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.
In homothallic cells of Saccharomyces cerevisiae, a or alpha mating type information at the mating type locus (MAT) is replaced by the transposition of the opposite mating type allele from HML alpha or HMRa. The rad52-1 mutation, which reduces mitotic and abolishes meiotic recombination, also affects homothallic switching (Malone and Esposito, Proc. Natl. Acad. Sci. U.S.A. 77:503-507, 1980). We have found that both HO rad52 MATa and HO rad52 MAT alpha cells die. This lethality is suppressed by mutations that substantially reduce but do not eliminate homothallic conversions. These mutations map at or near the MAT locus (MAT alpha inc, MATa-inc, MATa stk1) or are unlinked to MAT (HO-1 and swi1). These results suggest that the switching event itself is involved in the lethality. With the exception of swi1, HO rad52 strains carrying one of the above mutations cannot convert mating type at all. MAT alpha rad52 HO swi1 strains apparently can switch MAT alpha to MATa. However, when we analyzed these a maters, we found that few, if any, of them were bona fide MATa cells. These a-like cells were instead either deleted for part of chromosome III distal to and including MAT or had lost the entire third chromosome. Approximately 30% of the time, an a-like cell could be repaired to a normal MATa genotype if the cell was mated to a RAD52 MAT alpha-inc strain. The effects of rad52 were also studied in mata/MAT alpha-inc rad52/rad52 ho/HO diploids. When this diploid attempted to switch mata to MATa, an unstable broken chromosome was generated in nearly every cell. These studies suggest that homothallic switching involves the formation of a double-stranded deoxyribonucleic acid break or a structure which is labile in rad52 cells and results in a broken chromosome. We propose that the production of a double-stranded deoxyribonucleic acid break is the lethal event in rad52 HO cells.  相似文献   

2.
Meiotic recombination between a circular and a linear chromosome in Saccharomyces cerevisiae has been investigated. The circle was a haploid-viable derivative of chromosome III constructed by joining regions near the two chromosome ends via a recombinant DNA construction: (HMR/MAT-URA3-pBR322-MAT/HML) and was also deleted for MAL2 (which therefore uniquely marks a linear chromosome III). Recombination along chromosome III was measured for eight intervals spanning the entire length of the circular derivative. Only 25% of all tetrads from a ring/rod diploid contained four viable spores. These proved to be cases in which there was either no recombination along chromosome III or in which there were two-strand double crossovers or higher order crossovers that would not produce a dicentric chromosome.--At least half of the tetrads with three viable spores included one Ura+ Mal+ spore that was genetically highly unstable. The Ura+ Mal+ spore colonies gave rise to as many as seven genetically distinct, stable ("healed") derivatives, some of which had lost either URA3 or MAL2. Analysis of markers on chromosome III suggests that dicentric chromosomes frequently do not break during meiosis but are inherited intact into a haploid spore. In mitosis, however, the dicentric chromosome is frequently broken, giving rise to a variety of genetically distinct derivatives. We have also shown that dicentric ring chromosomes exhibit similar behavior: at least half the time they are not broken during meiosis but are broken and healed during mitosis.--The ring/rod diploid can also be used to determine the frequency of sister chromatid exchange (SCE) along an entire yeast ring chromosome. We estimate that an unequal number of SCE events occurs in approximately 15% of all cells undergoing meiosis. In contrast, the mitotic instability (and presumably SCE events) of a ring chromosome is low, occurring at a rate of about 1.2 X 10(-3) per cell division.  相似文献   

3.
Double-strand breaks can initiate meiotic recombination in S. cerevisiae   总被引:29,自引:0,他引:29  
A L Kolodkin  A J Klar  F W Stahl 《Cell》1986,46(5):733-740
We investigated the effects of double-strand breaks on meiotic recombination in yeast. A double-strand break was introduced at the MATa locus by sporulation of a MAT alpha inc/MATa diploid under inducing conditions for the HO-encoded endonuclease; 14% of the resulting tetrads had undergone 4 alpha:0a conversion. Conversion at MAT was associated with co-conversion of a closely linked marker and an increased recombination frequency for flanking markers. We also studied the sporulation products of a diploid heterozygous at the HIS4 locus for an insertion of a 100 bp fragment of MATa containing the HO endonuclease cut site. Under inducing conditions, a significant number of tetrads were formed that had undergone gene conversions in favor of the HIS4+ allele. Although double-strand breaks can initiate meiotic recombination in yeast, the data suggest that they do not normally do so.  相似文献   

4.
Assembly and disassembly of Rad51 and Rad52 complexes were monitored by immunofluorescence during homologous recombination initiated by an HO endonuclease-induced double-strand break (DSB) at the MAT locus. DSB-induced Rad51 and Rad52 foci colocalize with a TetR-GFP focus at tetO sequences adjacent to MAT. In strains in which HO cleaves three sites on chromosome III, we observe three distinct foci that colocalize with adjacent GFP chromosome marks. We compared the kinetics of focus formation with recombination intermediates and products when HO-cleaved MATalpha recombines with the donor, MATa. Rad51 assembly occurs 1 h after HO cleavage. Rad51 disassembly occurs at the same time that new DNA synthesis is initiated after single-stranded (ss) MAT DNA invades MATa. We present evidence for three distinct roles for Rad52 in recombination: a presynaptic role necessary for Rad51 assembly, a synaptic role with Rad51 filaments, and a postsynaptic role after Rad51 dissociates. Additional biochemical studies suggest the presence of an ssDNA complex containing both Rad51 and Rad52.  相似文献   

5.
Mating-type gene (MAT) switching in budding yeast exhibits donor preference. MATa preferentially recombines with HML near the left telomere of chromosome III, whereas MATalpha prefers HMR near the right telomere. Donor preference is controlled by the recombination enhancer (RE) located proximal to HML. To test if HML is constrained in pairing with MATalpha, we examined live-cell mobility of LacI-GFP-bound lactose operator (lacO) arrays inserted at different chromosomal sites. Without induction of recombination, lacO sequences adjacent to HML are strongly constrained in both MATalpha and RE-deleted MATa strains, compared with MATa. In contrast, chromosome movement at HMR or near a telomere of chromosome V is mating-type independent. HML is more constrained in MATa Deltare and less constrained in MATa RE+ compared with other sites. Although HML and MATa are not prealigned before inducing recombination, the three-dimensional configuration of MAT, HML, and HMR is mating-type dependent. These data suggest there is constitutive tethering of HML, which is relieved in MATa cells through the action of RE.  相似文献   

6.
Using the Saccharomyces cerevisiae MATa/MATalpha ORF deletion collection, homozygous deletion strains were identified that undergo mating with MATa or MATalpha haploids. Seven homozygous deletions were identified that confer enhanced mating. Three of these, lacking CTF8, CTF18, and DCC1, mate at a low frequency with either MATa or MATalpha haploids. The products of these genes form a complex involved in sister chromatid cohesion. Each of these strains also exhibits increased chromosome loss rates, and mating likely occurs due to loss of one copy of chromosome III, which bears the MAT locus. Three other homozygous diploid deletion strains, ylr193cDelta/ylr193cDelta, yor305wDelta/yor305wDelta, and ypr170cDelta/ypr170cDelta, mate at very low frequencies with haploids of either or both mating types. However, an ist3Delta/ist3Delta strain mates only with MATa haploids. It is shown that IST3, previously linked to splicing, is required for efficient processing of the MATa1 message, particularly the first intron. As a result, the ist3Delta/ist3Delta strain expresses unbalanced ratios of Matalpha to Mata proteins and therefore mates with MATa haploids. Accordingly, mating in this diploid can be repressed by introduction of a MATa1 cDNA. In summary, this study underscores and elaborates upon predicted pathways by which mutations restore mating function to yeast diploids and identifies new mutants warranting further study.  相似文献   

7.
The agalpha1 mutant MAT alpha cells specifically lack the cell surface alpha-type sexual agglutination substance, which is also called alpha-agglutinin. Because the mutant cells (MATalpha agalpha1) can not form aggregates with MATa cells, MATalpha agalpha1 cells are unable to mate with MATa cells when they are co-inoculated in a liquid medium, and the mating is attenuated on solid medium. The attenuated mating ability shown in the previous studies gave us a vague idea about a physiological function of the sexual agglutinability. In order to solve the question, mating behavior of MATalpha agalpha1 cells was investigated here under conditions where the contact between MATa and MAT alpha cells is assisted by physical methods. A synthetic mutation agalpha1::URA3 was constructed and used as well as agalpha1-1 for this study to ensure the genetic defect. When a mixture of MATa and MAT alpha cells was kept on filter membrane placed on relatively dry agar medium, even agalpha1::URA3 mutant cells mated as efficiently as the wild type (AGalpha1) cells did. On filter membrane placed on moist agar medium, agalpha1 mutants mated 10-fold less efficiently than wild type cells did. The mutant cells mated 10000-time less efficiently than the wild type cells in a pellet formed by brief low speed centrifugation. In contrast, the wild type MATalpha cells mated well under all conditions tested. Under the pellet condition, a mixture of MATa and MATalpha AG alpha1 cells formed an extended and cotton-like pellet while a mixture of MATa and MATalpha agalpha1 cells formed a compact and tight pellet. These results suggest that sexual cell agglutination contributes not only to cell contact between MATa and MAT alpha cells thereby stabilizing a-alpha cell pairs, but also to construction of a uniquely organized ultra structure favorable for zygote formation and subsequent growth of diploid cells. The mating specific extended pellet formation was observed also in 4 pairs of a and alpha strains in ascosporogenous yeast genera Hansenula and Pichia.  相似文献   

8.
9.
Arbel T  Shemesh R  Simchen G 《Genetics》1999,153(4):1583-1590
A single truncated chromosome fragment (TCF) in diploid cells undergoes frequent ectopic recombination during meiosis between markers located near the ends of the fragment. Tetrads produced by diploids with a single TCF show frequent loss of one of the two markers. This marker loss could result either from recombination of the TCF with one of the two copies of the chromosome from which it was derived or from ectopic recombination between the ends of the TCF. The former would result in shortening of a normal chromosome and lethality in one of the four spores. The high frequency of marker loss in tetrads with four viable spores supports recombination between the TCF ends as the main source of marker loss. Most of the spore colonies that display TCF marker loss contained a TCF with the same marker on both ends. Deletion of most of the pBR322 sequences distal to the marker at one of the subtelomeric regions of the TCF did not reduce the overall frequency of recombination between the ends, but affected the loss of one marker significantly more than the other. We suggest that the mechanism by which the duplication of one end marker and loss of the other occurs is based on association and recombination between the ends of the TCF.  相似文献   

10.
Coïc E  Richard GF  Haber JE 《Genetics》2006,173(3):1197-1206
Saccharomyces mating-type (MAT) switching occurs by gene conversion using one of two donors, HMLalpha and HMRa, located near the ends of the same chromosome. MATa cells preferentially choose HMLalpha, a decision that depends on the recombination enhancer (RE) that controls recombination along the left arm of chromosome III (III-L). When RE is inactive, the two chromosome arms constitute separate domains inaccessible to each other; thus HMRa, located on the same arm as MAT, becomes the default donor. Activation of RE increases HMLalpha usage, even when RE is moved 50 kb closer to the centromere. If MAT is inserted into the same domain as HML, RE plays little or no role in activating HML, thus ruling out any role for RE in remodeling the silent chromatin of HML in regulating donor preference. When the donors MAT and RE are moved to chromosome V, RE increases HML usage, but the inaccessibility of HML without RE apparently depends on other chromosome III-specific sequences. Similar conclusions were reached when RE was placed adjacent to leu2 or arg4 sequences engaged in spontaneous recombination. We propose that RE's targets are anchor sites that tether chromosome III-L in MATalpha cells thus reducing its mobility in the nucleus.  相似文献   

11.
Shiverer gene maps near the distal end of chromosome 18 in the house mouse   总被引:7,自引:0,他引:7  
Several mouse mutations cause unstable locomotion, tremor, seizures, and a reduced lifespan because of deficient myelin formation in the central nervous system. Mutant alleles at the shiverer (shi) locus are the only ones in this series with a selective molecular defect, namely, in myelin basic proteins (MBPs), which are virtually absent in shi homozygotes and 50% reduced in heterozygotes. In the present study, backcross and intercross matings indicate recombination of 21.2 +/- 3.3% between myelin deficient, shimld, and fused phalanges, syfp, a marker near the middle of chromosome 18. Recombination of shimld with twirler (Tw), a marker near the centromere, is 45.7 +/- 4.9%. Thus, the shi locus maps near the distal end of mouse chromosome 18 and is the first available marker for this region. Given the evidence of other workers that an MBP locus maps to the same mouse chromosome, and that part of this chromosome may be syntenic with an MBP-PEPA region on human chromosome 18, it is likely that shi is in or near an MBP gene.  相似文献   

12.
Homologous recombination, an essential process for preserving genomic integrity, uses intact homologous sequences to repair broken chromosomes. To explore the mechanism of homologous pairing in vivo, we tagged two homologous loci in diploid yeast Saccharomyces cerevisiae cells and investigated their dynamic organization in the absence and presence of DNA damage. When neither locus is damaged, homologous loci occupy largely separate regions, exploring only 2.7% of the nuclear volume. Following the induction of a double-strand break, homologous loci co-localize ten times more often. The mobility of the cut chromosome markedly increases, allowing it to explore a nuclear volume that is more than ten times larger. Interestingly, the mobility of uncut chromosomes also increases, allowing them to explore a four times larger volume. We propose a model for homology search in which increased chromosome mobility facilitates homologous pairing. Finally, we find that the increase in DNA dynamics is dependent on early steps of homologous recombination.  相似文献   

13.
Summary Polymorphism for six C-bands on chromosome 1R was used to study the frequency and distribution of recombination along the chromosome in a diploid rye (Secale cereale L.) and in a hexaploid triticale (X Triticosecale Wittmack) derived from it. In rye, the total recombination frequency in five segments of chromosome 1R was 93.7%. Recombination was concentrated in the distal regions of both chromosome arms and was infrequent in the proximal regions. In hexaploid triticale the total recombination frequency in the same chromosome was reduced to 51.7%. In both backgrounds the distal half of the long arm showed similar recombination frequencies, 51.4% and 45.7% for rye and triticale, respectively. The remaining about two-thirds of the chromosome length showed 42.3% recombination in rye but only 6% recombination in triticale. The results demonstrate that the genetic background in which mapping is performed not only affects the total amount of recombination, but also its distribution along the chromosome length.  相似文献   

14.
X. Wu  C. Wu    J. E. Haber 《Genetics》1997,147(2):399-407
Mating type (MAT) switching in Saccharomyces cerevisiae is initiated by a double-strand break (DSB) created at MAT by HO endonuclease. MATa cells activate the entire left arm of chromosome III; thus MATa preferentially recombines with the silent donor HML. In contrast, MATα cells inactivate the left arm, including HML, and thus preferentially recombine with HMR, 100 kb to the right of MAT. We present a novel competition assay, in which the DSB at MAT can be repaired either by MAT switching or by single-strand annealing (SSA) between two URA3 genes flanking MAT. With preferred donors, MATa or MATα switching occurs 65-70% of the time in competition with SSA. When HML is deleted, 40% of MATa cells recombine with the ``wrong' donor HMR; however, when HMR is deleted, only 18% of MATα cells recombine with HML. In interchromosomal switching, with donors on chromosome III and MAT on chromosome V, MATa retains its strong preference for HML and switching is efficient, when the chromosome III recombination enhancer is present. However, MATα donor preference is lost and interchromosomal switching is very inefficient. These experiments demonstrate the utility of using competition between two outcomes to measure the relative efficiency of recombination.  相似文献   

15.
Hiraoka M  Watanabe K  Umezu K  Maki H 《Genetics》2000,156(4):1531-1548
To obtain a broad perspective of the events leading to spontaneous loss of heterozygosity (LOH), we have characterized the genetic alterations that functionally inactivated the URA3 marker hemizygously or heterozygously situated either on chromosome III or chromosome V in diploid Saccharomyces cerevisiae cells. Analysis of chromosome structure in a large number of LOH clones by pulsed-field gel electrophoresis and PCR showed that chromosome loss, allelic recombination, and chromosome aberration were the major classes of genetic alterations leading to LOH. The frequencies of chromosome loss and chromosome aberration were significantly affected when the marker was located in different chromosomes, suggesting that chromosome-specific elements may affect the processes that led to these alterations. Aberrant-sized chromosomes were detected readily in approximately 8% of LOH events when the URA3 marker was placed in chromosome III. Molecular mechanisms underlying the chromosome aberrations were further investigated by studying the fate of two other genetic markers on chromosome III. Chromosome aberration caused by intrachromosomal rearrangements was predominantly due to a deletion between the MAT and HMR loci that occurred at a frequency of 3.1 x 10(-6). Another type of chromosome aberration, which occurred at a frequency slightly higher than that of the intrachromosomal deletion, appeared to be caused by interchromosomal rearrangement, including unequal crossing over between homologous chromatids and translocation with another chromosome.  相似文献   

16.
17.
Transformation experiments with Bacillus subtilis strains carrying trpE26 (the marker responsible for the detection of merodiploid clones after transformation or transduction) have established the precise position of this marker on the "aromatic region" of the chromosome, at the distal end of the anthranilate synthetase locus. Integration efficiency of the mutant allele (trpE26) seems to be very low. Co-transfer of markers situated on either side of it is almost nil when both donor and recipient carry this mutation. The "exclusion" of trpE26 does not, however, affect recombination frequencies for nearby markers. To explain these facts we considered the hypothesis of a preferential breakage of the deoxyribonucleic acid (DNA) at the trpE26 site or that of an insertion mutation. These studies have also demonstrated the establishment of physical linkage of a marker from the exogenote (hisH2) to a resident marker (tyrA1) in stable and unstable merodiploid clones, thus confirming integration of the donor DNA segment into a genetic structure of the recipient. Furthermore, duplication was shown in merodiploid clones (through reversion and transformation) for a locus of the recipient (tyrA) which was not involved in the initial transformation. This suggests that the diploid condition in this region extends beyond the transformed area. Interpretation of the genetic constitution of these partial diploids calls for postulation of the existence of long duplications, a second (incomplete) chromosome, or an episome-like element.  相似文献   

18.
Mating procedures for detection of mobilization of the Proteus mirabilis chromosome were re-investigated. The chromosome was mobilized by plasmid D, the previously used hybrid between plasmids P-lac and R1drd19. About a 40-fold increase in recombinant recovery correlated with the absence of swarming during mating and a lower temperature of incubation. The modification introduced was that conjugation was allowed to proceed on a non-selective supplemented minimal medium at 30 degrees C before washing and plating on selective media. Final incubation was also at 30 degrees C. This technique enabled eight additional chromosomal markers to be mapped. Polarized transfer of the chromosome was shown by gradient of transmission experiments using a previously described marker as reference, by linkage analysis with reference to proximal and distal markers and (less successfully) by interrupted mating on solid medium. Markers of plasmid D transferred at high frequency to all recombinants. The plasmid was stable in recombinants and could transfer itself and chromosomal markers of the new hosts in further matings. Resulting recombination of markers occurred at usual frequencies. The marker order, his-1, ser-2, ura-2, pyrB1, trp-3, cysA1, ade-2, ilv-2, cysG1, gly-1, cysC1, argA2, metF2, nalA1, thr-1, leuB2, did not resemble the order of these markers in Escherichia coli.  相似文献   

19.
Ajima J  Umezu K  Maki H 《Mutation research》2002,504(1-2):157-172
The SGS1 gene of Saccharomyces cerevisiae is a member of the RecQ helicase family, which includes the human BLM, WRN and RECQL4 genes responsible for Bloom and Werner's syndrome and Rothmund-Thomson syndrome, respectively. Cells defective in any of these genes exhibit a higher incidence of genome instability. We previously demonstrated that various genetic alterations were detectable as events leading to loss of heterozygosity (LOH) in S. cerevisiae diploid cells, utilizing a hemizygous URA3 marker placed at the center of the right arm of chromosome III. Analyses of chromosome structure in LOH clones by pulse field gel electrophoresis (PFGE) and PCR, coupled with a genetic method, allow identification of genetic alterations leading to the LOH. Such alterations include chromosome loss, chromosomal rearrangements at various locations and intragenic mutation. In this work, we have investigated the LOH events occurring in cells lacking the SGS1 gene. The frequencies of all types of LOH events, excluding intragenic mutation, were increased in sgs1 null mutants as compared to the wild-type cells. Loss of chromosome III and chromosomal rearrangements were increased 13- and 17-fold, respectively. Further classification of the chromosomal rearrangements confirmed that two kinds of events were especially increased in the sgs1 mutants: (1) ectopic recombination between chromosomes, that is, unequal crossing over and translocation (46-fold); and (2) allelic crossing over associated with chromosome loss (40-fold). These findings raise the possibility that the Sgs1 protein is involved in the processing of recombination intermediates as well as in the prevention of recombination repair during chromosome DNA replication. On the other hand, intrachromosomal deletions between MAT and HMR were increased only slightly (2.9-fold) in the sgs1 mutants. These results clearly indicate that defects in the SGS1 gene function lead to an elevated incidence of LOH in multiple ways, including chromosome loss and interchromosomal rearrangements, but not intrachromosomal deletion.  相似文献   

20.
G Bosco  J E Haber 《Genetics》1998,150(3):1037-1047
In yeast, broken chromosomes can be repaired by recombination, resulting in nonreciprocal translocations. In haploid cells suffering an HO endonuclease-induced, double-strand break (DSB), nearly 2% of the broken chromosome ends recombined with a sequence near the opposite chromosome end, which shares only 72 bp of homology with the cut sequence. This produced a repaired chromosome with the same 20-kb sequence at each end. Diploid strains were constructed in which the broken chromosome shared homology with the unbroken chromosome only on the centromere-proximal side of the DSB. More than half of these cells repaired the DSB by copying sequences distal to the break from the unbroken template chromosome. All these events were RAD52 dependent. Pedigree analysis established that DSBs occurring in G1 were repaired by a replicative mechanism, producing two identical daughter cells. We discuss the implications of these data in understanding telomerase-independent replication of telomeres, gene amplification, and the evolution of chromosomal ends.  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