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 diploid human cells, the DNA precursor pool equilibration times for exogenous thymidine are about twice those for the thymidine analogue 5-bromodeoxyuridine (BUdR); in cells that were either transformed chemically or derived from malignant tumours, the pool equilibration times are the same for thymidine and 5-bromodeoxyuridine and are closer in value to the shorter (bromodeoxyuridine) times of the diploid cells. Thymidine, if present in the culture medium with BUdR, is incorporated into DNA preferentially in diploid cells (by 2 or 3 to 1). Discrimination against bromodeoxyuridine is evident within 2 h of incubation of the two precursors with diploid cells, but is not observed even after 24 h in any of the transformed cell lines tested. Experiments were performed to test the effect of inhibitors of the mammalian DNA polymerases alpha (N-ethylmaleimide) and beta (incubation of cells at 45 °C) upon the ability of cells to synthesise DNA and to incorporate thymidine preferentially when present with equimolar BUdR. In diploid cells, overall in vivo DNA synthesis is more sensitive to N-ethylmaleimide and more resistant to 45 °C treatment than is DNA synthesis in the transformed cell lines. N-Ethylmaleimide decreases the capacity of diploid cells to discriminate against BUdR, whereas heating increases it. Transformed cells treated with N-ethylmaleimide remain unable to discriminate against BUdR; some transformed lines, when heated at 45 °C, become less incapable of such discrimination.  相似文献   

2.
Candida species exhibit a variety of ploidy states and modes of sexual reproduction. Most species possess the requisite genes for sexual reproduction, recombination, and meiosis, yet only a few have been reported to undergo a complete sexual cycle including mating and sporulation. Candida albicans, the most studied Candida species and a prevalent human fungal pathogen, completes its sexual cycle via a parasexual process of concerted chromosome loss rather than a conventional meiosis. In this study, we examine ploidy changes in Candida tropicalis, a closely related species to C. albicans that was recently revealed to undergo sexual mating. C. tropicalis diploid cells mate to form tetraploid cells, and we show that these can be induced to undergo chromosome loss to regenerate diploid forms by growth on sorbose medium. The diploid products are themselves mating competent, thereby establishing a parasexual cycle in this species for the first time. Extended incubation (>120 generations) of C. tropicalis tetraploid cells under rich culture conditions also resulted in instability of the tetraploid form and a gradual reduction in ploidy back to the diploid state. The fitness levels of C. tropicalis diploid and tetraploid cells were compared, and diploid cells exhibited increased fitness relative to tetraploid cells in vitro, despite diploid and tetraploid cells having similar doubling times. Collectively, these experiments demonstrate distinct pathways by which a parasexual cycle can occur in C. tropicalis and indicate that nonmeiotic mechanisms drive ploidy changes in this prevalent human pathogen.  相似文献   

3.
We previously showed that fusion between hepatocytes lacking a crucial liver enzyme, fumarylacetoacetate hydrolase (FAH), and wild-type blood cells resulted in hepatocyte reprogramming. FAH expression was restored in hybrid hepatocytes and, upon in vivo expansion, ameliorated the effects of FAH deficiency. Here, we show that fusion-derived polyploid hepatocytes can undergo ploidy reductions to generate daughter cells with one-half chromosomal content. Fusion hybrids are, by definition, at least tetraploid. We demonstrate reduction to diploid chromosome content by multiple methods. First, cytogenetic analysis of fusion-derived hepatocytes reveals a population of diploid cells. Secondly, we demonstrate marker segregation using ß-galactosidase and the Y-chromosome. Approximately 2–5% of fusion-derived FAH-positive nodules were negative for one or more markers, as expected during ploidy reduction. Next, using a reporter system in which ß-galactosidase is expressed exclusively in fusion-derived hepatocytes, we identify a subpopulation of diploid cells expressing ß-galactosidase and FAH. Finally, we track marker segregation specifically in fusion-derived hepatocytes with diploid DNA content. Hemizygous markers were lost by ≥50% of Fah-positive cells. Since fusion-derived hepatocytes are minimally tetraploid, the existence of diploid hepatocytes demonstrates that fusion-derived cells can undergo ploidy reduction. Moreover, the high degree of marker loss in diploid daughter cells suggests that chromosomes/markers are lost in a non-random fashion. Thus, we propose that ploidy reductions lead to the generation of genetically diverse daughter cells with about 50% reduction in nuclear content. The generation of such daughter cells increases liver diversity, which may increase the likelihood of oncogenesis.  相似文献   

4.
Ko Harada  Edward G. Buss 《Genetics》1981,98(2):335-345
In the early stages of parthenogenetic development in turkey eggs, many blastoderms are mosaics of haploid, diploid and polyploid cells. The genome composition of these blastoderms can be identified by C-banding. They may be generally described as either A-Z/2A-ZZ/nA-nZ or A-W/2A-WW/nA-nW and are found in a nearly 1:1 ratio. The blastoderms showing the W body (W+) become lethal within two days of incubation. The haploid cell proportion decreases rapidly during the early stage of development, and, as haploid cells decrease, the proportion of polyploid cells appears to increase. At six days of incubation, various kinds of parthenogenetic development can be observed. Their genome compositions are either diploid (2A-ZZ) or mosaic (A-Z/2A-ZZ). These findings suggest that diploid parthenogenesis occurs by either suppression of meiosis II or chromosome doubling some time after the first cleavage division. The frequent occurrence of mosaic blastoderms indicates that the majority, if not all, of the parthenogenetic embryos initiate their development in haploid ova.  相似文献   

5.
Klar AJ 《Genetics》1980,94(3):597-605
Given a nutritional regime marked by a low nitrogen level and the absence of fermentable carbon sources, conventional a/α diploid cells of Saccharomyces cerevisiae exhibit a complex developmental sequence that includes a round of premeiotic DNA replication, commitment to meiosis and the elaboration of mature tetrads containing viable ascospores. Ordinarily, haploid cells and diploid cells of genotype a/a and α/α fail to display these reactions under comparable conditions. Here, we describe a simple technique for sporulation of α/α and a/a cells. Cells of genotype α/α are mated to haploid a cells carrying the kar1 (karyogamy defective) mutation to yield heterokaryons containing the corresponding diploid and haploid nuclei. The kar1 strains mate normally, but nuclei in the resultant zygotes do not fuse. When heterokaryotic cells are inoculated into sporulation media, they produce asci with six spores. Four spores carry genotypes derived from the diploid nucleus and the other two possess the markers originating from the haploid nucleus, i.e., the diploid nucleus divides meiotically while the haploid nucleus apparently divides mitotically. Similarly, the a/a genome is "helped" to sporulate as a consequence of mating with α kar1 strains. The results allow us to conclude that the mating-type functions essential for meiosis and sporulation are communicated and act through the cytoplasm and that sporulation can be dissociated from typical meiosis. This procedure will facilitate the genetic analysis of strains that are otherwise unable to sporulate.  相似文献   

6.
The number of chromosomes in cells of the root meristem of seedlings of wild and cultivated species of Cruciferae plants capable of hybridizing with rapeseed Brassica napus is studied. Only diploid metaphases are observed in seedlings of Brassica juncea, Diplotaxis tenuifolia, and Raphanus raphanistrum. In B. napus and B. cretica roughly 5% of the seedlings are mixoploid. Diploid cells are dominant in the mixoploids, though hypo-and hyperdiploids are also encountered. Nearly 20% of the seedlings of B. campestris and R. sativum are mixoploid, a significant fraction of which contains di-triploid chimeras. In B. nigra less than one-half of the seedlings are truly diploid, the majority of the plants being mixoploid. Seedlings containing preferentially tetraploid and triploid cells are dominant. The biological significance and possible causes of the newly discovered mixoploidy are discussed.  相似文献   

7.
By crossing of a heterothallic haploid yeast, Saccharomyces rouxii, we have succeeded in obtaining diploid hybrids. This paper shows one possible method of breeding heterothallic haploid yeasts for industrial application. S. rouxii is highly salt-tolerant and plays an important role in shoyu and miso fermentation. Therefore, genetic improvements of the properties are of commercial importance. Since newly isolated S. rouxii could neither conjugate nor sporulate on sporulation media commonly used, a suitable medium for conjugation and sporulation of S. rouxii was firstly investigated. A 5% NaCl Shoyu-koji extract agar was found to be most efficient. Next, we tried to get diploid strains by mass culture of two mating types on the conjugation medium, but several phenomena made this difficult: (i) zygotes quickly sporulated before budding; (ii) several zygotes showed terminal budding, but the buds could not grow into diploid cells, suggesting they would be heterocaryon; and (iii) a few zygotes lost their viability. After trying to isolate and cultivate a large number of zygotes in various combinations of crossing by micromanipulation, we fortunately recognized that large cells arose from some combinations. The analysis of ploidy suggested that the large cells would be diploid. Also, they showed sporulation of typical Saccharomyces, i.e., two to four spores in an unconjugated ascus. The diploid strains thus obtained were highly salt-tolerant and stable in liquid medium. Therefore, the procedure presented here would be effective for breeding salt-tolerant S. rouxii.  相似文献   

8.
Previous studies of adaptation to the glucose analog, 2-deoxyglucose, by Saccharomyces cerevisiae have utilized haploid cells. In this study, diploid cells were used in the hope of identifying the distinct genetic mechanisms used by diploid cells to acquire drug resistance. While haploid cells acquire resistance to 2-deoxyglucose primarily through recessive alleles in specific genes, diploid cells acquire resistance through dominant alleles, haploinsufficiency, gene duplication and aneuploidy. Dominant-acting, missense alleles in all three subunits of yeast AMP-activated protein kinase confer resistance to 2-deoxyglucose. Dominant-acting, nonsense alleles in the REG1 gene, which encodes a negative regulator of AMP-activated protein kinase, confer 2-deoxyglucose resistance through haploinsufficiency. Most of the resistant strains isolated in this study achieved resistance through aneuploidy. Cells with a monosomy of chromosome 4 are resistant to 2-deoxyglucose. While this genetic strategy comes with a severe fitness cost, it has the advantage of being readily reversible when 2-deoxyglucose selection is lifted. Increased expression of the two DOG phosphatase genes on chromosome 8 confers resistance and was achieved through trisomies and tetrasomies of that chromosome. Finally, resistance was also mediated by increased expression of hexose transporters, achieved by duplication of a 117 kb region of chromosome 4 that included the HXT3, HXT6 and HXT7 genes. The frequent use of aneuploidy as a genetic strategy for drug resistance in diploid yeast and human tumors may be in part due to its potential for reversibility when selection pressure shifts.  相似文献   

9.
Crandall M  Caulton JH 《Genetics》1979,93(4):903-916
Diploids of the yeast Hansenula wingei are nonagglutinative and do not form zygotes in mixed cultures with either sexually agglutinative haploid mating type. However, a low frequency of diploid x haploid cell fusions (about 10-3) is detectable by prototrophic selection. This frequency of rare diploid x haploid matings is not increased after the diploid culture is induced for sexual agglutination. Therefore, we conclude that genes that repress mating are different from those that repress sexual agglutination.——Six prototrophs isolated from one diploid x haploid cross had an average DNA value (µg DNA per 108 cells) of 6.19, compared to 2.53 and 4.35 for the haploid and diploid strains, respectively. Four prototrophs were clearly cell-fusion products because they contained genes from both the diploid and the haploid partners. However, genetic analysis of the prototrophs yielded results inconsistent with triploid meiosis; all six isolates yielded a 2:2 segregation for the mating-type alleles and linked genes.——Mitotic segregation of monosomic (2n-1) cells lacking one homolog of the chromosome carrying the mating-type locus is proposed to explain the rare production of sexually active cells in the diploid cultures. Fusion between such monosomic cells and normal haploids is thought to have produced 3n-1 cells, disomic for the chromosome carrying the mating-type locus. We conclude that in the diploid strain we studied, the physiological mechanisms repressing sexual agglutination and conjugation function efficiently, but events occuring during mitosis lead to a low frequency of genetically altered cells in the population.  相似文献   

10.
Adult chimeric epidermal structures were obtained following transplantation of haploid nuclei from haploid donor embryos of Drosophila into genetically marked diploid embryos. The haploid nuclei either remained haploid or became diploid. Where possible, physical measurements indicated that the haploid cells were smaller and produced smaller cuticular structures than did diploid cells. An increase in the number of pattern elements was observed in many patches which, by various criteria, were judged to be formed by haploid cells. The observation of altered pattern element spacing in haploid patches is in agreement with the conclusion, reached by L. I. Held (1979, Wilhelm Roux's Arch.187, 105–127) in triploid flies, that bristle spacing is a function of cell size.  相似文献   

11.
Survival curves were obtained for haploid and diploid yeasts, Saccharmyces cerevisiae, of wild-type strains and radiosensitive mutants exposed to γ-rays and α-particles. A correlation between the values of the relative biological effectiveness (RBE) of high-LET radiation and cell-repair capacity was found. The difference in radiosensitivities of the wild-type diploid strain and homozygous rad mutants incapable of recovery was significantly higher after low-LET radiation than after high-LET radiation. Possible reasons for the observed radiation responses to low- and high-LET exposure of yeast cells with various genotype are discussed.  相似文献   

12.
Campbell DA 《Genetics》1973,74(2):243-258
Mitotic recombination in Saccharomyces cerevisiae was examined by means of experiments in which one of the haploid parents was X-irradiated prior to zygote formation. By this method radiation-induced lesions are restricted to only one of the two non-sister chromatids that may be expected to undergo mitotic exchange in the diploid. The principal results of this work are: (1) X-irradiated haploid cells that are incapable of further vegetative growth (colony formation) are efficiently rescued into viable diploids by mating with unirradiated haploid cells. (2) X-rays delivered to only one of the two haploid parents are recombinogenic in the resultant diploid. The frequency of detected recombinational events increases as a probable linear function of the X-ray dose. (3) A majority of the induced recombinational events are nonreciprocal in nature (mitotic gene conversion). These results complement those obtained from X-irradiation of the vegetative diploid itself, where the induced genetic exchanges are principally reciprocal.  相似文献   

13.
Two protocols are presented that allow the selection of mutations mapping to micronuclear chromosome 5 in Tetrahymena thermophilia . One protocol involves crossing mutagenized diploid cells directly to a strain nullisomic for chromosome 5 and screening the monosomic progeny for a mutant phenotype. The second protocol first takes the mutagenized diploid cells through round I of genomic exclusion to create useful and reusable mutant heterokaryons, which are then assayed for the presence of mutations on chromosome 5 by crossing to the nullisomic 5 strain. Of 14 putative chromosome 5 mutations obtained by these two methods, seven are shown by genetic analysis to be recessive mutations on chromosome 5; one mapped elsewhere in the genome; six were infertile or failed to yield progeny in some of the diagnostic crosses and, thus, their genetic nature could not be determined with certainty.  相似文献   

14.
R DeMars 《Mutation research》1974,24(3):335-364
In vitro enumeration of diploid human cell variants that are resistant to purine analogues is a possible method of detecting mutagenesis. Their incidences can be increased by the known mutagens, X-rays and N-methyl-N′-nitro-N-nitrosoguanidine (MNNG). Usefulness of this method depends on the kinds of hereditary changes that confer analogue-resistance on somatic cells. If resistance usually results from changes in genetic material, in vitro studies could be useful indicators of mutagenic effects on somatic cells and germ cells in vivo. If epigenetic changes are primarily responsible for analogue-resistant variants, their enumeration might not provide information relevant to germinal mutations but would still be a useful way to detect induction of general kinds of stable phenotypic changes that could cause cancer. This article outlines hypothetical epigenetic and genetic causes of somatic cell variation and a prospective genetic analysis of human cell variants that are resistant to 8-azaguanine (AG) or 2,6-diaminopurine ( (DAP).Recent evidences and arguments favoring epigenetic origins of resistance to base-analogues are inconclusive. The often cited high rate of changes causing impermeability to BUdR in hamster cells is based on one improperly executed determination. Comparisons of rates of variation conferring BUdR-resistance on cultured haploid and diploid frog cells included diploid variants that did not behave as mutants and ignored major sources of error in estimating mutation rates. AG-resistance could result from recessive mutations in X-chromosomal genes but comparisons of rates of mutation in hamster cells of different ploidies did not provide information about the numbers of X-chromosomes in the variants. Reports that normal rodent HGPRT reappeared in hybrids of enzyme-deficient rodent cells and HGPRT-containing cells of other species or in the rodent cells alone in response to the conditions of cell hybridization did not include adequate controls for reversions in mutant genes of the rodent cells. Questions about the epigenetic and genetic origins of analogue-resistance are mostly unanswered. It remains possible that some kinds of abnormal epigenetic changes cause somatic disease. Specific methods for detecting their occurrence and responsiveness to environmental factors should be devised by focusing efforts on traits that are normally subject to epigenetic regulation. Derepression of genes on the inactive X-chromosome and of liver phenylalanine hydroxylase production are presented as possible examples of abnormal epigenetic changes that could be quantitatively studied by direct selection in vitro.  相似文献   

15.
Newly released zoospores fromEnteromorpha linza (L.) J. Ag. lack significant cellulose cell wall material and are suitable for treatment as protoplasts in a parasexual fusion process using high pH-Ca+ +, PEG and centrifugation. Treated zoospores settled on glass cover slips within 3 h and were examined microscopically at 1000 ×. Presumptive fusion products were identified by their larger size and presence of twin chloroplasts and eyespots. Unfused zoospores adjacent to fusion cells were killed by 2–3 min exposure to blue light (410–490 nm) from a high pressure mercury illuminator. Unexposed fusion cells developed into uniseriate germlings within 10 days at which stage they could be readily identified at 60 × with a dissecting microscope and isolated by micropipette. Ten-day germlings from both unfused zoospores and fusion cells were stained with the DNA-localizing fluorochrome hydroethidine and relative nuclear DNA content determined with epi-(incident) UV illumination. All germlings were found to be uninucleate. Germlings from unfused zoospores had haploid nuclei with 1N = 10 and 1C and 2C levels of DNA, while germlings from fusion cells had diploid nuclei with 2N = 20 and 2C and 4C levels of DNA. These result are interpreted as evidence of karyogamy following parasexual zoospore fusions. Isolated diploid germlings, cultured for 10 weeks were found to conserve their 2N chromosome complements and elevated levels of nuclear DNA. Although most diploid germlings were morphologically similar to haploid control plants, some exhibited ‘gigas’ characteristics, including larger cells, chloroplasts, and nuclei. These results are discussed in terms of unique phenotypes that result when nuclear and organellar genes are combined in different ways.  相似文献   

16.
Oshima T  Takano I 《Genetics》1980,96(4):841-857
Reverse and forward mutation, induced by photoaddition of 8-methoxypsoralen (8-MOP) and 3-carbethoxypsoralen (3-CPs) or ultraviolet light (UV), are reduced in three pso mutants of Saccharomyces cerevisiae. The pso1–1 strain exhibits a lower frequency of spontaneous reversion (antimutator) and is almost entirely unaffected by the three agents in both the haploid and diploid states. The pso2–1 strain demonstrates very reduced frequencies of 8-MOP and 3-CPs plus 365 nm radiation-induced mutations in happloid and diploid cells. UV-induced mutations are slightly reduced, whereas survival is almost normal. The pso3–1 strain is mutable by 8-MOP and 3-CPs photoaddition only in the low-dose range. After UV treatment, survival of pso3–1 is nearly normal, whereas the frequencies of induced mutants are diminished as compared to the normal PSO+. An analogue of adenine, 6-N-hydroxyaminopurine, is capable of inducing reversions in wild type, as well as in pso and rad6–1 mutant strains, indicating that this drug may act as a direct mutagen in yeast. The comparison of photoaddition of the bifunctional agent (8-MOP) to that of the monofunctional one (3-CPs) confirms that cross-links, as well as monoadditions, are mutagenic in S. cerevisiae. Repair, of the recombinational type, taking place in diploid cells or in haploid cells in G2 phase leads to higher survival, but appears to be error-free.  相似文献   

17.
Tetraploid (4N) cells are considered important in cancer because they can display increased tumorigenicity, resistance to conventional therapies, and are believed to be precursors to whole chromosome aneuploidy. It is therefore important to determine how tetraploid cancer cells arise, and how to target them. P53 is a tumor suppressor protein and key regulator of tetraploidy. As part of the “tetraploidy checkpoint”, p53 inhibits tetraploid cell proliferation by promoting a G1-arrest in incipient tetraploid cells (referred to as a tetraploid G1 arrest). Nutlin-3a is a preclinical drug that stabilizes p53 by blocking the interaction between p53 and MDM2. In the current study, Nutlin-3a promoted a p53-dependent tetraploid G1 arrest in two diploid clones of the HCT116 colon cancer cell line. Both clones underwent endoreduplication after Nutlin removal, giving rise to stable tetraploid clones that showed increased resistance to ionizing radiation (IR) and cisplatin (CP)-induced apoptosis compared to their diploid precursors. These findings demonstrate that transient p53 activation by Nutlin can promote tetraploid cell formation from diploid precursors, and the resulting tetraploid cells are therapy (IR/CP) resistant. Importantly, the tetraploid clones selected after Nutlin treatment expressed approximately twice as much P53 and MDM2 mRNA as diploid precursors, expressed approximately twice as many p53-MDM2 protein complexes (by co-immunoprecipitation), and were more susceptible to p53-dependent apoptosis and growth arrest induced by Nutlin. Based on these findings, we propose that p53 plays novel roles in both the formation and targeting of tetraploid cells. Specifically, we propose that 1) transient p53 activation can promote a tetraploid-G1 arrest and, as a result, may inadvertently promote formation of therapy-resistant tetraploid cells, and 2) therapy-resistant tetraploid cells, by virtue of having higher P53 gene copy number and expressing twice as many p53-MDM2 complexes, are more sensitive to apoptosis and/or growth arrest by anti-cancer MDM2 antagonists (e.g. Nutlin).  相似文献   

18.
The satellite DNAs of Drosophila virilis have been examined in diploid and polyploid tissues by isopycnic ultracentrifugation and thermal denaturation experiments. Previous work has established that the satellite DNAs are under replicated in the polytene chromosomes of the salivary glands of D. virilis. The results of the present experiments demonstrate that this underreplication also takes place in the ovaries which contain nurse cells and follicle cells. These tissues are polyploid but do not show polytene chromosomes.  相似文献   

19.
Organisms show great variation in ploidy level. For example, chromosome copy number varies among cells, individuals and species. One particularly widespread example of ploidy variation is found in haplodiploid taxa, wherein males are typically haploid and females are typically diploid. Despite the prevalence of haplodiploidy, the regulatory consequences of having separate haploid and diploid genomes are poorly understood. In particular, it remains unknown whether epigenetic mechanisms contribute to regulatory compensation for genome dosage. To gain greater insights into the importance of epigenetic information to ploidy compensation, we examined DNA methylation differences among diploid queen, diploid worker, haploid male and diploid male Solenopsis invicta fire ants. Surprisingly, we found that morphologically dissimilar diploid males, queens and workers were more similar to one another in terms of DNA methylation than were morphologically similar haploid and diploid males. Moreover, methylation level was positively associated with gene expression for genes that were differentially methylated in haploid and diploid castes. These data demonstrate that intragenic DNA methylation levels differ among individuals of distinct ploidy and are positively associated with levels of gene expression. Thus, these results suggest that epigenetic information may be linked to ploidy compensation in haplodiploid insects. Overall, this study suggests that epigenetic mechanisms may be important to maintaining appropriate patterns of gene regulation in biological systems that differ in genome copy number.  相似文献   

20.
The occurrence of polysomaty was noted in a diploid cytotype ofUrginea indica with B chromosomes. These accessories could be clearly distinguished from the other chromosomes in their stainability, very small size, morphology, variability and irregular behaviour. The variant nuclei were hyperdiploid, hypertriploid and hypertetraploid. B chromosomes were noted only in diploid cells and no accessories could be recorded in cells showing higher chromosome number. The amount of DNA content as indicated by the absorbance is rather low as compared to the diploid without accessories which may be due to the shorter size of the chromosomes. Under ultrastructural study subterminal centromeric region of B chromosome show low electron dense appearance.  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