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.
2.
Aneuploid cells are characterized by incomplete chromosome sets. The resulting imbalance in gene dosage has phenotypic consequences that are specific to each karyotype. Even in the case of Down syndrome, the most viable and studied form of human aneuploidy, the mechanisms underlying the connected phenotypes remain mostly unclear. Because of their tolerance to aneuploidy, plants provide a powerful system for a genome-wide investigation of aneuploid syndromes, an approach that is not feasible in animal systems. Indeed, in many plant species, populations of aneuploid individuals can be easily obtained from triploid individuals. We phenotyped a population of Arabidopsis thaliana aneuploid individuals containing 25 different karyotypes. Even in this highly heterogeneous population, we demonstrate that certain traits are strongly associated with the dosage of specific chromosome types and that chromosomal effects can be additive. Further, we identified subtle developmental phenotypes expressed in the diploid progeny of aneuploid parent(s) but not in euploid controls from diploid lineages. These results indicate long-term phenotypic consequences of aneuploidy that can persist after chromosomal balance has been restored. We verified the diploid nature of these individuals by whole-genome sequencing and discuss the possibility that trans-generational phenotypic effects stem from epigenetic modifications passed from aneuploid parents to their diploid progeny.THE genome of aneuploid individuals contains incomplete chromosome sets. The balance between chromosome types, and the genes they encode, is compromised, resulting in altered expression of many genes, including genes with dosage-sensitive effects on phenotypes. In humans, only a few types of aneuploid karyotypes are viable (Hassold and Hunt 2001), highlighting the deleterious effect of chromosome imbalance. The most commonly known viable form of aneuploidy in humans is Down syndrome, which results from a trisomy of chromosome 21 in an otherwise diploid background. Down syndrome patients exhibit many specific phenotypes, sometimes visible only in a subset of patients (Antonarakis et al. 2004). For phenotypes found in all Down syndrome patients, the penetrance of each phenotype varies between patients (Antonarakis et al. 2004). Despite the increasing amount of information available about the human genome and the availability of a mouse model for Down syndrome (O''Doherty et al. 2005), the genes responsible for most of the phenotypes associated with Down syndrome are still unknown (Patterson 2007; Korbel et al. 2009; Patterson 2009). Recently, detailed phenotypic analyses of as many as 30 aneuploid patients have allowed the identification of susceptibility regions for several specific phenotypes (Patterson 2007, 2009; Korbel et al. 2009; Lyle et al. 2009), but the specific genes remain to be identified. Understanding the physiology of aneuploidy is not only relevant to those individuals with aneuploid genomes but also to understanding cancer since most cancerous cells are aneuploid (Matzke et al. 2003; Pihan and Doxsey 2003; Storchova and Pellman 2004; Holland and Cleveland 2009; Williams and Amon 2009) or the consequences of copy number variation and dosage sensitivity (Dear 2009; Henrichsen et al. 2009).Plants are more tolerant of aneuploidy than animals (Matzke et al. 2003) for reasons that remain unclear. Since the discovery of the Datura trisomic “chromosome mutants” by Blakeslee (1921, 1922), viable trisomics of each chromosome type have been described in numerous species. Trisomics exhibit phenotypes specific to the identity of the triplicated chromosome (Blakeslee 1922; Khush 1973; Koornneef and Van der Veen 1983; Singh 2003). More complex aneuploids, i.e., individuals carrying more than one additional chromosome, can be viable as well and have been observed in many plants species, especially among the progeny of triploid individuals (McClintock 1929; Levan 1942; Johnsson 1945; Khush 1973). Some species appear to be more tolerant of complex aneuploidies than others, suggesting a genetic basis for aneuploidy tolerance (Satina and Blakeslee 1938; Khush 1973; Ramsey and Schemske 2002; Henry et al. 2009). Aneuploid individuals frequently appear spontaneously within polyploid plant populations, presumably due to a failure to equally partition the multiple chromosome sets at meiosis (Randolph 1935; Doyle 1986). These aneuploids exhibit few or subtle phenotypic abnormalities and can often compete with their euploid progenitors (Ramsey and Schemske 1998). Plants therefore provide an excellent opportunity for a genome-wide investigation of aneuploid syndromes: sample size is not limited, phenotypes can be described and assessed in detail, and plant aneuploid populations provide a complex mixture of viable karyotypes.In this article, we report our investigation of the relationship between phenotype and karyotype in populations of aneuploid Arabidopsis thaliana plants. All simple trisomics of A. thaliana have been previously isolated and phenotypically characterized (Steinitz-Sears 1962; Lee-Chen and Steinitz-Sears 1967; Steinitz-Sears and Lee-Chen 1970; Koornneef and Van der Veen 1983), demonstrating that they are tolerated in A. thaliana. We previously reported that aneuploid swarms—populations of aneuploid individuals of varying aneuploid karyotypes—could be obtained from the progeny of triploid A. thaliana individuals (Henry et al. 2005, 2009). Using a combination of a quantitative PCR-based method and flow cytometry, we were able to derive the full aneuploid karyotype of each of these individuals (Henry et al. 2006). We further crossed triploid A. thaliana to diploid or tetraploid individuals and demonstrated that at least 44 of the 60 possible aneuploid karyotypes that could result from these crosses (aneuploid individuals carrying between 11 and 19 chromosomes) were viable and successfully produced adult plants. Taken together, these populations and methods make it possible to explore the basis of aneuploid syndromes in A. thaliana. In this study, we were able to phenotypically characterize at least one individual from 25 different aneuploid karyotypes falling between diploidy and tetraploidy. We demonstrated that specific phenotypes are affected by the dosage of specific chromosome types. The effect of the dosage of specific chromosome types on traits was additive and could be used to predict the observed phenotype. The availability of multiple generations of aneuploid and euploid individuals allowed us to investigate potential long-term effects of aneuploidy as well as parent-of-origin effects on aneuploid phenotypes.  相似文献   

3.
《Genetics》2010,185(4):1519-1534
The distal arm of the fourth (“dot”) chromosome of Drosophila melanogaster is unusual in that it exhibits an amalgamation of heterochromatic properties (e.g., dense packaging, late replication) and euchromatic properties (e.g., gene density similar to euchromatic domains, replication during polytenization). To examine the evolution of this unusual domain, we undertook a comparative study by generating high-quality sequence data and manually curating gene models for the dot chromosome of D. virilis (Tucson strain 15010–1051.88). Our analysis shows that the dot chromosomes of D. melanogaster and D. virilis have higher repeat density, larger gene size, lower codon bias, and a higher rate of gene rearrangement compared to a reference euchromatic domain. Analysis of eight “wanderer” genes (present in a euchromatic chromosome arm in one species and on the dot chromosome in the other) shows that their characteristics are similar to other genes in the same domain, which suggests that these characteristics are features of the domain and are not required for these genes to function. Comparison of this strain of D. virilis with the strain sequenced by the Drosophila 12 Genomes Consortium (Tucson strain 15010–1051.87) indicates that most genes on the dot are under weak purifying selection. Collectively, despite the heterochromatin-like properties of this domain, genes on the dot evolve to maintain function while being responsive to changes in their local environment.EUKARYOTIC genomes are packaged into two major types of chromatin: euchromatin is gene rich and has a diffuse appearance during interphase, while heterochromatin is gene poor and remains densely packaged throughout the cell cycle (Grewal and Elgin 2002). The distal 1.2 Mb of the fourth chromosome of Drosophila melanogaster, known as the dot chromosome or Muller F element, is unusual in exhibiting an amalgamation of heterochromatic and euchromatic properties. This domain has a gene density that is similar to the other autosomes (Bartolomé et al. 2002; Slawson et al. 2006). However, it appears heterochromatic by many criteria, including late replication and very low levels of meiotic recombination (Wang et al. 2002; Arguello et al. 2010). It exhibits high levels of association with heterochromatin protein 1 (HP1) and histone H3 di- and trimethylated at lysine 9 (H3K9me2/3), as shown by immunofluorescent staining of the polytene chromosomes (Riddle and Elgin 2006; Slawson et al. 2006). This association with heterochromatin marks has recently been confirmed by the modENCODE Project [N. C. Riddle, A. Minoda, P. V. Kharchenko, A. A. Alekseyenko, Y. B. Schwartz, M. Y. Tolstorukov, A. A. Gorchakov, C. Kennedy, D. Linder-Basso, J. D. Jaffe, G. Shanower, M. I. Kuroda, V. Pirrotta, P. J. Park, S. C. R. Elgin, G. H. Karpen, and the modENCODE Consortium (http://www.modencode.org), unpublished results]. To understand this unique domain and to examine the evolution of a region with very low levels of recombination, we have undertaken a comparative study using the dot chromosome of D. virilis, a species that diverged from D. melanogaster 40–60 million years ago (Powell and Desalle 1995). We sequenced and improved the assembly of the D. virilis dot chromosome and created a manually curated set of gene models to ensure that both the assembly and the gene annotations are at a quality comparable to those in D. melanogaster. We then compared the sequence organization and gene characteristics of the distal portion of the D. virilis dot chromosome with the corresponding region from the D. melanogaster dot chromosome.In addition to examining the long-term dot chromosome evolution, we also investigated the short-term dot chromosome evolution by comparing the genomic sequences from two different strains of D. virilis. Agencourt Biosciences (AB) has previously produced a whole genome shotgun assembly of Tucson strain 15010–1051.87, while we have sequenced Tucson strain 15010–1051.88 of D. virilis [the Genomics Education Partnership (GEP) assembly]. The AB assembly has been improved by the Drosophila 12 Genomes Consortium and released as part of the comparative analysis freeze 1 (CAF1) assembly (Drosophila 12 Genomes Consortium et al. 2007).Using the GEP and CAF1 assemblies from D. virilis, and the high-quality D. melanogaster assembly and its gene annotations from FlyBase (Crosby et al. 2007), we compared the gene properties and sequence organization of the dot chromosomes and reference euchromatic and heterochromatic domains. The dot chromosomes from D. melanogaster and D. virilis are distinct from the heterochromatic and euchromatic regions of the two genomes, both in organization (e.g., repeat density) and in characteristics of the genes (e.g., size, codon bias). The two dot chromosomes resemble each other by most criteria and differ only in the types of repetitive sequences present and in relative gene order and orientation. Despite the very low rate of meiotic recombination, comparison of the two D. virilis strains shows that dot chromosome genes are under weak purifying selection. Our analysis of genes that are present in a euchromatic chromosome arm in one species and on the dot chromosome in the other (the “wanderer” genes) shows that this set of genes evolves to maintain function while responding to the changes in the local chromosomal environment.  相似文献   

4.
Andrea L. Sweigart 《Genetics》2010,184(3):779-787
Postzygotic reproductive isolation evolves when hybrid incompatibilities accumulate between diverging populations. Here, I examine the genetic basis of hybrid male sterility between two species of Drosophila, Drosophila virilis and D. americana. From these analyses, I reach several conclusions. First, neither species carries any autosomal dominant hybrid male sterility alleles: reciprocal F1 hybrid males are perfectly fertile. Second, later generation (backcross and F2) hybrid male sterility between D. virilis and D. americana is not polygenic. In fact, I identified only three genetically independent incompatibilities that cause hybrid male sterility. Remarkably, each of these incompatibilities involves the Y chromosome. In one direction of the cross, the D. americana Y is incompatible with recessive D. virilis alleles at loci on chromosomes 2 and 5. In the other direction, the D. virilis Y chromosome causes hybrid male sterility in combination with recessive D. americana alleles at a single QTL on chromosome 5. Finally, in contrast with findings from other Drosophila species pairs, the X chromosome has only a modest effect on hybrid male sterility between D. virilis and D. americana.SPECIATION occurs when populations evolve one or more barriers to interbreeding (Dobzhansky 1937; Mayr 1963). One such barrier is intrinsic postzygotic isolation, which typically evolves when diverging populations accumulate different alleles at two or more loci that are incompatible when brought together in hybrid genomes; negative epistasis between these alleles renders hybrids inviable or sterile (Bateson 1909; Dobzhansky 1937; Muller 1942). Classical and recent studies in diverse animal taxa have provided support for two evolutionary patterns that often characterize the genetics of postzygotic isolation (Coyne and Orr 1989a). The first, Haldane''s rule, observes that when there is F1 hybrid inviability or sterility that affects only one sex, it is almost always the heterogametic sex (Haldane 1922). Over the years, many researchers have tried to account for this pattern, but only two ideas are now thought to provide a general explanation: the “dominance theory,” which posits that incompatibility alleles are generally recessive in hybrids, and the “faster-male theory,” which posits that genes causing hybrid male sterility diverge more rapidly than those causing hybrid female sterility (Muller 1942; Wu and Davis 1993; Turelli and Orr 1995; reviewed in Coyne and Orr 2004). In some cases, however, additional factors might contribute to Haldane''s rule, including meiotic drive, a faster-evolving X chromosome, dosage compensation, and Y chromosome incompatibilities (reviewed in Laurie 1997; Turelli and Orr 2000; Coyne and Orr 2004).The second broad pattern affecting the evolution of postzygotic isolation is the disproportionately large effect of the X chromosome on heterogametic F1 hybrid sterility (Coyne 1992). This “large X effect” has been documented in genetic analyses of backcross hybrid sterility (e.g., Dobzhansky 1936; Grula and Taylor 1980; Orr 1987; Masly and Presgraves 2007) and inferred from patterns of introgression across natural hybrid zones (e.g., Machado et al. 2002; Saetre et al. 2003; Payseur et al. 2004). However, in only one case has the cause of the large X effect been unambiguously determined: incompatibilities causing hybrid male sterility between Drosophila mauritiana and D. sechellia occur at a higher density on the X than on the autosomes (Masly and Presgraves 2007). Testing the generality of this pattern will require additional high-resolution genetic analyses in diverse taxa (Presgraves 2008). But whatever its causes, there is now general consensus that the X chromosome often plays a special role in the evolution of postzygotic isolation (Coyne and Orr 2004).The contribution of the Y chromosome to animal speciation is less clear. Y chromosomes have far fewer genes than the X or autosomes, and most of these genes are male specific (Lahn and Page 1997; Carvalho et al. 2009). In Drosophila species, the Y chromosome is typically required for male fertility, but not for viability (Voelker and Kojima 1971). How often, then, does the Y chromosome play a role in reproductive isolation? In crosses between Drosophila species, hybrid male sterility is frequently caused by incompatibilities between the X and Y chromosomes (Schafer 1978; Heikkinen and Lumme 1998; Mishra and Singh 2007) or between the Y and heterospecific autosomal alleles (Patterson and Stone 1952; Vigneault and Zouros 1986; Lamnissou et al. 1996). In crosses between D. yakuba and D. santomea, the Y chromosome causes F1 hybrid male sterility, and accordingly, shows no evidence for recent introgression across a species hybrid zone (Coyne et al. 2004; Llopart et al. 2005). In mammals, reduced introgression of Y-linked loci (relative to autosomal loci) has been shown across natural hybrid zones of mice (Tucker et al. 1992) and rabbits (Geraldes et al. 2008), suggesting that the Y chromosome contributes to reproductive barriers.Here I examine the genetic basis of hybrid male sterility between two species of Drosophila, D. virilis and D. americana. These species show considerable genetic divergence (Ks ∼0.11, Morales-Hojas et al. 2008) and are currently allopatric: D. virilis is a human commensal worldwide with natural populations in Asia, and D. americana is found in riparian habitats throughout much of North America (Throckmorton 1982; McAllister 2002). Nearly 70 years ago, Patterson et al. (1942) showed that incompatibilities between the D. americana Y chromosome and the second and fifth chromosomes from D. virilis cause hybrid male sterility, a result that was confirmed in a more recent study (Lamnissou et al. 1996). Another study suggested that the X chromosome might play the predominant role in causing hybrid male sterility between D. virilis and D. americana (Orr and Coyne 1989). But because previous genetic analyses had to rely on only a few visible markers to map hybrid male sterility, they lacked the resolution to examine the genomic distribution of incompatibility loci.Using the D. virilis genome sequence, I have developed a dense set of molecular markers to investigate the genetic architecture of hybrid male sterility between D. virilis and D. americana. In this study, I perform a comprehensive set of crosses to address several key questions: What is the effect of the X chromosome on hybrid male sterility between D. virilis and D. americana? What is the effect of the Y chromosome? Approximately how many loci contribute to hybrid male sterility between these Drosophila species? Perhaps surprisingly, the answers to these questions differ dramatically from what has been found for other Drosophila species, including the well-studied D. melanogaster group.  相似文献   

5.
Telomere binding proteins protect chromosome ends from degradation and mask chromosome termini from checkpoint surveillance. In Saccharomyces cerevisiae, Cdc13 binds single-stranded G-rich telomere repeats, maintaining telomere integrity and length. Two additional proteins, Ten1 and Stn1, interact with Cdc13 but their contributions to telomere integrity are not well defined. Ten1 is known to prevent accumulation of aberrant single-stranded telomere DNA; whether this results from defective end protection or defective telomere replication is unclear. Here we report our analysis of a new group of ten1 temperature-sensitive (ts) mutants. At permissive temperatures, ten1-ts strains display greatly elongated telomeres. After shift to nonpermissive conditions, however, ten1-ts mutants accumulate extensive telomeric single-stranded DNA. Cdk1 activity is required to generate these single-stranded regions, and deleting the EXO1 nuclease partially suppresses ten1-ts growth defects. This is similar to cdc13-1 mutants, suggesting ten1-ts strains are defective for end protection. Moreover, like Cdc13, our analysis reveals Ten1 promotes de novo telomere addition. Interestingly, in ten1-ts strains at high temperatures, telomeric single-stranded DNA and Rad52-YFP repair foci are strongly induced despite Cdc13 remaining associated with telomeres, revealing Cdc13 telomere binding is not sufficient for end protection. Finally, unlike cdc13-1 mutants, ten1-ts strains display strong synthetic interactions with mutations in the POLα complex. These results emphasize that Cdc13 relies on Ten1 to execute its essential function, but leave open the possibility that Ten1 has a Cdc13-independent role in DNA replication.GENOME stability is critically dependent upon functional telomeres. DNA ends that lack telomeres, or that have dysfunctional telomeres, are metabolized by DNA repair processes; without an appropriate repair template, such chromosome ends can be resected or joined inappropriately with other chromosome ends. Thus, genomic integrity can be significantly compromised by telomere dysfunction, particularly in proliferating cells where cycles of instability may ensue due to creation of dicentric chromosomes (Bailey and Murnane 2006). Protein complexes that bind to the duplex and single-stranded telomere repeats are key for stabilizing the chromosome ends (de Lange 2005). In proliferating cells, this job is complicated not only because the terminal chromatin must be opened during the process of chromosome replication, but also because additional processes that metabolize DNA ends are active. For example, while nonhomologous end joining processes are preferentially used in repair of DNA double-strand breaks in G1, homologous recombination is preferentially used for this repair in S and G2 (Ferreira and Cooper 2004; Zierhut and Diffley 2008). Given these complexities, it is not surprising that our molecular understanding of how telomere proteins protect chromosomes ends is incomplete.Budding yeast has been useful for dissecting how cells correctly metabolize their chromosome ends. In Saccharomyces cerevisiae, the terminal DNA comprises approximately 300 bp of TG1-3/C1-3A sequences, ending with a short single-stranded overhang of the G-rich repeats. This 3′ overhang is ∼12–14 nucleotides, although during the late S/G2 phase of the cell cycle, it becomes longer, >30 nucleotides in length (Wellinger et al. 1993b; Dionne and Wellinger 1996; Larrivee et al. 2004). Central among factors that prevent inappropriate telomere degradation in S. cerevisiae is Cdc13, a protein that binds to single-stranded telomere G-rich repeats (Garvik et al. 1995; Lin and Zakian 1996; Nugent et al. 1996). Reducing Cdc13 function through either the cdc13-1 temperature sensitive (ts) allele or the cdc13-td conditional null (degron) allele results in telomere C-strand loss, with degradation continuing into the subtelomeric chromosomal regions (Garvik et al. 1995; Vodenicharov and Wellinger 2006). Correspondingly, homologous recombination at chromosome termini increases in cdc13-1 strains (Carson and Hartwell 1985; Garvik et al. 1995). The loss of Cdc13 unmasks the telomeres, provoking activation of the DNA damage checkpoint (Weinert and Hartwell 1993; Garvik et al. 1995). This protective role of Cdc13 is most likely its essential function.A thorough, mechanistic understanding of how Cdc13 mediates chromosome end protection is hampered in part because the activities responsible for the loss of the telomere C strand are not fully known. At normal telomeres, the Mre11-Rad50-Xrs2 complex has a role regulating resection required for telomere addition, whereas the Exo1 nuclease, Rad9 and Rad24 checkpoint proteins each influence the resection process at uncapped telomeres (Lydall and Weinert 1995; Maringele and Lydall 2002; Larrivee et al. 2004; Zubko et al. 2004). The 5′-to-3′ resection of both normal and uncapped telomeres is regulated by the activity of Cdk1, the yeast cyclin-dependent kinase (Frank et al. 2006; Vodenicharov and Wellinger 2006). Similar to the activities that promote 5′-to-3′ degradation of DNA ends at double-strand breaks (Aylon et al. 2004; Ira et al. 2004), the activities that lead to telomere resection are active in late S and G2 cell cycle phases (Wellinger et al. 1993a, 1996; Marcand et al. 2000; Vodenicharov and Wellinger 2006). Interestingly, Cdc13 is required to prevent degradation at telomeres only in proliferating cells and not when cells are blocked in stationary phase (Vodenicharov and Wellinger 2006). Additional factors, such as the S. cerevisiae Rap1 protein, prevent chromosome fusions by nonhomologous recombination during the G1 phase of the cell cycle (Pardo and Marcand 2005; Marcand et al. 2008).At least two additional proteins, Stn1 and Ten1, aid the capping role of Cdc13. Like CDC13, both STN1 and TEN1 are essential, and loss of their function leads to excessive single-stranded telomeric DNA (Grandin et al. 1997, 2001; Petreaca et al. 2007). STN1 was originally identified as a high copy suppressor of cdc13-1 temperature sensitivity (Grandin et al. 1997), and TEN1 was similarly isolated as a dosage suppressor of stn1-13 (Grandin et al. 2001). Combining either the cdc13-1 allele with stn1 mutations or the ten1-31 allele with stn1-13 is lethal (Grandin et al. 2001; Petreaca et al. 2007). The essential nature of these genes makes it difficult to clearly differentiate whether these genes operate in the same, or in parallel pathways to protect telomeres. A compelling argument that Cdc13, Stn1, and Ten1 likely function in a common pathway is that, in addition to these genetic interactions, Stn1 and Ten1 proteins interact with one another both in vivo and in vitro (Grandin et al. 2001; Gao et al. 2007), and each associates with Cdc13 in the yeast two-hybrid assay (Grandin et al. 1997, 2001; Petreaca et al. 2007). From these data, Cdc13, Stn1, and Ten1 are suggested to function as a single complex that mediates chromosome end protection in S. cerevisiae. Such a complex would share some similarities with the single-stranded DNA binding complex RPA (Gao et al. 2007). Whether these proteins normally operate exclusively as a heterotrimeric complex is still not entirely clear. Stn1 and Ten1 can make contributions to capping that are independent of Cdc13, as shown in experiments where overproducing the Stn1 essential domain with Ten1 replaced the essential function of Cdc13 (Petreaca et al. 2006). In addition, while the Schizosaccharomyces pombe Stn1 and Ten1 homologs are critical for telomere protection, they do not interact with Pot1, the single-stranded telomere binding protein that is also critical for telomere capping (Martin et al. 2007).The role of Ten1 in maintaining both telomere integrity and length homeostasis is not understood. It has been assumed that Stn1 and Ten1 play the same role as Cdc13 in maintaining telomere integrity, namely, preventing inappropriate terminal resection. However, whether this is in fact the case is not entirely clear. For one, disrupting the DNA replication machinery can give rise to an excess of terminal single-stranded DNA, although in this case, the ssDNA accumulation is attributed to a failure to synthesize the lagging DNA strand rather than removing a block to telomere resection (Diede and Gottschling 1999; Adams Martin et al. 2000). Although both Cdc13 and Stn1 are thought to act as capping proteins, each can interact with Polα subunits (Qi et al. 2003; Grossi et al. 2004; Petreaca et al. 2006), making it important to evaluate Ten1 function more carefully. Our goal here was to compare how Cdc13 and Ten1 promote chromosome end protection, first by testing whether Ten1 acts to prevent telomere resection from activities comparable to those that degrade telomeres in cdc13-1, and second by determining the impact of ten1 dysfunction upon Cdc13. The cdc13-1 allele has been extremely useful in analyzing the CDC13 essential function; TEN1 analysis has been hindered by a lack of equivalent genetic reagents. Here we have created a collection of ten1-ts alleles useful for probing the essential role of TEN1. Analysis of these alleles, which show constitutive telomere elongation, reveals that Ten1 promotes telomere capping with a similar cell cycle dependency as Cdc13, protecting ends during the period in which mitotic forms of Cdk1 are active. Critically, by showing that single-stranded DNA is generated in ten1-ts strains under conditions where semi-conservative replication is complete, we conclude that Ten1 truly can function as a capping protein. Moreover, the ten1-ts strains fail to restrain degradation of chromosome ends and induce formation of Rad52 repair foci, despite the association of wild-type Cdc13 with telomeres, indicating not only that Cdc13 binds telomeres independent of Ten1 function, but also that Cdc13 telomere localization is not sufficient for end protection. Finally, although the ten1-ts capping-deficient phenotypes parallel cdc13-1, only the ten1-ts strains are highly sensitive to impaired POL1 function, leaving open the possibility that TEN1 function additionally impacts terminal replication.  相似文献   

6.
7.
8.
9.
10.
11.
Naturally transformable bacteria acquire chromosomal DNA from related species at lower frequencies than from cognate DNA sources. To determine how genome location affects heterogamic transformation in bacteria, we inserted an nptI marker into random chromosome locations in 19 different strains of the Acinetobacter genus (>24% divergent at the mutS/trpE loci). DNA from a total of 95 nptI-tagged isolates was used to transform the recipient Acinetobacter baylyi strain ADP1. A total of >1300 transformation assays revealed that at least one nptI-tagged isolate for each of the strains/species tested resulted in detectable integration of the nptI marker into the ADP1 genome. Transformation frequencies varied up to ∼10,000-fold among independent nptI insertions within a strain. The location and local sequence divergence of the nptI flanking regions were determined in the transformants. Heterogamic transformation depended on RecA and was hampered by DNA mismatch repair. Our studies suggest that single-locus-based studies, and inference of transfer frequencies from general estimates of genomic sequence divergence, is insufficient to predict the recombination potential of chromosomal DNA fragments between more divergent genomes. Interspecies differences in overall gene content, and conflicts in local gene organization and synteny are likely important determinants of the genomewide variation in recombination rates between bacterial species.HORIZONTAL gene transfer (HGT) contributes to bacterial evolution by providing access to DNA evolved and retained in separate species or strains (Cohan 1994a,b; Bergstrom et al. 2000; Ochman et al. 2000; Feil et al. 2001; Koonin 2003; Lawrence and Hendrickson 2003; Fraser et al. 2007). Multilocus sequence typing (MLST) has provided strong evidence for frequent transfer and recombination of chromosomal DNA between related bacterial strains within the same species (Maiden et al. 1998; Enright et al. 2002). HGT occurring by natural transformation allows bacteria to exploit the presence of nucleic acids in their environment for the purposes of nutrition, DNA repair, reacquisition of lost genes, and/or acquisition of novel genetic diversity (Redfield 1993; Mehr and Seifert 1998; Dubnau 1999; Claverys et al. 2000; Szöllösi et al. 2006; Johnsen et al. 2009). It can be inferred from observations of the presence of extracellular DNA in most environments that bacteria are constantly exposed to DNA from a variety of sources, without such exposure necessarily producing observable changes in the genetic compositions of bacterial populations over evolutionary time (Thomas and Nielsen 2005; Nielsen et al. 2007a,b).The absence of sequence similarity between the donor DNA and the DNA of the recipient bacterium is the strongest barrier to the horizontal acquisition of chromosomal genes in bacteria (Matic et al. 1996; Vulic et al. 1997; Majewski 2001; Townsend et al. 2003) as illegitimate recombination occurs only at extremely low frequencies in bacteria (Hülter and Wackernagel 2008a). Single-locus transfer models have been extensively applied and have demonstrated a log-linear decrease in recombination frequencies with increasing sequence divergence for Bacillus subtilis (Roberts and Cohan, 1993; Zawadzki et al. 1995), Acinetobacter baylyi (Young and Ornston 2001), Escherichia coli (Shen and Huang 1986; Vulic et al. 1997), and Streptococcus pneumoniae (Majewski et al. 2000). For instance, heterogamic transformation between nonmutator isolates at the rpoB locus of B. mojavensis is undetectable at sequence divergences >16.7% (Zawadzki et al. 1995) and between S. pneumoniae isolates with sequence divergences >18% (Majewski et al. 2000). In A. baylyi, the nonmutator sequence divergence limit for detectable transformation at the pcaH locus of strain ADP1 was found to be 20% (Young and Ornston 2001), and up to 24% overall divergence yielded transformants at 16S rRNA loci in strain DSM587 (Strätz et al. 1996).Several recent studies also show that short stretches (<200 bp) of DNA sequence identity can facilitate additive or substitutive integration of longer stretches (>1000 bp) of heterologous DNA in bacteria (Prudhomme et al. 1991, 2002; de Vries and Wackernagel 2002; Hülter and Wackernagel 2008a). Thus, the uptake of DNA in bacteria can facilitate larger substitutions within gene sequences and the integration of additional DNA material on the basis of recombination initiated in flanking DNA stretches (either at one or both ends) with high sequence similarity (Nielsen et al. 2000). On the other hand, segments of heterologous DNA interrupting the synteny of homologous DNA have also been shown to be a barrier in intraspecies transformation in S. pneumoniae (Pasta and Sicard 1996, 1999).The various studies of the interspecies transfer potential of single genes demonstrate that the immediate local sequence divergence of the transferred locus is of high importance in determining recombination frequencies in hosts up to 20% divergent (at the housekeeping gene level). However, it can be hypothesized that the broader structural, organizational, and biochemical properties of the genome region surrounding a particular locus will determine its transfer potential to more divergent host species (Cohan 2001; Lawrence 2002). The interspecies transfer potential of various genome regions/loci between more diverged species (>20% at the housekeeping gene level) may therefore differ substantially from a log-linear model (determined experimentally for more closely related species) as local gene organization becomes less conserved with evolutionary time. The barriers to gene exchange between divergent bacterial species is likely a combination of inefficient recombination due to both mismatched base pairs (the main determinator in the log-linear model) and conflicting gene order and organization across the local recombining DNA regions. In addition, selective barriers due to negative effects on host fitness of the transferred DNA regions may become increasingly important for the removal of recombination events from the bacterial population. Recent bioinformatics-based genome analysis of E. coli and Salmonella genomes suggests various parts of the bacterial genome may have different suceptibilities to undergo evolutionarily successful recombination leading to temporal fragmentation of speciation (Lawrence 2002; Retchless and Lawrence 2007). Nevertheless, few studies have experimentally tested the effect of variable species and chromosome locations of genes on their transfer potential between bacteria (Ravin and Chen 1967; Ravin and Chakrabarti 1975; Siddiqui and Goldberg 1975; Cohan et al. 1991; Huang et al. 1991; Fall et al. 2007).Here, we determine to what extent genome location contributes to sexual isolation between the recipient A. baylyi strain ADP1 and 19 sequence divergent (24–27% divergent at the mutS/trpE loci) donor Acinetobacter strains and species (carrying a selectable nptI gene in a total of 95 random genome locations).  相似文献   

12.
13.
During lagging-strand DNA replication in eukaryotic cells primers are removed from Okazaki fragments by the flap endonuclease and DNA ligase I joins nascent fragments. Both enzymes are brought to the replication fork by the sliding clamp proliferating cell nuclear antigen (PCNA). To understand the relationship among these three components, we have carried out a synthetic lethal screen with cdc9-p, a DNA ligase mutation with two substitutions (F43A/F44A) in its PCNA interaction domain. We recovered the flap endonuclease mutation rad27-K325* with a stop codon at residue 325. We created two additional rad27 alleles, rad27-A358* with a stop codon at residue 358 and rad27-pX8 with substitutions of all eight residues of the PCNA interaction domain. rad27-pX8 is temperature lethal and rad27-A358* grows slowly in combination with cdc9-p. Tests of mutation avoidance, DNA repair, and compatibility with DNA repair mutations showed that rad27-K325* confers severe phenotypes similar to rad27Δ, rad27-A358* confers mild phenotypes, and rad27-pX8 confers phenotypes intermediate between the other two alleles. High-copy expression of POL30 (PCNA) suppresses the canavanine mutation rate of all the rad27 alleles, including rad27Δ. These studies show the importance of the C terminus of the flap endonuclease in DNA replication and repair and, by virtue of the initial screen, show that this portion of the enzyme helps coordinate the entry of DNA ligase during Okazaki fragment maturation.CELLULAR maintenance of genomic integrity is essential for the continued viability of all organisms. The fidelity of DNA replication has to be maintained and DNA insults have to be repaired to ensure that deleterious mutations are not passed on to progeny or cause cancerous growth. A number of cellular proteins have multiple roles in DNA replication, mutation avoidance, and repair. In Saccharomyces cerevisiae, the flap endonuclease, proliferating cell nuclear antigen (PCNA), and DNA ligase I encoded by RAD27, POL30, and CDC9, respectively, are all required for proper replication and also function to avoid mutation and to facilitate repair.The flap endonuclease, FEN-1 in humans, is a highly conserved structure-specific nuclease that has both endonuclease and 5′–3′ exonuclease activity. During lagging-strand replication these activities function to remove primers from Okazaki fragments, either by endonucleolytic cleavage of a flap made by strand displacement (Liu et al. 2004) or by sequential exonucleolytic removal of single nucleotides at the 5′ end of the primer (Murante et al. 1994).While deletion of RAD27 is not lethal to yeast cells, the rad27Δ mutant exhibits temperature-sensitive growth, is a mutator, and undergoes genomic instability (Johnson et al. 1995; Reagan et al. 1995; Tishkoff et al. 1997b; Chen and Kolodner 1999). In addition, its sensitivity to low doses of the methylating agent methylmethane sulfonate (MMS) implicates the participation of the enzyme in base excision repair (BER) (Reagan et al. 1995; Wu and Wang 1999). rad27Δ mutants have been reported to be either mildly sensitive to UV light or not sensitive to UV light (Reagan et al. 1995; Sommers et al. 1995). In the strain background that the mutant is mildly sensitive, its combination with rad2Δ yields a double mutant more sensitive than each single mutant, implying that the enzyme does not participate in RAD2-mediated nucleotide excision repair (NER) (Reagan et al. 1995). The flap endonuclease has also been implicated in double-strand break (DSB) repair by virtue of the incompatibility of rad27Δ with mutations of the DSB repair pathways (Tishkoff et al. 1997b; Symington 1998). In addition, either the yeast enzyme or its human ortholog has been shown to participate in reactions of homologous recombination, nonhomologous end joining, and telomere maintenance (Parenteau and Wellinger 1999, 2002; Wu et al. 1999; Wang et al. 2004; Kikuchi et al. 2005). Curiously, the rad27Δ mutant is not sensitive to gamma radiation but is sensitive to high doses of MMS that are thought to act as a radiomimetic agent (Reagan et al. 1995; Sommers et al. 1995).PCNA is the replicative clamp that acts as a scaffold to facilitate the loading of DNA replication and repair proteins, including DNA ligase I and the flap endonuclease to DNA (Warbrick 2000, 2006; Maga and Hubscher 2003). PCNA (POL30) is essential for cell viability, which is indicative of its central role in DNA metabolism. Biochemical characterization of its effect on the flap endonuclease shows that it stimulates its activity ∼50-fold, evidencing the productive nature of the interaction (Gomes and Burgers 2000; Tom et al. 2000; Frank et al. 2001; Stucki et al. 2001). The ability of DNA ligase to efficiently catalyze the formation of phosphodiester bonds in the DNA backbone may also be facilitated by its binding to PCNA. Tom et al. (2001) showed that, in vitro, PCNA enhances the ligation reaction 5-fold and that the stable association of DNA ligase with nicked duplex DNA requires PCNA.Both DNA ligase and the flap endonuclease bind to PCNA via their respective PCNA interactive peptide domains (PIP box). The PIP box is a conserved sequence motif of the amino acids QXXLXXFF. The PIP box fits into the interdomain connector loop (IDCL) of PCNA to provide a protein–protein interaction surface (Gomes and Burgers 2000; Chapados et al. 2004; Sakurai et al. 2005; Pascal et al. 2006). Mutations in the PIP box or the IDCL that impair the interaction of DNA ligase and the flap endonuclease to PCNA lead to genomic instability (Amin and Holm 1996; Eissenberg et al. 1997; Gary et al. 1999; Refsland and Livingston 2005; Subramanian et al. 2005). We have reported that the double mutants made by combinations of cdc9-p, rad27-p, and pol30-90—mutations with alterations of the PIP box or the IDCL in the respective proteins—have synergistic phenotypes with respect to MMS sensitivity and to trinucleotide repeat instability (Refsland and Livingston 2005). These results suggest that the two enzymes function in a concerted manner that is facilitated by PCNA.The precise nature of how PCNA coordinates the entry of the flap endonuclease and DNA ligase into the replication fork is not well understood. Biochemical and structural studies have begun to elucidate a possible ordering of these PCNA-mediated interactions. The possibility of such an ordering is underscored by the observation that DNA ligase adopts a toroidal conformation by completely encircling duplex DNA while interacting with PCNA (Pascal et al. 2004). Moreover, both PCNA and DNA ligase may be loaded onto the DNA in a mechanism utilizing the replication clamp loader replication factor C (RFC) (Levin et al. 2004; Vijayakumar et al. 2009), again suggesting a complete encirclement of the DNA by DNA ligase as well as by PCNA. PCNA and DNA ligase are similar in size and their interaction is likely to extend along the face of PCNA in a manner that would prevent other proteins such as the flap endonuclease from binding to the IDCL (Pascal et al. 2004, 2006). A biochemical study with purified yeast proteins showed that the two enzymes cannot bind simultaneously to PCNA (Subramanian et al. 2005). These studies suggest that a coordinated sequential interaction among PCNA, DNA ligase, and the flap endonuclease is important for replication and repair.Alternatively, both the flap endonuclease and DNA ligase may bind to the same molecule of PCNA. Since PCNA is a homotrimer, DNA ligase can potentially bind to one monomer while the flap endonuclease binds to another, using its extended C-terminal tail in a conformation allowing it to be tethered to PCNA concurrently with DNA ligase (Gomes and Burgers 2000; Sakurai et al. 2005). DNA ligase could also bind to PCNA in an extended conformation while the flap endonuclease cleaves the DNA. Sulfolobus solfataricus DNA ligase has been shown to have an open, extended conformation while binding to PCNA (Pascal et al. 2006). Presumably, once the flap endonuclease has removed the 5′ flap, DNA ligase acquires a closed, ring-shaped conformation to catalyze the joining of Okazaki fragments (Pascal et al. 2006).Exactly how the interaction of these enzymes with PCNA is coordinated in vivo, whether singly or concurrently, is not well understood. To further elucidate how the interaction of DNA ligase with PCNA is ordered, we performed a genetic screen to identify mutations that are synthetically lethal with cdc9-p (F44A/F35A), an allele of DNA ligase that has impaired binding to PCNA (Refsland and Livingston 2005; Subramanian et al. 2005). We postulated that genes recovered from this screen would function in DNA repair, replication, and recombination or would be involved in ordering the DNA ligase–PCNA interaction. From the screen we recovered a truncated allele of RAD27, rad27-K325*. This allele encodes a protein that lacks the PIP box and the entire C-terminal domain of the enzyme but retains the N terminus containing the nuclease activities. We have characterized this allele and compared it to two other rad27 alleles in which we have created different alterations of the C-terminal end of the flap endonuclease.  相似文献   

14.
We present the results of surveys of diversity in sets of >40 X-linked and autosomal loci in samples from natural populations of Drosophila miranda and D. pseudoobscura, together with their sequence divergence from D. affinis. Mean silent site diversity in D. miranda is approximately one-quarter of that in D. pseudoobscura; mean X-linked silent diversity is about three-quarters of that for the autosomes in both species. Estimates of the distribution of selection coefficients against heterozygous, deleterious nonsynonymous mutations from two different methods suggest a wide distribution, with coefficients of variation greater than one, and with the average segregating amino acid mutation being subject to only very weak selection. Only a small fraction of new amino acid mutations behave as effectively neutral, however. A large fraction of amino acid differences between D. pseudoobscura and D. affinis appear to have been fixed by positive natural selection, using three different methods of estimation; estimates between D. miranda and D. affinis are more equivocal. Sources of bias in the estimates, especially those arising from selection on synonymous mutations and from the choice of genes, are discussed and corrections for these applied. Overall, the results show that both purifying selection and positive selection on nonsynonymous mutations are pervasive.SURVEYS of DNA sequence diversity and divergence are shedding light on a number of questions in evolutionary genetics (for recent reviews, see Akey 2009; Sella et al. 2009). Two of the most important questions of this kind concern the distribution of selection coefficients against deleterious mutations affecting protein sequences and the proportion of amino acid sequence differences between related species that have been fixed by positive selection. Several different methods have been proposed for studying each of these questions, using different features of data on polymorphism and divergence at nonsynonymous and silent sites.For example, the parameters of the distribution of selection coefficients against deleterious amino acid mutations have been estimated by contrasting the numbers of nonsynonymous and silent within-species polymorphisms and fixed differences between species (Sawyer and Hartl 1992; Bustamante et al. 2002; Piganeau and Eyre-Walker 2003; Sawyer et al. 2007); by fitting the frequency spectra of nonsynonymous and silent variants to models of selection, mutation, and drift (Akashi 1999; Eyre-Walker et al. 2006; Keightley and Eyre-Walker 2007; Kryukov et al. 2007; Boyko et al. 2008; Eyre-Walker and Keightley 2009); or by comparing levels of nonsynonymous and silent diversities between species with different population sizes (Loewe and Charlesworth 2006; Loewe et al. 2006). The results of these different approaches generally agree in suggesting that there is a wide distribution of selection coefficients against nonsynonymous mutations and that the mean selection coefficient against heterozygous carriers of such mutations is very small. The results imply that a typical individual from a human population carries several hundred weakly deleterious mutations (Eyre-Walker et al. 2006; Kryukov et al. 2007; Boyko et al. 2008); for a typical Drosophila population, with its much higher level of variability, the number is probably an order of magnitude greater (Loewe et al. 2006; Keightley and Eyre-Walker 2007).The presence of this large load of slightly deleterious mutations in human and natural populations, most of which are held at low frequencies by natural selection, has many implications. From the point of view of understanding human genetic disease, it means that we have to face the likelihood that susceptibility to a disease can be influenced by variants at many loci, each with small effects (Kryukov et al. 2007). The pervasive presence of deleterious mutations throughout the genome contributes to inbreeding depression (Charlesworth and Willis 2009) and may mean that the effective population size is reduced by background selection effects, even in regions of the genome with normal levels of genetic recombination (Loewe and Charlesworth 2007). Their presence may contribute so strongly to Hill–Robertson effects (Hill and Robertson 1966; Felsenstein 1974) that they cause severely reduced levels of diversity and adaptation in low-recombination regions of the genome (Charlesworth et al. 2010) and create a selective advantage to maintaining nonzero levels of recombination (Keightley and Otto 2006; Charlesworth et al. 2010). In addition, having an estimate of the distribution of selection coefficients against deleterious nonsynonymous mutations allows their contribution to between-species divergence to be predicted, providing a way of estimating the fraction of fixed nonsynonymous differences caused by positive selection (Loewe et al. 2006; Boyko et al. 2008; Eyre-Walker and Keightley 2009).It is thus important to collect data that shed light on the properties of selection against nonsynonymous mutations in a wide range of systems and also to compare the results from different methods of estimation, since they are subject to different sources of difficulty and biases. In a previous study, we proposed the use of a comparison between two related species with different effective population sizes for this purpose (Loewe and Charlesworth 2006; Loewe et al. 2006), using Drosophila miranda and D. pseudoobscura as material. These are well suited for this type of study, as they are closely related, live together in similar habitats, and yet have very different levels of silent nucleotide diversity, indicating different effective population sizes (Ne). This study was hampered by our inability to compare the same set of loci across the two species and by the small number of loci that could be used. We here present the results of a much larger study of DNA variation at X-linked and autosomal loci for these two species, using D. affinis as a basis for estimating divergence. We compare the results, applying the method of Loewe et al. (2006) with that of Eyre-Walker and Keightley (2009) for estimating the distribution of deleterious selection coefficients and with McDonald–Kreitman test-based methods for estimating the proportion of nonsynonymous differences fixed by positive selection. While broadly confirming the conclusions from earlier studies, we note some possible sources of bias and describe methods for minimizing their effects.  相似文献   

15.
Empirical knowledge of the fitness effects of mutations is important for understanding many evolutionary processes, yet this knowledge is often hampered by several sources of measurement error and bias. Most of these problems can be solved using site-directed mutagenesis to engineer single mutations, an approach particularly suited for viruses due to their small genomes. Here, we used this technique to measure the fitness effect of 100 single-nucleotide substitutions in the bacteriophage f1, a filamentous single-strand DNA virus. We found that approximately one-fifth of all mutations are lethal. Viable ones reduced fitness by 11% on average and were accurately described by a log-normal distribution. More than 90% of synonymous substitutions were selectively neutral, while those affecting intergenic regions reduced fitness by 14% on average. Mutations leading to amino acid substitutions had an overall mean deleterious effect of 37%, which increased to 45% for those changing the amino acid polarity. Interestingly, mutations affecting early steps of the infection cycle tended to be more deleterious than those affecting late steps. Finally, we observed at least two beneficial mutations. Our results confirm that high mutational sensitivity is a general property of viruses with small genomes, including RNA and single-strand DNA viruses infecting animals, plants, and bacteria.MUTATIONAL fitness effects are relevant to many evolutionary processes. For instance, they determine the fraction of mutations that evolves neutrally (Ohta 1992), the amount of genetic variation at the mutation–selection balance (Haldane 1937), processes of fitness decay, such as Muller''s ratchet (Butcher 1995), mutational meltdown (Lynch et al. 1993), or lethal mutagenesis (Bull et al. 2007), the ability of organisms to fix beneficial mutations and evolve novel functions (Wagner 2005), or the origin of sex and recombination (Peck et al. 1997; de Visser et al. 2003). Considerable progress has been made in characterizing mutational fitness effects using model organisms or studying genetic variation in natural populations (Eyre-Walker and Keightley 2007). For instance, mutation–accumulation experiments suggest that the average effect of spontaneous deleterious mutations is 1% or lower (Kibota and Lynch 1996) in Escherichia coli, while roughly 90% of engineered gene knockouts are viable (Baba et al. 2006) and transposon insertions reduce fitness by 3% or less on average (Elena et al. 1998). In yeast, mutation–accumulation and chemical mutagenesis experiments have shown that mutations reduce fitness by 1–4% on average in diploid strains (Zeyl and de Visser 2001; Szafraniec et al. 2003; Joseph and Hall 2004). In nematodes most mutations have fitness effects lower than 1% (Keightley and Caballero 1997; Davies et al. 1999), in Drosophila the average effect of mutations ranges from 0.5 to 3.5% (Mukai et al. 1972; Ohnishi 1977; Fernández and López-Fanjul 1996; Fry et al. 1999), and, in humans, most segregating amino acid substitutions have fitness effects lower than 10% (Eyre-Walker and Keightley 2007).Although mutation–accumulation studies provide valuable information about the average effects of deleterious mutations, their power to infer the entire distribution of mutational effects, including neutral and lethal mutations, is more limited. Also, excluding bias due to selection can be problematic, and the precise location and nature of each mutation is often unknown. On the other hand, studies based on engineering mutations have been generally restricted to large deletions or insertions, which are probably infrequent in nature compared to point mutations. A direct and powerful approach that helps us to solve these difficulties consists of introducing single-nucleotide substitutions by site-directed mutagenesis. Due to their small genome sizes, viruses are excellent systems for achieving this goal. In previous work, this technique has been used for studying mutational fitness effects in several RNA viruses (Sanjuán et al. 2004; Carrasco et al. 2007; Domingo-Calap et al. 2009). However, less is known for DNA viruses—but see Domingo-Calap et al. (2009). Here, we use this approach to characterize the distribution of mutational fitness effects in the bacteriophage f1, an inovirus of the bacteriophage m13 clade, making two important improvements over previous work: first, the number of mutations tested is higher (100) and second, the contribution of experimental error to the observed distribution is explicitly accounted for. We show that one-fifth of single-nucleotide substitutions are lethal, while viable ones reduce fitness by 11% on average and can be described by a heavy-tail two-parameter distribution such as the log-normal. Interestingly, the fraction of beneficial mutations is unexpectedly high. We also compare the average effects of different mutation types and of mutations affecting different genes.  相似文献   

16.
A small cluster of dioecious species in the plant genus Silene has evolved chromosomal sex determination and sex chromosomes relatively recently, within the last 10 million years (MY). Five dioecious Silene species (section Elisanthe) are very closely related (1–2 MY of divergence) and it was previously thought that all five have similar sex chromosomes. Here we demonstrate that in one of these species, Silene diclinis, the sex chromosomes have been significantly rearranged, resulting in the formation of neo-sex chromosomes. Fluorescence in situ hybridization with genic and repetitive probes revealed that in S. diclinis a reciprocal translocation has occurred between the ancestral Y chromosome and an autosome, resulting in chromosomes designated Y1 and Y2. Both Y1 and Y2 chromosomes are male specific. Y1 pairs with the X chromosome and with the autosome (the neo-X), which cosegregates with X. Y2 pairs only with the neo-X, forming a chain X-Y1-neo-X-Y2 in male meiosis. Despite very recent formation of the neo-sex chromosomes in S. diclinis, they are present in all surveyed individuals throughout the species range. Evolution of neo-sex chromosomes may be the cause of partial reproductive isolation of this species and could have been the isolating mechanism that drove speciation of S. diclinis.PAIRING of homologous chromosomes during meiosis, in the majority of diploid plants and animals, leads to the formation of bivalents at first metaphase and subsequently the correct segregation of the chromosomes. Chromosomal translocations that produce multivalents usually result in unbalanced segregation, which consequently affects fertility. However, chain or ring configurations appear to be stably inherited in some species. An extreme example is found in the plant genus Oenothera, where many species display a ring involving all 14 chromosomes (Cleland 1972). In animals these configurations may include sex chromosomes, resulting in the formation of multiple X and Y chromosomes. For example, the monotreme platypus possesses five X and five Y chromosomes that form a chain of alternating X and Y chromosomes in male meiosis (Bick and Sharman 1975; Gruetzner et al. 2006). Such chains are formed due to several interchromosomal translocation events, including sex chromosome–autosome translocations (Gruetzner et al. 2006). Since sex chromosomes are rare in plants, examples of plant sex-linked chromosome multiples have been reported on only a few occasions. A chain of four X and five Y has been identified in an East African mistletoe Viscum fischeri (Wiens and Barlow 1975) and a chain of two X and two Y has been found in Humulus lupulus ssp. cordifolius (Shephard et al. 2000). Trivalent formation comprising Y1 X Y2 has been observed both in H. japonicus (Shephard et al. 2000) and in a number of dioecious species in the genus Rumex (Cunado et al. 2007; Navajas-Perez et al. 2009). Here we report that the plant species Silene diclinis has multiple sex chromosomes that form a chain of four during meiosis metaphase I.S. diclinis is a member of a small group of dioecious species (having separate male and female plants) in section Elisanthe in the plant genus Silene (Caryophyllaceae). The other members of this group are S. latifolia, S. dioica, S. heuffelii, and S. marizii (Prentice 1978). The presence of large heteromorphic sex chromosomes in S. latifolia and S. dioica has been known for many years (Westergaard 1958). Due to the ease of cytogenetic identification of the sex chromosomes, the clear morphological difference between the sexes and the short generation time, S. latifolia was used in early genetic research concerning sex determination in plants. The male was shown to be the heterogametic sex (XY) with the larger Y chromosome having a decisive role in sex determination (Westergaard 1958). Since then, S. latifolia has become a species of choice for studies in plant genetics, ecology, and evolution (Bernasconi et al. 2009). It is particularly useful for studies of sex chromosome evolution because the sex chromosomes in Silene are of relatively recent origin compared to those of mammals (Charlesworth 2002; Ming and Moore 2007; Marais et al. 2008).Experimental crosses involving all five dioecious species in Silene section Elisanthe in various pairwise combinations have produced viable hybrids and, although some combinations were less successful than others, the formation of these hybrids suggests a close relationship within this group (Prentice 1978). This close relationship is also illustrated by DNA sequence comparisons that show that interspecific silent divergence between these species does not exceed 2%, which is comparable to intraspecific polymorphism in S. latifolia (Ironside and Filatov 2005). S. diclinis is a rare and restricted endemic, found only in Southern Valencia, Spain in an area smaller than 18 × 9 km (Prentice 1976; Montesinos et al. 2006). Of the other four Elisanthe species, only S. latifolia occurs in this region, and experimental crosses between these two species are the least successful (Prentice 1978). Hybrids between S. latifolia and S. dioica occur naturally in regions where their populations coincide (Baker 1948) but no natural hybrids of S. diclinis and S. latifolia have been reported.Cytogenetic analysis of S. diclinis has been limited. Examination of mitotic metaphase spreads in root tip squash preparations from adult male and female plants indicated that the male had one X and one Y chromosome. Both chromosomes were large but the difference between them was slight (van Nigtevecht and Prentice 1985). Regular pairing of chromosomes with 12 bivalents at metaphase I in pollen mother cells has been reported (Morisset and Bozman 1969). However, these observations were made without the benefit of a marker for the Y chromosome. Recently, sequences with homology to an Ogre retrotransposon have been isolated from S. latifolia and used as probes in fluorescence in situ hybridization (FISH) experiments on mitotic (Cermak et al. 2008) and both mitotic and meiotic (Filatov et al. 2009) chromosome spreads. The pattern of hybridization showed that these sequences are widespread over the X chromosome and all of the autosomes but are mainly confined to a small section at the pairing region of the Y chromosome in S. latifolia. Therefore, these probes “paint” all the chromosomes apart from the Y, providing a “negative paint” for the Y chromosome. By using one of these probes (clone 4.2) on meiotic spreads of S. dioica and S. marizii, we confirmed that these species have sex chromosomes similar to those of S. latifolia (Filatov et al. 2009). The X and Y formed a rod bivalent and the Y chromosome was larger than both the X and autosomes.In this article we report our FISH experiments with S. diclinis using the negative paint probe together with probes containing S. latifolia sex-linked gene sequences. We demonstrate that S. diclinis males have two Y chromosomes that differ in the distribution of the paint signal and these gene sequences. In meiotic metaphase I, one Y pairs with the X and an autosome while the second Y pairs with the other arm of this autosome, forming a chain of four chromosomes. We suggest that an autosome–Y reciprocal translocation was involved in the evolution of neo-sex chromosomes in this species.  相似文献   

17.
18.
19.
20.
Using zinc-finger nucleases (ZFNs) to cleave the chromosomal target, we have achieved high frequencies of gene targeting in the Drosophila germline. Both local mutagenesis through nonhomologous end joining (NHEJ) and gene replacement via homologous recombination (HR) are stimulated by target cleavage. In this study we investigated the mechanisms that underlie these processes, using materials for the rosy (ry) locus. The frequency of HR dropped significantly in flies homozygous for mutations in spnA (Rad51) or okr (Rad54), two components of the invasion-mediated synthesis-dependent strand annealing (SDSA) pathway. When single-strand annealing (SSA) was also blocked by the use of a circular donor DNA, HR was completely abolished. This indicates that the majority of HR proceeds via SDSA, with a minority mediated by SSA. In flies deficient in lig4 (DNA ligase IV), a component of the major NHEJ pathway, the proportion of HR products rose significantly. This indicates that most NHEJ products are produced in a lig4-dependent process. When both spnA and lig4 were mutated and a circular donor was provided, the frequency of ry mutations was still high and no HR products were recovered. The local mutations produced in these circumstances must have arisen through an alternative, lig4-independent end-joining mechanism. These results show what repair pathways operate on double-strand breaks in this gene targeting system. They also demonstrate that the outcome can be biased toward gene replacement by disabling the major NHEJ pathway and toward simple mutagenesis by interfering with the major HR process.EXPERIMENTAL gene targeting relies on cellular DNA repair activities. When a donor DNA carrying the desired sequence modifications is introduced into cells or organisms, successful gene replacement depends on cellular capabilities for homologous recombination (HR).We have developed a very efficient gene targeting procedure for Drosophila based on target cleavage by designed zinc-finger nucleases (ZFNs) (Bibikova et al. 2002, 2003; Beumer et al. 2006). Because the DNA-binding domain consists of Cys2His2 zinc fingers, these hybrid proteins are very flexible in their recognition capabilities. Each finger makes contact primarily with 3 bp of DNA, and arrays of three to four fingers provide sufficient affinity for in vivo binding. Since two ZFNs are required to cleave any single target, a pair of three-finger proteins provides adequate specificity, in principle, to attack a unique genomic sequence.When a double-strand break (DSB) is created at a specific site in the genome, DNA sequence changes result either from HR with a marked donor DNA or from inaccurate nonhomologous end joining (NHEJ). In this study we set out to determine which cellular activities support each of these processes and to learn whether the repair outcome could be biased by elimination of one or another pathway.Earlier studies showed that Drosophila uses DSB repair mechanisms that are very similar to other eukaryotic organisms (Wyman and Kanaar 2006). In the realm of HR, homologs of the Rad51 (spnA) and Rad54 (okr) proteins are required for the break-initiated meiotic recombination events needed for proper chromosome segregation in females (Kooistra et al. 1997, 1999; Ghabrial et al. 1998; Staeva-Vieira et al. 2003). Mutations in both these genes sensitize somatic cells in early developmental stages to ionizing radiation (IR) and to other DNA damaging agents. In yeast, mutations in the RAD51 gene sensitize cells to IR and lead to severe sporulation defects (Symington 2002). Mutations in RAD54 also confer sensitivity to DNA damaging agents, but are less severely affected in meiosis. In mice absence of the Rad51 protein is lethal in early embryonic development (Lim and Hasty 1996; Tsuzuki et al. 1996). Absence of Rad54 is tolerable, but confers sensitivity to IR and other agents (Essers et al. 1997).The Drosophila genome encodes components of the major NHEJ pathway, including DNA ligase IV (lig4), Xrcc4, and the Ku proteins (ku70, ku80). Loss of Lig4 sensitizes early developmental stages to ionizing radiation, and this effect is more severe in the absence of Rad54 (Gorski et al. 2003). In other assays a considerable amount of end joining still occurs in lig4 mutants (McVey et al. 2004c; Romeijn et al. 2005), suggesting a secondary or backup pathway, as has been observed in other organisms (Nussenzweig and Nussenzweig 2007). Yeasts rely more heavily on HR for DSB repair, so lig4 mutations have little effect unless HR is impaired. In contrast, lig4−/− mice die early in embryogenesis (Barnes et al. 1998), although they can be rescued by elimination of p53 (Frank et al. 2000).The molecular process of DSB repair by HR has been studied in Drosophila by introducing a single break at a unique target either by P-element excision or by I-SceI cleavage. The evidence strongly points to an invasion and copying mechanism called synthesis-dependent strand annealing (SDSA) (see below) (Kurkulos et al. 1994; Nassif et al. 1994; McVey et al. 2004a). These events are largely dependent on spnA (McVey et al. 2004a; Johnson-Schlitz et al. 2007; Wei and Rong 2007), okr (Johnson-Schlitz et al. 2007; Wei and Rong 2007), and other factors, including mus309 (the Drosophila Bloom syndrome protein, DmBlm) (Adams et al. 2003; McVey et al. 2004b, 2007; Johnson-Schlitz and Engels 2006). When the break site is surrounded by direct repeats, repair proceeds efficiently by single-strand annealing (SSA) (Rong and Golic 2003; Preston et al. 2006).The key difference between SDSA and SSA is the mechanistic requirement for strand invasion in the former. SSA has rather modest genetic dependencies and is independent of Rad51 and Rad54, but requires that all participating molecules have ends (Symington 2002; Wyman and Kanaar 2006; Johnson-Schlitz et al. 2007; Wei and Rong 2007). In yeast, SSA is reduced in rad52 mutants, but Drosophila has no identified homolog of this gene.In this study we examined the effects of null mutations in the spnA (Rad51), okr (Rad54), and lig4 genes on ZFN-induced targeting of the Drosophila rosy (ry) locus (Beumer et al. 2006). To reveal the role of SSA, we also compared linear and circular presentation of the donor DNA.  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