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 78 毫秒
1.
2.
Despite the widespread study of genetic variation in admixed human populations, such as African-Americans, there has not been an evaluation of the effects of recent admixture on patterns of polymorphism or inferences about population demography. These issues are particularly relevant because estimates of the timing and magnitude of population growth in Africa have differed among previous studies, some of which examined African-American individuals. Here we use simulations and single-nucleotide polymorphism (SNP) data collected through direct resequencing and genotyping to investigate these issues. We find that when estimating the current population size and magnitude of recent growth in an ancestral population using the site frequency spectrum (SFS), it is possible to obtain reasonably accurate estimates of the parameters when using samples drawn from the admixed population under certain conditions. We also show that methods for demographic inference that use haplotype patterns are more sensitive to recent admixture than are methods based on the SFS. The analysis of human genetic variation data from the Yoruba people of Ibadan, Nigeria and African-Americans supports the predictions from the simulations. Our results have important implications for the evaluation of previous population genetic studies that have considered African-American individuals as a proxy for individuals from West Africa as well as for future population genetic studies of additional admixed populations.STUDIES of archeological and genetic data show that anatomically modern humans originated in Africa and more recently left Africa to populate the rest of the world (Tishkoff and Williams 2002; Barbujani and Goldstein 2004; Garrigan and Hammer 2006; Reed and Tishkoff 2006; Campbell and Tishkoff 2008; Jakobsson et al. 2008; Li et al. 2008). Given the central role Africa has played in the origin of diverse human populations, understanding patterns of genetic variation and the demographic history of populations within Africa is important for understanding the demographic history of global human populations. The availability of large-scale single-nucleotide polymorphism (SNP) data sets coupled with recent advances in statistical methodology for inferring parameters in population genetic models provides a powerful means of accomplishing these goals (Keinan et al. 2007; Boyko et al. 2008; Lohmueller et al. 2009; Nielsen et al. 2009).It is important to realize that studies of African demographic history using genetic data have come to qualitatively different conclusions regarding important parameters. Some recent studies have found evidence for ancient (>100,000 years ago) two- to fourfold growth in African populations (Adams and Hudson 2004; Marth et al. 2004; Keinan et al. 2007; Boyko et al. 2008). Other studies have found evidence of very recent growth (Pluzhnikov et al. 2002; Akey et al. 2004; Voight et al. 2005; Cox et al. 2009; Wall et al. 2009) or could not reject a model with a constant population size (Pluzhnikov et al. 2002; Voight et al. 2005). It is unclear why studies found such different parameter estimates. However, these studies all differ from each other in the amount of data considered, the types of data used (e.g., SNP genotypes vs. full resequencing), the genomic regions studied (e.g., noncoding vs. coding SNPs), and the types of demographic models considered (e.g., including migration vs. not including migration postseparation of African and non-African populations).Another important way in which studies of African demographic history differ from each other is in the populations sampled. Some studies have focused on genetic data from individuals sampled from within Africa (Pluzhnikov et al. 2002; Adams and Hudson 2004; Voight et al. 2005; Keinan et al. 2007; Cox et al. 2009; Wall et al. 2009), while other studies included American individuals with African ancestry (Adams and Hudson 2004; Akey et al. 2004; Marth et al. 2004; Boyko et al. 2008). While there is no clear correspondence between those studies which sampled native African individuals (as opposed to African-Americans) and particular growth scenarios, it is clear from previous studies that African-American populations do differ from African populations in their recent demographic history. In particular, genetic studies suggest that there is wide variation in the degree of European admixture in most African-American individuals in the United States and that they have, on average, ∼80% African ancestry and 20% European ancestry (Parra et al. 1998; Pfaff et al. 2001; Falush et al. 2003; Patterson et al. 2004; Tian et al. 2006; Lind et al. 2007; Reiner et al. 2007; Price et al. 2009; Bryc et al. 2010). Furthermore, both historical records and genetic evidence suggest that the admixture process began quite recently, within the last 20 generations (Pfaff et al. 2001; Patterson et al. 2004; Seldin et al. 2004; Tian et al. 2006). Recent population admixture can alter patterns of genetic variation in a discernible and predictable way. For example, recently admixed populations will exhibit correlation in allele frequencies (i.e., linkage disequilibrium) among markers that differ in frequency between the parental populations. This so-called admixture linkage disequilibrium (LD) (Chakraborty and Weiss 1988) can extend over long physical distances (Lautenberger et al. 2000) and decays exponentially with time the since the admixture process began (i.e., recently admixed populations typically exhibit LD over a longer physical distance than anciently admixed populations).While it is clear that African-American populations have a different recent demographic history than do African populations from within Africa and that admixture tracts can be identified in admixed individuals (Falush et al. 2003; Patterson et al. 2004; Tang et al. 2006; Sankararaman et al. 2008a,b; Price et al. 2009; Bryc et al. 2010), the effect that admixture has on other patterns of genetic variation remains unclear. For example, Xu et al. (2007) found similar LD decay patterns when comparing African-American and African populations. It is also unclear whether the recent admixture affects our ability to reconstruct ancient demographic events (such as expansions that predate the spread of humans out of Africa) from whole-genome SNP data. Most studies of demographic history have summarized the genome-wide SNP data by allele frequency or haplotype summary statistics. If these summary statistics are not sensitive to the recent European admixture, then the African-American samples may yield estimates of demographic parameters that are close to the true demographic parameters for the ancestral, unsampled, African populations. This would suggest that the differences in growth parameter estimates obtained from African populations cannot be explained by certain studies sampling African-American individuals and others sampling African individuals from within Africa. However, if these statistics are sensitive to recent admixture, then they may give biased estimates of growth parameters.Here, we examine the effect of recent admixture on the estimation of population demography. In particular, we estimate growth parameters from simulated data sets using SNP frequencies as well as a recently developed haplotype summary statistic (Lohmueller et al. 2009). We compare the demographic parameter estimates made from the admixed and nonadmixed populations and find that some parameter estimates are qualitatively similar between the two populations when inferred using allele frequencies. Inferences of growth using haplotype-based approaches appear to be more sensitive to recent admixture than inferences based on SNP frequencies. We discuss implications that our results have for interpreting studies of demography in admixed populations.  相似文献   

3.
4.
The importance of genes of major effect for evolutionary trajectories within and among natural populations has long been the subject of intense debate. For example, if allelic variation at a major-effect locus fundamentally alters the structure of quantitative trait variation, then fixation of a single locus can have rapid and profound effects on the rate or direction of subsequent evolutionary change. Using an Arabidopsis thaliana RIL mapping population, we compare G-matrix structure between lines possessing different alleles at ERECTA, a locus known to affect ecologically relevant variation in plant architecture. We find that the allele present at ERECTA significantly alters G-matrix structure—in particular the genetic correlations between branch number and flowering time traits—and may also modulate the strength of natural selection on these traits. Despite these differences, however, when we extend our analysis to determine how evolution might differ depending on the ERECTA allele, we find that predicted responses to selection are similar. To compare responses to selection between allele classes, we developed a resampling strategy that incorporates uncertainty in estimates of selection that can also be used for statistical comparisons of G matrices.THE structure of the genetic variation that underlies phenotypic traits has important consequences for understanding the evolution of quantitative traits (Fisher 1930; Lande 1979; Bulmer 1980; Kimura 1983; Orr 1998; Agrawal et al. 2001). Despite the infinitesimal model''s allure and theoretical tractability (see Orr and Coyne 1992; Orr 1998, 2005a,b for reviews of its influence), evidence has accumulated from several sources (artificial selection experiments, experimental evolution, and QTL mapping) to suggest that genes of major effect often contribute to quantitative traits. Thus, the frequency and role of genes of major effect in evolutionary quantitative genetics have been a subject of intense debate and investigation for close to 80 years (Fisher 1930; Kimura 1983; Orr 1998, 2005a,b). Beyond the conceptual implications, the prevalence of major-effect loci also affects our ability to determine the genetic basis of adaptations and species differences (e.g., Bradshaw et al. 1995, 1998).Although the existence of genes of major effect is no longer in doubt, we still lack basic empirical data on how segregating variation at such genes affects key components of evolutionary process (but see Carrière and Roff 1995). In other words, How does polymorphism at genes of major effect alter patterns of genetic variation and covariation, natural selection, and the likely response to selection? The lack of data stems, in part, from the methods used to detect genes of major effect: experimental evolution (e.g., Bull et al. 1997; Zeyl 2005) and QTL analysis (see Erickson et al. 2004 for a review) often detect such genes retrospectively after they have become fixed in experimental populations or the species pairs used to generate the mapping population. The consequences of polymorphism at these genes on patterns of variation, covariation, selection, and the response to selection—which can be transient (Agrawal et al. 2001)—are thus often unobserved.A partial exception to the absence of data on the effects of major genes comes from artificial selection experiments, in which a substantial evolutionary response to selection in the phenotype after a plateau is often interpreted as evidence for the fixation of a major-effect locus (Frankham et al. 1968; Yoo 1980a,b; Frankham 1980; Shrimpton and Robertson 1988a,b; Caballero et al. 1991; Keightley 1998; see Mackay 1990 and Hill and Caballero 1992 for reviews). However, many of these experiments report only data on the selected phenotype (e.g., bristle number) or, alternatively, the selected phenotype and some measure of fitness (e.g., Frankham et al. 1968, Yoo 1980b; Caballero et al. 1991; Mackay et al. 1994; Fry et al. 1995; Nuzhdin et al. 1995; Zur Lage et al. 1997), making it difficult to infer how a mutation will affect variation, covariation, selection, and evolutionary responses for a suite of traits that might affect fitness themselves. One approach is to document how variation at individual genes of major effect affects the genetic variance–covariance matrix (“G matrix”; Lande 1979), which represents the additive genetic variance and covariance between traits.Although direct evidence for variation at major-effect genes altering patterns of genetic variation, covariation, and selection is rare, there is abundant evidence for the genetic mechanisms that could produce these dynamics. A gene of major effect could have these consequences due to any of at least three genetic mechanisms: (1) pleiotropy, where a gene of major effect influences several traits, including potentially fitness, simultaneously, (2) physical linkage or linkage disequilibrium (LD), in which a gene of major effect is either physically linked or in LD with other genes that influence other traits under selection, and (3) epistasis, in which the allele present at a major-effect gene alters the phenotypic effect of other loci and potentially phenotypes under selection. Evidence for these three evolutionary genetic mechanisms leading to changes in suites of traits comes from a variety of sources, including mutation accumulation experiments (Clark et al. 1995; Fernandez and Lopez-Fanjul 1996), mutation induction experiments (Keightley and Ohnishi 1998), artificial selection experiments (Long et al. 1995), and transposable element insertions (Rollmann et al. 2006). For pleiotropy in particular, major-effect genes that have consequences on several phenotypic traits are well known from the domestication and livestock breeding literature [e.g., myostatin mutations in Belgian blue cattle and whippets (Arthur 1995; Grobet et al. 1997; Mosher et al. 2007), halothane genes in pigs (Christian and Rothschild 1991; Fujii et al. 1991), and Booroola and Inverdale genes in sheep (Amer et al. 1999; Visscher et al. 2000)]. While these data suggest that variation at major-effect genes could—and probably does—influence variation, covariation, and selection on quantitative traits, data on the magnitude of these consequences remain lacking.Recombinant inbred line (RIL) populations are a promising tool for investigating the influence of major-effect loci. During advancement of the lines from F2''s to RILs, alternate alleles at major-effect genes (and most of the rest of the genome) will be made homozygous, simplifying comparisons among genotypic classes. Because of the high homozygosity, individuals within RILs are nearly genetically identical, facilitating phenotyping of many genotypes under a range of environments. In addition, because of recombination, alternative alleles are randomized across genetic backgrounds—facilitating robust comparisons between sets of lines differing at a major-effect locus.Here we investigate how polymorphism at an artificially induced mutation, the erecta locus in Arabidopsis thaliana, affects the magnitude of these important evolutionary genetic parameters under ecologically realistic field conditions. We use the Landsberg erecta (Ler) × Columbia (Col) RIL population of A. thaliana to examine how variation at a gene of major effect influences genetic variation, covariation, and selection on quantitative traits in a field setting. The Ler × Col RIL population is particularly suitable, because it segregates for an artificially induced mutation at the erecta locus, which has been shown to influence a wide variety of plant traits. The Ler × Col population thus allows a powerful test of the effects of segregating variation at a gene—chosen a priori—with numerous pleiotropic effects. The ERECTA gene is a leucine-rich receptor-like kinase (LRR-RLK) (Torii et al. 1996) and has been shown to affect plant growth rates (El-Lithy et al. 2004), stomatal patterning and transpiration efficiency (Masle et al. 2005; Shpak et al. 2005), bacterial pathogen resistance (Godiard et al. 2003), inflorescence and floral organ size and shape (Douglas et al. 2002; Shpak et al. 2003, 2004), and leaf polarity (Xu et al. 2003; Qi et al. 2004).Specifically, we sought to answer the following questions: (1) Is variation at erecta significantly associated with changes to the G matrix? (2) Is variation at erecta associated with changes in natural selection on genetically variable traits? And (3) is variation at erecta associated with significantly different projected evolutionary responses to selection?  相似文献   

5.
6.
7.
While mitochondria are renowned for their role in energy production, they also perform several other integral functions within the cell. Thus, it is not surprising that mitochondrial dysfunction can negatively impact cell viability. Although mitochondria have received an increasing amount of attention in recent years, there is still relatively little information about how proper maintenance of mitochondria and its genomes is achieved. The Neurospora crassa mus-10 mutant was first identified through its increased sensitivity to methyl methanesulfonate (MMS) and was thus believed to be defective in some aspect of DNA repair. Here, we report that mus-10 harbors fragmented mitochondria and that it accumulates deletions in its mitochondrial DNA (mtDNA), suggesting that the mus-10 gene product is involved in mitochondrial maintenance. Interestingly, mus-10 begins to senesce shortly after deletions are visualized in its mtDNA. To uncover the function of MUS-10, we used a gene rescue approach to clone the mus-10 gene and discovered that it encodes a novel F-box protein. We show that MUS-10 interacts with a core component of the Skp, Cullin, F-box containing (SCF) complex, SCON-3, and that its F-box domain is essential for its function in vivo. Thus, we provide evidence that MUS-10 is part of an E3 ubiquitin ligase complex involved in maintaining the integrity of mitochondria and may function to prevent cellular senescence.THE mus-10 mutant was isolated from a screen aimed at identifying Neurospora crassa strains that were sensitive to MMS and therefore likely to lack proper DNA repair mechanisms (Kafer and Perlmutter 1980). Epistasis analyses involving mus-10 suggested that it belonged to the uvs-6 epistasis group, which functions in recombination repair (Kafer and Perlmutter 1980; Kafer 1983). However, mus-10 did not display several phenotypes common to other members of the uvs-6 epistasis group: chromosomal instability, a high sensitivity to histidine, and the inability to produce viable ascospores in homozygous crosses (Newmeyer et al. 1978; Newmeyer and Galeazzi 1978; Kafer and Perlmutter 1980; Kafer 1981; Schroeder 1986; Watanabe et al. 1997; Handa et al. 2000; Sakuraba et al. 2000). Furthermore, the frequencies of spontaneous and radiation-induced mutation observed in mus-10 were similar to those of a wild-type strain (Kafer 1981). Past efforts to uncover the nature of these discrepancies or the function of the mus-10 gene product have been uninformative.The majority of cellular ATP is produced in mitochondria through aerobic respiration, which couples electron flow through respiratory complexes within the mitochondrial inner membrane with oxidative phosphorylation. Besides their role in ATP synthesis, mitochondria are also involved in many other cellular processes including beta-oxidation (Bartlett and Eaton 2004), calcium homeostasis (Gunter et al. 2004; Rimessi et al. 2008), production of iron-sulfur clusters (Zheng et al. 1998; Gerber and Lill 2002; Lill and Muhlenhoff 2005; Rouault and Tong 2005), and apoptosis (Green 2005; Antignani and Youle 2006; Xu and Shi 2007). Although virtually all mitochondrial proteins are encoded within the nucleus, a small number of proteins are encoded by mitochondrial DNA (mtDNA). The integrity of the mitochondrial genome may affect cell survival as mutations in mtDNA accumulate in patients suffering from severe neurological diseases including Alzheimer''s, Huntington''s and Parkinson''s, as well as several types of cancer (Chatterjee et al. 2006; Higuchi 2007; Krishnan et al. 2007; Reeve et al. 2008). The number of mtDNA mutations also increases with age, suggesting a link between mitochondrial dysfunction and ageing (Cortopassi and Arnheim 1990; Corral-Debrinski et al. 1992; Cortopassi et al. 1992; Simonetti et al. 1992; Reeve et al. 2008). Contrary to the single genome in the nucleus, there are several copies of mtDNA in each mitochondrion. Thus, defects in a few mitochondrial genomes do not necessarily lead to mitochondrial dysfunction. Many patients suffering from mitochondrial diseases exhibit heteroplasmy, a phenomenon in which a mixture of wild-type and mutant mtDNAs exist in a single cell. The ratio of wild-type to mutant mtDNAs is critical in determining the penetrance of the genetic defect, where mutant loads >60% are required to cause respiratory chain dysfunction within an individual cell (Boulet et al. 1992; Chomyn et al. 1992; Sciacco et al. 1994).Even though N. crassa strains are generally deemed immortal if they can be subcultured ∼50 times, a wild-type strain was recently reported to senesce after 12,000 hr of growth, implying that this fungus undergoes natural or programmed ageing (Maheshwari and Navaraj 2008; Kothe et al. 2010). However, replicative life span is also influenced by genetic background as certain mutations can cause progressive deterioration of growth, ultimately leading to death. One such example is the nuclear-encoded natural death (nd), which when mutant causes a senescence phenotype correlating with the accumulation of multiple mtDNA deletions (Sheng 1951; Seidel-Rogol et al. 1989). The deletions of mtDNA in nd occurred between two 70- to 701-bp direct repeats, suggesting that the nd gene product regulates recombination, repair, or replication of mtDNA (Bertrand et al. 1993). Another nuclear mutation, senescence (sen), was isolated from N. intermedia and introgressed into N. crassa (Navaraj et al. 2000). Deletions were also observed in the mtDNA of sen mutants, but unlike those occurring in nd were flanked by 6- to 10-bp repeats typically associated with GC-rich palindromic sequences (D''Souza et al. 2005). The nature of the sequences that flanked the mtDNA deletions in these two mutants supported the existence of two distinct systems of mtDNA recombination in N. crassa: a general system of homologous recombination (system I) and a site-specific mechanism (system II), mediated in part by nd and sen, respectively (Bertrand et al. 1993; D''Souza et al. 2005). The nd and sen mutations have been mapped to linkage groups I and V, respectively, but neither gene has been cloned and the precise function of their gene products remains unclear. Two ultraviolet (UV)-sensitive mutants, uvs-4 and uvs-5, are thought to undergo senescence, but unfortunately, these strains have not been studied in great detail (Schroeder 1970; Perkins et al. 1993; Hausner et al. 2006). Premature senescence has also been observed in cytoplasmic mutants of N. crassa including the E35 and ER-3 stopper mutants that harbor large mtDNA deletions, as well as strains that accumulate mitochondrial plasmids capable of inserting into mtDNA through homologous recombination (de Vries et al. 1986; Akins et al. 1989; Myers et al. 1989; Niagro and Mishra 1989; Court et al. 1991; Alves and Videira 1998).While trying to establish the role of MUS-10 in DNA repair, we discovered that the mus-10 mutant exhibited a shortened life span, an abnormal mitochondrial morphology and mtDNA instability. We cloned the mus-10 gene through its ability to complement the MMS sensitivity of the mus-10 mutant and revealed that it encoded a novel F-box protein. This suggested that MUS-10 is part of an Skp, Cullin, F-box containing (SCF) E3 ubiquitin ligase complex that targets proteins for degradation by the 26S proteasome. The data we present in this article offer proof that an SCF complex can regulate both mitochondrial maintenance and cellular senescence.  相似文献   

8.
9.
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   

10.
It is widely recognized that the mixed linear model is an important tool for parameter estimation in the analysis of complex pedigrees, which includes both pedigree and genomic information, and where mutually dependent genetic factors are often assumed to follow multivariate normal distributions of high dimension. We have developed a Bayesian statistical method based on the decomposition of the multivariate normal prior distribution into products of conditional univariate distributions. This procedure permits computationally demanding genetic evaluations of complex pedigrees, within the user-friendly computer package WinBUGS. To demonstrate and evaluate the flexibility of the method, we analyzed two example pedigrees: a large noninbred pedigree of Scots pine (Pinus sylvestris L.) that includes additive and dominance polygenic relationships and a simulated pedigree where genomic relationships have been calculated on the basis of a dense marker map. The analysis showed that our method was fast and provided accurate estimates and that it should therefore be a helpful tool for estimating genetic parameters of complex pedigrees quickly and reliably.MUCH effort in genetics has been devoted to revealing the underlying genetic architecture of quantitative or complex traits. Traditionally, the polygenic model has been used extensively to estimate genetic variances and breeding values of natural and breeding populations, where an infinite number of genes is assumed to code for the trait of interest (Bulmer 1971; Falconer and Mackay 1996). The genetic variance of a quantitative trait can be decomposed into an additive part that corresponds to the effects of individual alleles and a part that is nonadditive because of interactions between alleles. Attention has generally been focused on the estimation of additive genetic variance (and heritability), since additive variation is directly proportional to the response of selection via the breeder''s equation (Falconer and Mackay 1996, Chap. 11). However, to estimate additive genetic variation and heritability accurately, it can be important to identify potential nonadditive sources in genetic evaluations (Misztal 1997; Ovaskainen et al. 2008; Waldmann et al. 2008), especially if the pedigree being analyzed contains a large proportion of full-sibs and clones, as these in particular give rise to nonadditive genetic relationships (Lynch and Walsh 1998, pp. 145). The polygenic model using pedigree and phenotypic information, i.e., the animal model (Henderson 1984), has been the model of choice for estimating genetic parameters in breeding and natural populations (Abney et al. 2000; Sorensen and Gianola 2002; O′Hara et al. 2008).Recent breakthroughs in molecular techniques have made it possible to create genome-wide, single nucleotide polymorphism (SNP) maps. These maps have helped to uncover a vast amount of new loci responsible for trait expression and have provided general insights into the genetic architecture of quantitative traits (e.g., Valdar et al. 2006; Visscher 2008; Flint and Mackay 2009). These insights can help when calculating disease risks in humans, when attempting to increase the yield from breeding programs, and when estimating relatedness in conservation programs. High-density SNPs of many species of importance to science and agriculture can now be scored quickly and relatively cheaply, for example, in mice (Valdar et al. 2006), chickens (Muir et al. 2008), and dairy cattle (VanRaden et al. 2009).In the analysis of populations of breeding stock, the inclusion of dense marker data has improved the predictive ability (i.e., reliability) of genetic evaluations compared to the traditional phenotype model, both in simulations (Meuwissen et al. 2001; Calus et al. 2008; Hayes et al. 2009) and when using real data (Legarra et al. 2008; VanRaden et al. 2009; González-Recio et al. 2009). Meuwissen et al. (2001) suggested that the effect of all markers should first be estimated, and then summed, to obtain genomic estimated breeding values (GEBVs). An alternative procedure, where all markers are used to compute the genomic relationship matrix (in place of the additive polygenic relationship matrix) has also been suggested (e.g., Villanueva et al. 2005; VanRaden 2008; Hayes et al. 2009); this matrix is then incorporated into the statistical analysis to estimate GEBVs. A comparison of both procedures (VanRaden 2008) yielded similar estimates of GEBVs in cases where the effect of an individual allele was small. In addition, if not all pedigree members have marker information, a combined relationship matrix derived from both genotyped and ungenotyped individuals could be computed; this has been shown to increase the accuracy of GEBVs (Legarra et al. 2009; Misztal et al. 2009). Another plausible option to incorporate marker information is to use low-density SNP panels within families and to trace the effect of SNPs from high-density genotyped ancestors, as suggested by Habier et al. (2009) and Weigel et al. (2009). However, fast and powerful computer algorithms, which can use the marker information as efficiently as possible in the analysis of quantitative traits, are needed to obtain accurate GEBVs from genome-wide marker data.This study describes the development of an efficient Bayesian method for incorporating general relationships into the genetic evaluation procedure. The method is based on expressing the multivariate normal prior distribution as a product of one-dimensional normal distributions, each conditioned on the descending variables. When evaluating the genetic parameters of natural and breeding populations, high-dimensional distributions are often used as prior distributions of various genetic effects, such as the additive polygenic effect (Wang et al. 1993), multivariate additive polygenic effects (Van Tassell and Van Vleck 1996), and quantitative trait loci (QTL) effects via the identical-by-decent matrix (Yi and Xu 2000). A Bayesian framework is adopted to obtain posterior distributions of all unknown parameters, estimated by using Markov chain Monte Carlo (MCMC) sampling algorithms in the software package WinBUGS (Lunn et al. 2000, 2009). By performing prior calculations in the form of the factorized product of simple univariate conditional distributions, the computational time of the MCMC estimation procedure is reduced considerably. This feature permits rapid inference for both the polygenic model and the genomic relationship model. Moreover, the decomposition allows for inbreeding of varying degree, since the correct genetic covariance structure can be inferred into the analysis. In this article, we test the method on two previously published pedigree data sets: phenotype data from a large pedigree of Scots pine, incorporation of information on both additive and dominance genetic relationships (Waldmann et al. 2008); and genomic information obtained from a genome-wide scan of a simulated animal population (Lund et al. 2009).  相似文献   

11.
12.
Mlh1p forms three heterodimers that are important for mismatch repair (Mlh1p/Pms1p), crossing over during meiosis (Mlh1p/Mlh3p), and channeling crossover events into a specific pathway (Mlh1p/Mlh2p). All four proteins contain highly conserved ATPase domains and Pms1p has endonuclease activity. Studies of the functional requirements for Mlh1p/Pms1p in Saccharomyces cerevisae revealed an asymmetric contribution of the ATPase domains to repairing mismatches. Here we investigate the functional requirements of the Mlh1p and Mlh3p ATPase domains in meiosis by constructing separation of function mutations in Mlh3p. These mutations are analogous to mutations of Mlh1p that have been shown to lead to loss of ATP binding and/or ATP hydrolysis. Our data suggest that ATP binding by Mlh3p is required for meiotic crossing over while ATP hydrolysis is dispensable. This has been seen previously for Mlh1p. However, when mutations that affect ATP hydrolysis by both Mlh3p and Mlh1p are combined within a single cell, meiotic crossover frequencies are reduced. These observations suggest that the function of the Mlh1p/Mlh3p heterodimer requires both subunits to bind ATP but only one to efficiently hydrolyze it. Additionally, two different amino acid substitutions to the same residue (G97) in Mlh3p affect the minor mismatch repair function of Mlh3p while only one of them compromises its ability to promote crossing over. These studies thus reveal different functional requirements among the heterodimers formed by Mlh1p.CROSSING over during meiosis not only generates variation but is also important for providing the necessary interactions between homologous chromosomes that ensure correct segregation at division I of meiosis. Recombination is initiated by the production of programmed double-strand breaks (DSBs), catalyzed by the covalently attached Spo11p (Bergerat et al. 1997; Keeney et al. 1997), aided by a number of proteins (reviewed in Keeney and Neale 2006). DSBs are made at a much higher frequency than crossovers, and designation of only a subset to yield crossovers is thought to occur during early stages of DSB repair (Borner et al. 2004). At least two distinct pathways contribute to the production of crossover events in Saccharomyces cerevisiae. The major pathway is dependent on Msh4p/Msh5p and the mismatch repair proteins Mlh1p and Mlh3p (Ross-MacDonald and Roeder 1994; Hollingsworth et al. 1995; Hunter and Borts 1997; Wang et al. 1999; Abdullah et al. 2004) and the second pathway is dependent on Mus81p/Mms4p endonuclease (de los Santos et al. 2001, 2003).Mitotic mismatch repair (MMR) is the process by which mutations that arise during DNA replication and recombination are recognized and removed (reviewed in Kolodner 1996; Harfe and Jinks-Robertson 2000). Msh2p forms a heterodimer with Msh6p (MutSα) to repair base–base mismatches and small insertions and/or deletions and with Msh3p (MutSβ) to repair large insertions and/or deletions (reviewed in Jiricny 2006). Mlh1p forms heterodimers with Pms1p, Mlh2p, and Mlh3p to coordinate the removal of these mismatches (Prolla et al. 1994; Wang et al. 1999). Mlh1p/Pms1p (MutLα) are involved in the repair of all types of mismatches in combination with MutSα and MutSβ, and in the absence of either protein a mutator phenotype is observed (Habraken et al. 1997, 1998). Mlh1p/Mlh2p (MutLβ) and Mlh1p/Mlh3p (MutLγ) are involved in the MutSβ pathway only, which repairs frameshift mutations caused by insertions or deletions. Consequently mlh3Δ mutants only exhibit a weak mutator phenotype, due to a lesser involvement in mismatch repair and a partial overlap in function with Pms1p (Flores-Rozas and Kolodner 1998; Harfe et al. 2000).Although the MutL homologs interact primarily through their C-terminal domains (Pang et al. 1997; Ban and Yang 1998), it is thought that the N-terminal domains must also interact for the complex to be fully functional (Ban and Yang 1998). Binding of ATP causes the proteins to undergo conformational changes, which are essential for the interaction between the N termini (Ban et al. 1999; Tran and Liskay 2000; Sacho et al. 2008). ATP hydrolysis and subsequent release of ADP is required to allow the protein complex to return to its initial state, completing the cycle so that the subunits are ready to bind ATP again if required. Using mutants of MLH1 and PMS1 that are presumed to be defective for ATP binding and/or ATP hydrolysis, it has been shown that both of these functions are essential for fully effective mismatch repair (Tran and Liskay 2000). However, the ATP binding and ATP hydrolysis mutants of PMS1 exhibited lower mitotic mutation rates than the corresponding MLH1 ATPase mutants, suggesting that there is functional asymmetry within the Mlh1p/Pms1p heterodimer (Tran and Liskay 2000; Hall et al. 2002). Another example of the asymmetry in the contributions of these subunits to function can be seen in assays that measure recombination between diverged sequences (homeologous recombination). The Mlh1p ATPase activity has been shown to be more important for the suppression of homeologous recombination than Pms1p ATPase activity (Welz-Voegele et al. 2002). This functional asymmetry is supported by in vitro biochemical analysis that demonstrated Pms1p has a lower ATP binding affinity than Mlh1p (Hall et al. 2002).As mentioned above, Mlh1p/Mlh3p function in the Msh4p/Msh5p pathway for meiotic recombination (Hunter and Borts 1997; Santucci-Darmanin et al. 2000). The Msh4p/Msh5p complex is thought to act in the stabilization of Holliday junction intermediates to allow their resolution in a crossover configuration (Snowden et al. 2004). The Mlh1p/Mlh3p complex has been suggested to act in the resolution of these structures, either directly or indirectly. Human Pms2 and its yeast homolog, Pms1p, have been shown to possess a latent endonuclease activity, conferred by a motif that is conserved among some of the MutL homologs, including Mlh3p (Kadyrov et al. 2006, 2007). Mutations in the DHQA(X)2E(X)4E motif in yeast MLH3 cause defects in both mismatch repair and meiotic recombination equivalent to mlh3Δ, suggesting that Mlh3p may also possess an endonuclease activity that is important for the generation of crossovers (Nishant et al. 2008).ATP binding by Mlh1p has been shown to be important for both of its meiotic functions (crossing over and repair of heteroduplex DNA) (Pang et al. 1997; Tran and Liskay 2000; Hoffmann et al. 2003). In contrast, the ATP hydrolysis mutant mlh1-E31A/mlh1-E31A appears to have no effect on meiotic recombination (Tran and Liskay 2000; Hoffmann et al. 2003). This may partly be explained by in vitro studies demonstrating that this mutant exhibits a low level of ATPase activity (Hall et al. 2002).The meiotic functions of MLH1 can be functionally separated as shown by mutating the same residue, G98, to different amino acids (Hoffmann et al. 2003). The residue G98 is situated in the ATPase motif in the GFRGEAL box (GYRGDAL in Mlh3p), which forms the lid of the ATP binding pocket. Mutations in this motif are predicted to affect ATP binding and/or heterodimerization with Pms1p (Ban and Yang 1998; Ban et al. 1999). Mutating the residue G98 in the ATP binding lid to alanine resulted in defective repair of heteroduplex DNA while crossing over was unaffected, but when the same residue was mutated to valine both mismatch repair and crossover functions were defective (Hoffmann et al. 2003). The mlh1-G98V mutant disrupts the interaction of Mlh1p with Pms1p, while mlh1-G98A does not (Pang et al. 1997). This may contribute to the difference observed in the effect on crossing over as Mlh1p is thought to interact with Pms1p and Mlh3p through the same residues (Wang et al. 1999; Kondo et al. 2001). Consequently if the interaction with Pms1p is affected then it is likely that the interaction with Mlh3p is also disrupted.We constructed mlh3 mutants corresponding to the ATP binding and ATP hydrolysis mutants of mlh1 to explore the role of Mlh3p in meiotic recombination. We also constructed mlh3-G97A and mlh3-G97V mutants, equivalent to the mlh1-G98A/V pair that has been shown to differentially affect the mitotic and meiotic functions of Mlh1p. All mutants were assayed for mitotic mismatch repair, meiotic heteroduplex repair, crossing over, and chromosome segregation.  相似文献   

13.
During meiosis, recombination is directed to occur between homologous chromosomes to create connections necessary for proper segregation at meiosis I. Partner choice is determined at the time of strand invasion and is mediated by two recombinases: Rad51 and the meiosis-specific Dmc1. In budding yeast, interhomolog bias is created in part by the activity of a meiosis-specific kinase, Mek1, which is localized to the protein cores of condensed sister chromatids. Analysis of meiotic double-strand break (DSB) repair in haploid and disomic haploid strains reveals that Mek1 suppresses meiotic intersister DSB repair by working directly on sister chromatids. Rec8 cohesin complexes are not required, however, either for suppression of intersister DSB repair or for the repair itself. Regulation of DSB repair in meiosis is chromosome autonomous such that unrepaired breaks on haploid chromosomes do not prevent interhomolog repair between disomic homologs. The pattern of DSB repair in haploids containing Dmc1 and/or Rad51 indicates that Mek1 acts on Rad51-specific recombination processes.IN eukaryotes, meiosis is a specialized type of cell division that produces the gametes required for sexual reproduction. In meiosis, one round of DNA replication is followed by two rounds of chromosome segregation, termed meiosis I and II. As a result of the two divisions, four haploid cells are produced, each containing half the number of chromosomes as the diploid parent. Proper segregation at meiosis I requires connections between homologous chromosomes that are created by a combination of sister chromatid cohesion and recombination (Petronczki et al. 2003). In vegetative cells, cohesion is mediated by multisubunit ring-shaped complexes that are removed by proteolysis of the kleisin subunit, Mcd1/Scc1 (Onn et al. 2008). In meiotic cells, introduction of a meiosis-specific kleisin subunit, Rec8, allows for a two-step removal of cohesion with loss of arm cohesion at anaphase I and centromere cohesion at anaphase II (Klein et al. 1999). Missegregation of chromosomes during meiosis causes abnormal chromosome numbers in gametes that may lead to infertility and genetic disorders such as trisomy 21 or Down''s syndrome.In mitotically dividing budding yeast cells, recombination is mediated by an evolutionarily conserved RecA-like recombinase, Rad51, and occurs preferentially between sister chromatids (Kadyk and Hartwell 1992). In contrast, recombination during meiosis is initiated by the deliberate formation of double-strand breaks (DSBs) by an evolutionarily conserved, topoisomerase-like protein, Spo11, and occurs preferentially between homologous chromosomes (Jackson and Fink 1985; Schwacha and Kleckner 1997; Keeney 2001). After DSB formation, the 5′ ends on either side of the breaks are resected, resulting in 3′ single stranded (ss) tails. Rad51, and the meiosis-specific recombinase Dmc1, bind to the 3′ ssDNA tails to form protein/DNA filaments that promote strand invasion of homologous chromosomes. DNA synthesis and ligation result in the formation of double Holliday junctions, which are then preferentially resolved into crossovers (Allers and Lichten 2001; Hunter 2007).The precise roles that the Rad51 and Dmc1 recombinase activities play in meiotic recombination have been unclear because experiments have indicated both overlapping and distinct functions for the two proteins (Sheridan and Bishop 2006; Hunter 2007). While both rad51Δ and dmc1Δ mutants reduce interhomolog recombination, other studies suggest that Rad51, in complex with the accessory protein Rad54, is involved primarily in intersister DSB repair. In contrast, Dmc1, in conjunction with the accessory protein Rdh54/Tid1 (a paralog of Rad54), effects DSB repair in meiotic cells by invasion of nonsister chromatids (Dresser et al. 1997; Schwacha and Kleckner 1997; Shinohara et al. 1997a,b; Arbel et al. 1999; Bishop et al. 1999; Hayase et al. 2004; Sheridan and Bishop 2006).The preference for recombination to occur between homologous chromosomes during meiosis is created in part by Dmc1. DSBs accumulate in dmc1Δ diploids due to a failure in strand invasion (Bishop et al. 1992; Hunter and Kleckner 2001). In the efficiently sporulating SK1 strain background, these unrepaired breaks trigger the meiotic recombination checkpoint, resulting in prophase arrest (Lydall et al. 1996; Roeder and Bailis 2000). In dmc1Δ mutants, Rad51 is present at DSBs, yet there is no strand invasion of sister chromatids (Bishop 1994; Shinohara et al. 1997a). These results suggest that in addition to Dmc1 promoting interhomolog strand invasion, Rad51 activity must also be suppressed.Recent studies have shown that during meiosis Rad51 recombinase activity is inhibited by two different mechanisms that decrease the formation of Rad51/Rad54 complexes: (1) binding of the meiosis-specific Hed1 protein to Rad51, thereby excluding interaction with Rad54, and (2) reduction in the affinity of Rad54 for Rad51 due to phosphorylation of Rad54 by Mek1 (Tsubouchi and Roeder 2006; Busygina et al. 2008; Niu et al. 2009). Mek1 is a meiosis-specific kinase that is activated in response to DSBs (Niu et al. 2005, 2007; Carballo et al. 2008). In addition to phosphorylating Rad54, Mek1 phosphorylation of an as yet undetermined substrate is required to suppress Rad51/Rad54-mediated strand invasion of sister chromatids (Niu et al. 2009).To dissect the mechanism by which Mek1 suppresses meiotic intersister DSB repair, we took advantage of the ability of yeast cells to undergo haploid meiosis. The lack of homologous chromosomes in haploid cells makes it possible to examine sister-chromatid-specific events in the absence of interhomolog recombination. De Massy et al. (1994) previously observed a delay in DSB repair in haploid cells and proposed that this delay was due to a constraint in using sister chromatids. We have shown that this delay is dependent on MEK1 and utilized the haploid system to determine various biological parameters required to suppress meiotic intersister DSB repair. Our results indicate that Rad51 and Dmc1 recombinase activities have distinct roles during meiosis and that interhomolog bias is established specifically on sister chromatids through regulation of Rad51, not Dmc1. rec8Δ diploids exhibit defects in meiotic DSB repair (Klein et al. 1999; Brar et al. 2009). Given that cohesin complexes are specific for sister chromatids, we investigated the role of REC8 in intersister DSB repair and found it is required neither for suppressing intersister DSB repair during meiosis nor for the repair itself.  相似文献   

14.
15.
Regulation of cytoskeletal structure and dynamics is essential for multiple aspects of cellular behavior, yet there is much to learn about the molecular machinery underlying the coordination between the cytoskeleton and its effector systems. One group of proteins that regulate microtubule behavior and its interaction with other cellular components, such as actin-regulatory proteins and transport machinery, is the plus-end tracking proteins (MT+TIPs). In particular, evidence suggests that the MT+TIP, CLASP, may play a pivotal role in the coordination of microtubules with other cellular structures in multiple contexts, although the molecular mechanism by which it functions is still largely unknown. To gain deeper insight into the functional partners of CLASP, we conducted parallel genetic and proteome-wide screens for CLASP interactors in Drosophila melanogaster. We identified 36 genetic modifiers and 179 candidate physical interactors, including 13 that were identified in both data sets. Grouping interactors according to functional classifications revealed several categories, including cytoskeletal components, signaling proteins, and translation/RNA regulators. We focused our initial investigation on the MT+TIP Minispindles (Msps), identified among the cytoskeletal effectors in both genetic and proteomic screens. Here, we report that Msps is a strong modifier of CLASP and Abl in the retina. Moreover, we show that Msps functions during axon guidance and antagonizes both CLASP and Abl activity. Our data suggest a model in which CLASP and Msps converge in an antagonistic balance in the Abl signaling pathway.COORDINATION of cytoskeletal dynamics is an essential process for the regulation of virtually all aspects of cellular behavior including cell shape changes, cell division, and cell motility (e.g., Rodriguez et al. 2003; Kodama et al. 2004). Not only is the cytoskeletal system coordinated with numerous cellular pathways to control cell behavior, it also functions as a central organizing scaffold for multiple effector protein complexes downstream of signaling pathways and cellular processes such as intracellular transport. Yet, little is known regarding the molecular machinery that governs the integrated coordination of various cytoskeletal components. Evidence suggests that the microtubule (MT) plus-end tracking protein CLASP [cytoplasmic linker protein (CLIP)-associated protein], which has been implicated in mitotic spindle formation (Inoue et al. 2000, 2004) and in linking MT ends to other cell structures such as the cell cortex and kinetochore (Akhmanova et al. 2001; Maiato et al. 2003; Mimori-Kiyosue et al. 2005; Reis et al. 2009), may play a pivotal role in the overall coordination of cytoskeletal networks. Not only does it affect MT dynamics, but CLASP may function as an actin-MT crosslinker as well, as it possesses actin-binding activity (Tsvetkov et al. 2007) and CLASP-bound microtubules appear to track along F-actin bundles in growth cones (Lee et al. 2004).We previously showed that CLASP functions downstream of Abelson (Abl) nonreceptor tyrosine kinase (Lee et al. 2004), which is a key signaling molecule that modulates the cytoskeleton downstream of numerous cell surface receptor inputs and plays essential roles in various contexts including cell motility and human disease (Van Etten 1999; Moresco and Koleske 2003; Bradley and Koleske 2009). While most cytoskeletal-related studies of Abl have focused on its regulation of actin dynamics (Lanier and Gertler 2000; Wills et al. 2002; Hernandez et al. 2004; Bradley and Koleske 2009), few studies have examined its MT effectors such as CLASP and how they may be involved in the coordination of both cytoskeletal networks. Additionally, we previously reported that Drosophila CLASP is necessary for accurate embryonic axon guidance at the central nervous system (CNS) midline where conserved guidance factors (Netrins and Slits) control growth cone navigation (Lee et al. 2004). In embryonic axons, we found that CLASP is required for Abl function (Lee et al. 2004); however, additional partner proteins required for CLASP activity during axon guidance are largely unknown.Recent studies of CLASP function focusing on cell culture and imaging have identified several CLASP-binding proteins, such as additional MT+TIPs (CLIP-170 and EB1) (Akhmanova et al. 2001; Mimori-Kiyosue et al. 2005) and cell cortex-associated proteins (LL5beta and ELKS) (Lansbergen et al. 2006). While investigation of the detailed interactions and functional significance of these types of molecules and their relation to CLASP has been important to understanding the CLASP molecular mechanism, elucidating how CLASP functions in a broader context will require expanding our awareness of the entire CLASP network, or “interactome.” As of yet, there has not been a comprehensive unbiased survey of CLASP functional interactors.A long history of molecular pathway dissection in multiple model systems and biological contexts has shown that genetic and proteomic interactome screens are powerful tools for defining the network of functional partners for any given gene of interest (Xu et al. 1990; Simon et al. 1991; Carthew et al. 1994; Karim et al. 1996; Rebay et al. 2000; St Johnston 2002). Determining the CLASP interaction network can not only define MT+TIP-associated proteins and regulators of MT biology, but it can also reveal new classes of molecules that interact with CLASP. For example, a recent study suggested that CLASPs function as actin-MT crosslinkers because they possess actin-binding activity (Tsvetkov et al. 2007), but few specific actin-binding CLASP interactors have been identified, and the functional relevance of CLASP–actin interaction is still unclear. A systematic approach to define the CLASP interactome has the potential for significantly increasing our ability to understand the CLASP mechanism.Therefore, to expand our knowledge of the CLASP functional mechanism, we have performed a multilevel genetic and proteomic screen for CLASP interactors in Drosophila. The single Drosophila CLASP ortholog has been given multiple names [orbit/multiple asters (MAST)/chromosome bows (chb)] (Fedorova et al. 1997; Inoue et al. 2000; Lemos et al. 2000), but here, we refer to it as CLASP. Our screen has identified novel and specific partners in several functional categories including cytoskeletal components, signaling proteins, and the unanticipated class of translation/RNA regulators. To validate the findings of this screen, we focused our initial investigation on the conserved MT+TIP identified among the cytoskeletal effectors in both the genetic and proteomic screens, Minispindles (Msps, ortholog of the human CKAP5 [cytoskeleton associated protein 5)/TOG (tumor overexpressed gene)/Xenopus Xmap215)]. Msps function and regulation of MT stability has been studied previously in the context of the mitotic spindle (Cullen et al. 1999; Lee et al. 2001; Barros et al. 2005) and in centrosomes (Popov et al. 2002; Cassimeris and Morabito 2004), although it has not been shown to functionally interact with CLASP nor play any role in the nervous system.Here, we report that Msps is an in vivo antagonist of CLASP and interacts strongly with Abl. Furthermore, we show that Msps functions during axon guidance. Our data suggest a model in which CLASP and Msps act antagonistically to provide the growth cone with a rapidly adaptable output for Abl-dependent responses to attractive and repulsive guidance cues.  相似文献   

16.
17.
Qiuheng Lu  Jie Yan  Paul N. Adler 《Genetics》2010,185(2):549-558
The conserved frizzled (fz) pathway regulates planar cell polarity in both vertebrate and invertebrate animals. This pathway has been most intensively studied in the wing of Drosophila, where the proteins encoded by pathway genes all accumulate asymmetrically. Upstream members of the pathway accumulate on the proximal, distal, or both cell edges in the vicinity of the adherens junction. More downstream components including Inturned and Multiple Wing Hairs accumulate on the proximal side of wing cells prior to hair initiation. The Mwh protein differs from other members of the pathway in also accumulating in growing hairs. Here we show that the two Mwh accumulation patterns are under different genetic control with the early proximal accumulation being regulated by the fz pathway and the latter hair accumulation being largely independent of the pathway. We also establish recruitment by proximally localized Inturned to be a putative mechanism for the localization of Mwh to the proximal side of wing cells. Genetically inturned (in) acts upstream of mwh (mwh) and is required for the proximal localization of Mwh. We show that Mwh can bind to and co-immunoprecipitate with Inturned. We also show that these two proteins can function in close juxtaposition in vivo. An In∷Mwh fusion protein provided complete rescue activity for both in and mwh mutations. The fusion protein localized to the proximal side of wing cells prior to hair formation and in growing hairs as expected if protein localization is a key for the function of these proteins.THE frizzled (fz) signaling pathway regulates tissue planar cell polarity (PCP) in the epidermis of both vertebrate and invertebrate animals (Lawrence et al. 2007; Montcouquiol 2007; Wang and Nathans 2007; Zallen 2007). PCP is dramatic in the cuticle of insects such as Drosophila, which is decorated with arrays of hairs and sensory bristles.The genetic basis for tissue polarity has been most extensively studied in the fly wing (Wong and Adler 1993). The Planar Polarity (PCP) genes of the fz pathway (also known as the core PCP genes), the planar polarity effector (PPE) genes and the multiple wing hairs (mwh) gene encode key components that regulate planar polarity in the wing. fz, disheveled (dsh), prickle/spiny leg (pk/sple), Van Gogh (Vang) (aka strabismus), starry night (stan) (aka flamingo) and diego (dgo) are members of the PCP group (Vinson and Adler 1987; Wong and Adler 1993; Taylor et al. 1998; Wolff and Rubin 1998; Chae et al. 1999; Gubb et al. 1999; Usui et al. 1999). A distinctive feature of these genes is that their protein products accumulate asymmetrically on the distal (Fz, Dsh, and Dgo) (Axelrod 2001; Feiguin et al. 2001; Shimada et al. 2001; Strutt 2001), proximal (Vang, Pk)(Tree et al. 2002; Bastock et al. 2003), or both distal and proximal (Stan) (Usui et al. 1999) sides of wing cells. These genes/proteins act as a functional group and are corequirements for the asymmetric accumulation of the others.The PPE includes inturned (in), fuzzy (fy), and fritz (frtz) (Park et al. 1996; Collier and Gubb 1997; Collier et al. 2005). These genes are thought to function downstream of the PCP genes and the proteins encoded by these genes also accumulate asymmetrically in wing cells (Adler et al. 2004; Strutt and Warrington 2008). As is the case for the PCP genes, the PPE genes/proteins also appear to be a functional group and to be corequirements for the asymmetric accumulation of the others. Several observations support the hypothesis that the PPE genes are essential downstream effectors of the PCP genes. The earliest appreciation of this came from careful observations of the mutant phenotypes. A common feature of mutations in all of these genes is that they do not result in a randomization of hair polarity, but rather in a similar complicated and abnormal stereotypic pattern (Gubb and Garcia-Bellido 1982; Adler et al. 2000). That the abnormal patterns were so similar suggested that these genes all functioned in the same process (Wong and Adler 1993). The mutant phenotypes differed in that the vast majority of PCP mutant wing cells form a single hair, while many PPE mutant wing cells form two or three hairs. Mutations in PPE genes are epistatic to both loss- and gain-of-function mutations in PCP genes (Wong and Adler 1993; Lee and Adler 2002). Further evidence that the PPE genes function downstream of the PCP genes comes from the analysis of protein localization. PPE gene function is not needed for the proper asymmetric localization of PCP proteins (Usui et al. 1999; Strutt 2001; Tree et al. 2002; Collier et al. 2005) but in contrast PCP gene function is essential for the asymmetric accumulation of PPE proteins (Adler et al. 2004; Strutt and Warrington 2008). Further, the PCP genes/proteins instruct the localization of the PPE proteins (Adler et al. 2004).The multiple-wing-hairs (mwh) gene is thought to function downstream of both the PCP and PPE genes (Wong and Adler 1993). This conclusion comes from analyses that are similar to those that established that the PPE genes function downstream of the PCP genes. The overall hair polarity pattern of mwh mutant wings shares the same complicated and abnormal stereotypic hair polarity pattern seen in PCP and PPE mutants. However, mwh cells differ by producing a larger number of hairs (typically three to four hairs) (Wong and Adler 1993). mwh mutations are epistatic to mutations in both the PCP and PPE genes and mwh is not required for the asymmetric accumulation of either PCP or PPE proteins (Usui et al. 1999; Strutt 2001; Adler et al. 2004; Strutt and Warrington 2008).The mwh gene was recently determined to encode a novel G protein binding–formin homology 3 (GBD-FH3) protein with a complex accumulation pattern in wing cells (Strutt and Warrington 2008; Yan et al. 2008). Prior to hair initiation Mwh accumulates along the proximal side of wing cells and during hair growth Mwh accumulates in the growing hair. Temperature-shift experiments with a temperature-sensitive allele provided evidence for two temporally separate mwh functions and it was proposed that the two accumulation patterns were associated with the two temporal functions (Yan et al. 2008). Here we show that the early proximal accumulation of Mwh requires the function of the PCP and PPE genes (a result also seen previously in Strutt and Warrington 2008), while the hair accumulation of Mwh is largely independent of these two groups of genes providing further genetic evidence for Mwh having two independent functions.How does the Mwh protein accumulate proximally? An obvious possibility is that Mwh interacts directly with one or more of the upstream proteins and in this way is recruited to the proximal side. The PPE proteins are strong candidates to interact directly with Mwh, as they function genetically in between the PCP gene and Mwh (Wong and Adler 1993). Consistent with this possibility we found that In and Mwh interacted in the yeast two-hybrid system and that these two proteins co-immunoprecipated from wing cells. This interaction was found not to be dependent on the function of the PCP genes consistent with the data from genetic studies that both in and mwh retain at least partial function in a fz mutant wing (Wong and Adler 1993). The hypothesis that Mwh is recruited to the proximal side by interacting with In predicts that these two proteins function in close proximity to one another. Consistent with these expectations we found that an In∷Mwh fusion protein provided both In and Mwh function.  相似文献   

18.
19.
A major question about cytokinesis concerns the role of the septin proteins, which localize to the division site in all animal and fungal cells but are essential for cytokinesis only in some cell types. For example, in Schizosaccharomyces pombe, four septins localize to the division site, but deletion of the four genes produces only a modest delay in cell separation. To ask if the S. pombe septins function redundantly in cytokinesis, we conducted a synthetic-lethal screen in a septin-deficient strain and identified seven mutations. One mutation affects Cdc4, a myosin light chain that is an essential component of the cytokinetic actomyosin ring. Five others cause frequent cell lysis during cell separation and map to two loci. These mutations and their dosage suppressors define a signaling pathway (including Rho1 and a novel arrestin) for repairing cell-wall damage. The seventh mutation affects the poorly understood RNA-binding protein Scw1 and severely delays cell separation when combined either with a septin mutation or with a mutation affecting the septin-interacting, anillin-like protein Mid2, suggesting that Scw1 functions in a pathway parallel to that of the septins. Taken together, our results suggest that the S. pombe septins participate redundantly in one or more pathways that cooperate with the actomyosin ring during cytokinesis and that a septin defect causes septum defects that can be repaired effectively only when the cell-integrity pathway is intact.THE fission yeast Schizosaccharomyces pombe provides an outstanding model system for studies of cytokinesis (McCollum and Gould 2001; Balasubramanian et al. 2004; Pollard and Wu 2010). As in most animal cells, successful cytokinesis in S. pombe requires an actomyosin ring (AMR). The AMR begins to assemble at the G2/M transition and involves the type II myosin heavy chains Myo2 and Myp2 and the light chains Cdc4 and Rlc1 (Wu et al. 2003). Myo2 and Cdc4 are essential for cytokinesis under all known conditions, Rlc1 is important at all temperatures but essential only at low temperatures, and Myp2 is essential only under stress conditions. As the AMR constricts, a septum of cell wall is formed between the daughter cells. The primary septum is sandwiched by secondary septa and subsequently digested to allow cell separation (Humbel et al. 2001; Sipiczki 2007). Because of the internal turgor pressure of the cells, the proper assembly and structural integrity of the septal layers are essential for cell survival.Septum formation involves the β-glucan synthases Bgs1/Cps1/Drc1, Bgs3, and Bgs4 (Ishiguro et al. 1997; Le Goff et al. 1999; Liu et al. 1999, 2002; Martín et al. 2003; Cortés et al. 2005) and the α-glucan synthase Ags1/Mok1 (Hochstenbach et al. 1998; Katayama et al. 1999). These synthases are regulated by the Rho GTPases Rho1 and Rho2 and the protein kinase C isoforms Pck1 and Pck2 (Arellano et al. 1996, 1997, 1999; Nakano et al. 1997; Hirata et al. 1998; Calonge et al. 2000; Sayers et al. 2000; Ma et al. 2006; Barba et al. 2008; García et al. 2009b). The Rho GTPases themselves appear to be regulated by both GTPase-activating proteins (GAPs) and guanine-nucleotide-exchange factors (GEFs) (Nakano et al. 2001; Calonge et al. 2003; Iwaki et al. 2003; Tajadura et al. 2004; Morrell-Falvey et al. 2005; Mutoh et al. 2005; García et al. 2006, 2009a,b). In addition, septum formation and AMR function appear to be interdependent. In the absence of a normal AMR, cells form aberrant septa and/or deposit septal materials at random locations, whereas a mutant defective in septum formation (bgs1) is also defective in AMR constriction (Gould and Simanis 1997; Le Goff et al. 1999; Liu et al. 1999, 2000). Both AMR constriction and septum formation also depend on the septation initiation network involving the small GTPase Spg1 (McCollum and Gould 2001; Krapp and Simanis 2008). Despite this considerable progress, many questions remain about the mechanisms and regulation of septum formation and its relationships to the function of the AMR.One major question concerns the role(s) of the septins. Proteins of this family are ubiquitous in fungal and animal cells and typically localize to the cell cortex, where they appear to serve as scaffolds and diffusion barriers for other proteins that participate in a wide variety of cellular processes (Longtine et al. 1996; Gladfelter et al. 2001; Hall et al. 2008; Caudron and Barral 2009). Despite the recent progress in elucidating the mechanisms of septin assembly (John et al. 2007; Sirajuddin et al. 2007; Bertin et al. 2008; McMurray and Thorner 2008), the details of septin function remain obscure. However, one prominent role of the septins and associated proteins is in cytokinesis. Septins concentrate at the division site in every cell type that has been examined, and in Saccharomyces cerevisiae (Hartwell 1971; Longtine et al. 1996; Lippincott et al. 2001; Dobbelaere and Barral 2004) and at least some Drosophila (Neufeld and Rubin 1994; Adam et al. 2000) and mammalian (Kinoshita et al. 1997; Surka et al. 2002) cell types, the septins are essential for cytokinesis. In S. cerevisiae, the septins are required for formation of the AMR (Bi et al. 1998; Lippincott and Li 1998). However, this cannot be their only role, because the AMR itself is not essential for cytokinesis in this organism (Bi et al. 1998; Korinek et al. 2000; Schmidt et al. 2002). Moreover, there is no evidence that the septins are necessary for AMR formation or function in any other organism. A further complication is that in some cell types, including most Caenorhabditis elegans cells (Nguyen et al. 2000; Maddox et al. 2007) and some Drosophila cells (Adam et al. 2000; Field et al. 2008), the septins do not appear to be essential for cytokinesis even though they localize to the division site.S. pombe has seven septins, four of which (Spn1, Spn2, Spn3, and Spn4) are expressed in vegetative cells and localize to the division site shortly before AMR constriction and septum formation (Longtine et al. 1996; Berlin et al. 2003; Tasto et al. 2003; Wu et al. 2003; An et al. 2004; Petit et al. 2005; Pan et al. 2007; Onishi et al. 2010). Spn1 and Spn4 appear to be the core members of the septin complex (An et al. 2004; McMurray and Thorner 2008), and mutants lacking either of these proteins do not assemble the others at the division site. Assembly of a normal septin ring also depends on the anillin-like protein Mid2, which colocalizes with the septins (Berlin et al. 2003; Tasto et al. 2003). Surprisingly, mutants lacking the septins are viable and form seemingly complete septa with approximately normal timing. These mutants do, however, display a variable delay in separation of the daughter cells, suggesting that the septins play some role(s) in the proper completion of the septum or in subsequent processes necessary for cell separation (Longtine et al. 1996; An et al. 2004; Martín-Cuadrado et al. 2005).It is possible that the septins localize to the division site and yet are nonessential for division in some cell types because their role is redundant with that of some other protein(s) or pathway(s). To explore this possibility in S. pombe, we screened for mutations that were lethal in combination with a lack of septins. The results suggest that the septins cooperate with the AMR during cytokinesis and that, in the absence of septin function, the septum is not formed properly, so that an intact system for recognizing and repairing cell-wall damage becomes critical for cell survival.  相似文献   

20.
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