination pathway; they reduce interchromosomal recombination to ~10% of normal levels. We demonstrate that a mutation in a gene (REC104) required for initiation of exchange is completely epistatic to a mutation in RED1. Finally, mutations in either HOP1 or RED1 reduce the number of double-strand breaks observed at the HIS2 meiotic recombination hotspot.  相似文献   

15.
16.
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   

17.
Using plasmids capable of undergoing intramolecular recombination, we have compared the rates and the molecular outcomes of recombination events in a wild-type and a rad52 strain of Saccharomyces cerevisiae. The plasmids contain his3 heteroalleles oriented in either an inverted or a direct repeat. Inverted repeat plasmids recombine approximately 20-fold less frequently in the mutant than in the wild-type strain. Most events from both cell types have continuous coconversion tracts extending along one of the homologous segments. Reciprocal exchange occurs in fewer than 30% of events. Direct repeat plasmids recombine at rates comparable to those of inverted repeat plasmids in wild-type cells. Direct repeat conversion tracts are similar to inverted repeat conversion tracts in their continuity and length. Inverted and direct repeat plasmid recombination differ in two respects. First, rad52 does not affect the rate of direct repeat recombination as drastically as the rate of inverted repeat recombination. Second, direct repeat plasmids undergo crossing over more frequently than inverted repeat plasmids. In addition, crossovers constitute a larger fraction of mutant than wild-type direct repeat events. Many crossover events from both cell types are unusual in that the crossover HIS3 allele is within a plasmid containing the parental his3 heteroalleles.  相似文献   

18.
L. C. Kadyk  L. H. Hartwell 《Genetics》1993,133(3):469-487
Homolog recombination and unequal sister chromatid recombination were monitored in rad1-1/rad1-1 diploid yeast cells deficient for excision repair, and in control cells, RAD1/rad1-1, after exposure to UV irradiation. In a rad1-1/rad1-1 diploid, UV irradiation stimulated much more sister chromatid recombination relative to homolog recombination when cells were irradiated in the G(1) or the G(2) phases of the cell cycle than was observed in RAD1/rad1-1 cells. Since sister chromatids are not present during G(1), this result suggested that unexcised lesions can stimulate sister chromatid recombination events during or subsequent to DNA replication. The results of mating rescue experiments suggest that unexcised UV dimers do not stimulate sister chromatid recombination during the G(2) phase, but only when they are present during DNA replication. We propose that there are two types of sister chromatid recombination in yeast. In the first type, unexcised UV dimers and other bulky lesions induce sister chromatid recombination during DNA replication as a mechanism to bypass lesions obstructing the passage of DNA polymerase, and this type is analogous to the type of sister chromatid exchange commonly observed cytologically in mammalian cells. In the second type, strand scissions created by X-irradiation or the excision of damaged bases create recombinogenic sites that result in sister chromatid recombination directly in G(2). Further support for the existence of two types of sister chromatid recombination is the fact that events induced in rad1-1/rad1-1 were due almost entirely to gene conversion, whereas those in RAD1/rad1-1 cells were due to a mixture of gene conversion and reciprocal recombination.  相似文献   

19.
M. Ajimura  S. H. Leem    H. Ogawa 《Genetics》1993,133(1):51-66
Mutants defective in meiotic recombination were isolated from a disomic haploid strain of Saccharomyces cerevisiae by examining recombination within the leu2 and his4 heteroalleles located on chromosome III. The mutants were classified into two new complementation groups (MRE2 and MRE11) and eight previously identified groups, which include SPO11, HOP1, REC114, MRE4/MEK1 and genes in the RAD52 epistasis group. All of the mutants, in which the mutations in the new complementation groups are homozygous and diploid, can undergo premeiotic DNA synthesis and produce spores. The spores are, however, not viable. The mre2 and mre11 mutants produce viable spores in a spo13 background, in which meiosis I is bypassed, suggesting that these mutants are blocked at an early step in meiotic recombination. The mre2 mutant does not exhibit any unusual phenotype during mitosis and it is, thus, considered to have a mutation in a meiosis-specific gene. By contrast, the mre11 mutant is sensitive to damage to DNA by methyl methanesulfonate and exhibits a hyperrecombination phenotype in mitosis. Among six alleles of HOP1 that were isolated, an unusual pattern of intragenic complementation was observed.  相似文献   

20.
Natural strains of Saccharomyces cerevisiae are prototrophic homothallic yeasts that sporulate poorly, are often heterozygous, and may be aneuploid. This genomic constitution may confer selective advantages in some environments. Different mechanisms of recombination, such as meiosis or mitotic rearrangement of chromosomes, have been proposed for wine strains. We studied the stability of the URA3 locus of a URA3/ura3 wine yeast in consecutive grape must fermentations. ura3/ura3 homozygotes were detected at a rate of 1 × 10−5 to 3 × 10−5 per generation, and mitotic rearrangements for chromosomes VIII and XII appeared after 30 mitotic divisions. We used the karyotype as a meiotic marker and determined that sporulation was not involved in this process. Thus, we propose a hypothesis for the genome changes in wine yeasts during vinification. This putative mechanism involves mitotic recombination between homologous sequences and does not necessarily imply 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