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 390 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.
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   

7.
Codon usage bias is the nonrandom use of synonymous codons for the same amino acid. Most population genetic models of codon usage evolution assume that the population is at mutation–selection–drift equilibrium. Natural populations, however, frequently deviate from equilibrium, often because of recent demographic changes. Here, we construct a matrix model that includes the effects of a recent change in population size on estimates of selection on preferred vs. unpreferred codons. Our results suggest that patterns of synonymous polymorphisms affecting codon usage can be quite erratic after such a change; statistical methods that fail to take demographic effects into account can then give incorrect estimates of important parameters. We propose a new method that can accurately estimate both demographic and codon usage parameters. The method also provides a simple way of testing for the effects of covariates such as gene length and level of gene expression on the intensity of selection, which we apply to a large Drosophila melanogaster polymorphism data set. Our analyses of twofold degenerate codons reveal that (i) selection acts in favor of preferred codons, (ii) there is mutational bias in favor of unpreferred codons, (iii) shorter genes and genes with higher expression levels are under stronger selection, and (iv) there is little evidence for a recent change in population size in the Zimbabwe population of D. melanogaster.CODONS specifying the same amino acid are called synonymous codons. These are often used nonrandomly, with some codons appearing more frequently than others. This biased usage of synonymous codons has been found in many organisms such as Drosophila, yeast, and bacteria (Ikemura 1985; Duret and Mouchiroud 1999; Hershberg and Petrov 2008). Conventionally, synonymous codons for a given amino acid are divided into two classes: preferred and unpreferred codons (Ikemura 1985; Akashi 1994; Duret and Mouchiroud 1999). Several observations indicate that codon usage is affected by natural selection. First, in species with codon usage bias, preferred codons generally correspond to the most abundant tRNA species (Ikemura 1981). Second, highly expressed genes usually have higher codon usage bias than genes with low expression (Sharp and Li 1986; Duret and Mouchiroud 1999; Hey and Kliman 2002). Third, the synonymous substitution rate of a gene has been shown to be negatively correlated with its degree of codon usage bias (Sharp and Li 1986; Bierne and Eyre-Walker 2006). The most commonly cited explanations of the apparent fitness differences between preferred and unpreferred codons are selection for translation efficiency, translational accuracy, and mRNA stability (Ikemura 1985; Eyre-Walker and Bulmer 1993; Akashi 1994; Drummond et al. 2005). Recently, it has been proposed that exon splicing also affects codon usage bias (Warnecke and Hurst 2007).From a population genetics perspective, the extent of codon usage bias is ultimately a product of the joint effects of mutation, selection, genetic drift, recombination, and demographic history. The Li–Bulmer model of drift, selection, and reversible mutation between preferred and unpreferred codons at a site is the most widely used model (Li 1987; Bulmer 1991; McVean and Charlesworth 1999). Applications of this model generally assume that the population is at mutation–selection–drift equilibrium. However, empirical studies have suggested that changes in the strengths of various driving forces may not be unusual. For example, in Drosophila melanogaster, there is evidence that the population size (Li and Stephan 2006; Thornton and Andolfatto 2006; Keightley and Eyre-Walker 2007; Stephan and Li 2007), recombinational landscape (Takano-Shimizu 1999), and mutational process (Takano-Shimizu 2001; Kern and Begun 2005) may have changed significantly over the species'' evolutionary history.Such changes cause departures from equilibrium. Theoretical models show that it takes a very long time, proportional to the reciprocal of the mutation rate, for the population to approach a new equilibrium state (Tachida 2000; Comeron and Kreitman 2002). Before reaching equilibrium, the population often shows counterintuitive patterns of evolution (Eyre-Walker 1997; Takano-Shimizu 1999, 2001; Comeron and Kreitman 2002; Comeron and Guthrie 2005; Charlesworth and Eyre-Walker 2007). Despite these theoretical results, details of the patterns of polymorphism and substitution rates following a recent change in population size, and their effects on estimates of strength of selection, have not been determined.The above findings point to the importance of incorporating nonequilibrium factors into the study of codon usage bias. To this end, we extend the Li–Bulmer model to allow population size to vary over time, by representing the evolutionary process by a transition matrix. By analyzing this matrix model, we show that a recent change in population size can result in erratic patterns of codon usage and that methods failing to take into account these demographic effects can give false estimates of the intensity of selection.To solve these problems, we propose a new method, which does not require polarizing ancestral vs. derived states using outgroup data (cf. Cutter and Charlesworth 2006), but requires only knowledge of preferred vs. unpreferred states defined by patterns of codon usage. We use information on both polymorphic and fixed sites, which enables both mutational bias and the strength of selection to be estimated, in contrast to previous methods that use information on polymorphisms alone. Simulations indicate that this method can accurately estimate both demographic and codon usage parameters and can distinguish between selection and demography. We use the new method to analyze a large D. melanogaster polymorphism data set (Shapiro et al. 2007) and find evidence for natural selection on synonymous codons. We use our approach to show that genes with shorter coding sequences and higher levels of expression are under significantly stronger selection than longer genes with lower expression.  相似文献   

8.
9.
10.
11.
The vertebrate tight junction is a critical claudin-based cell–cell junction that functions to prevent free paracellular diffusion between epithelial cells. In Drosophila, this barrier is provided by the septate junction, which, despite being ultrastructurally distinct from the vertebrate tight junction, also contains the claudin-family proteins Megatrachea and Sinuous. Here we identify a third Drosophila claudin, Kune-kune, that localizes to septate junctions and is required for junction organization and paracellular barrier function, but not for apical-basal polarity. In the tracheal system, septate junctions have a barrier-independent function that promotes lumenal secretion of Vermiform and Serpentine, extracellular matrix modifier proteins that are required to restrict tube length. As with Sinuous and Megatrachea, loss of Kune-kune prevents this secretion and results in overly elongated tubes. Embryos lacking all three characterized claudins have tracheal phenotypes similar to any single mutant, indicating that these claudins act in the same pathway controlling tracheal tube length. However, we find that there are distinct requirements for these claudins in epithelial septate junction formation. Megatrachea is predominantly required for correct localization of septate junction components, while Sinuous is predominantly required for maintaining normal levels of septate junction proteins. Kune-kune is required for both localization and levels. Double- and triple-mutant combinations of Sinuous and Megatrachea with Kune-kune resemble the Kune-kune single mutant, suggesting that Kune-kune has a more central role in septate junction formation than either Sinuous or Megatrachea.EPITHELIA are essential for separating physiologically distinct body compartments and regulating trafficking between them. For proper function, it is imperative that epithelia maintain effective barriers against free paracellular diffusion. To this end, epithelial cells contain occluding junctions, which regulate paracellular permeability. In vertebrates, this is accomplished by tight junctions (TJ), structures that are characterized by regions of close membrane apposition between adjacent cells known as “kissing points” (Tsukita and Furuse 2002). While the TJ is made up of at least 40 different components (Schneeberger and Lynch 2004), the core proteins responsible for the paracellular barrier are the claudins (Angelow et al. 2008).Claudins are four-transmembrane domain proteins that form homo- and heterophilic interactions within the same cell (Furuse et al. 1999; Blasig et al. 2006) and with claudins in adjacent cells (Furuse et al. 1999), thereby establishing the paracellular seal. There are 24 members of the claudin family in mammals, many of which display distinct, tissue-specific expression patterns (Kiuchi-Saishin et al. 2002; Angelow et al. 2008). Mutations in several claudins can cause significant paracellular permeability defects in mice. For example, mutations in claudin-14 increase TJ permeability in the organ of Corti and cause deafness (Ben-Yosef et al. 2003), while loss of claudin-1 compromises epidermal barrier function (Furuse et al. 2002).In Drosophila, primary (ectodermally derived) epithelia lack discernable TJs and instead use pleated septate junctions (SJ) for the paracellular barrier (Baumgartner et al. 1996; Lamb et al. 1998; Genova and Fehon 2003; Paul et al. 2003). However, despite sharing a common barrier function, vertebrate TJs and invertebrate SJs differ in several ways. While vertebrate TJs are positioned apical to adherens junctions (AJ) and contain conserved apical polarity proteins, SJs are basal to AJs and contain conserved basolateral polarity proteins (reviewed in Tepass 2003; Wu and Beitel 2004). In addition, SJs do not contain kissing points, but rather ladder-like septa that span the intermembrane space (Lane and Swales 1982; Tepass and Hartenstein 1994).Beyond their general epithelial barrier function, SJs are also required for several tissue-specific processes. Glial cells, for example, ensheath nerve fibers and use SJs to maintain the blood–brain barrier (Auld et al. 1995; Baumgartner et al. 1996; Schwabe et al. 2005). In the embryonic tracheal system, SJs are required for the apical secretion of the lumenal matrix modifying proteins, Vermiform (Verm) and Serpentine (Serp), which act through undefined pathways to restrict tube length (Wang et al. 2006). This secretory pathway appears to be specific for Verm and Serp, since other apical proteins are secreted normally in SJ mutants. SJ proteins have also been shown to play a role in morphogenesis of the heart tube, even though this tissue lacks typical SJ septa (Yi et al. 2008).Although SJs have clear differences from vertebrate TJs, SJs contain at least two claudins, Megatrachea (Mega) and Sinuous (Sinu), both of which are required for the paracellular barrier (Behr et al. 2003; Wu et al. 2004; Stork et al. 2008). In this article, we identify a third claudin, Kune-kune (Kune), that is an integral SJ protein. Like the other claudins, Kune is required for maintaining epithelial paracellular barrier and tracheal tube size control and is not required for apical-basal polarity. We also find that, of all three characterized claudins, Kune has a more severe SJ phenotype, suggesting that it is a more central player in SJ organization and function than previously characterized Drosophila claudins.  相似文献   

12.
13.
14.
Variation in maize for response to photoperiod is related to geographical adaptation in the species. Maize possesses homologs of many genes identified as regulators of flowering time in other species, but their relation to the natural variation for photoperiod response in maize is unknown. Candidate gene sequences were mapped in four populations created by crossing two temperate inbred lines to two photoperiod-sensitive tropical inbreds. Whole-genome scans were conducted by high-density genotyping of the populations, which were phenotyped over 3 years in both short- and long-day environments. Joint multiple population analysis identified genomic regions controlling photoperiod responses in flowering time, plant height, and total leaf number. Four key genome regions controlling photoperiod response across populations were identified, referred to as ZmPR1–4. Functional allelic differences within these regions among phenotypically similar founders suggest distinct evolutionary trajectories for photoperiod adaptation in maize. These regions encompass candidate genes CCA/LHY, CONZ1, CRY2, ELF4, GHD7, VGT1, HY1/SE5, TOC1/PRR7/PPD-1, PIF3, ZCN8, and ZCN19.MAIZE (Zea mays L. subsp. mays) was domesticated in southern Mexico and its center of diversity is in tropical Latin America (Goodman 1999; Matsuoka et al. 2002), where precipitation rates and day lengths cycle annually. The presumed ancestor of maize, teosinte (Zea mays L. subsp. parviglumis), likely evolved photoperiod sensitivity to synchronize its reproductive phases to the wetter, short-day growing season (Ribaut et al. 1996; Campos et al. 2006). A critical event in the postdomestication evolution of maize was its spread from tropical to temperate regions of the Americas (Goodman 1988), requiring adaptation to longer day lengths. The result of this adaptation process is manifested today as a major genetic differentiation between temperate and tropical maize (Liu et al. 2003) and substantially reduced photoperiod sensitivity of temperate maize (Gouesnard et al. 2002). Tropical maize exhibits delayed flowering time, increased plant height, and a greater total leaf number when grown in temperate latitudes with daily dark periods <11 hr (Allison and Daynard 1979; Warrington and Kanemasu 1983a,b). Identifying the genes underlying maize photoperiod sensitivity will provide insight into the postdomestication evolution of maize and may reduce barriers to the use of diverse tropical germplasm resources for improving temperate maize production (Holland and Goodman 1995; Liu et al. 2003; Ducrocq et al. 2009).Natural variation at key genes in flowering time pathways is related to adaptation and evolution of diverse plant species (Caicedo et al. 2004; Shindo et al. 2005; Turner et al. 2005; Cockram et al. 2007; Izawa 2007; Slotte et al. 2007). Identification of some of the genes controlling adaptation in numerous plant species relied on regulatory pathways elucidated in Arabidopsis (Simpson and Dean 2002). Many key genes in the Arabidopsis flowering time regulatory pathways are conserved across diverse plant species (Kojima et al. 2002; Hecht et al. 2007; Kwak et al. 2008), but their functions have diverged, resulting in unique regulatory pathways in some phylogenetic groups (Colasanti and Coneva 2009). For example, FRI and FLC control most natural variation for vernalization response in Arabidopsis (Caicedo et al. 2004; Shindo et al. 2005), but wheat and barley appear to lack homologs of these genes and regulate vernalization response with different genes (Yan et al. 2004).Maize exhibits tremendous natural variation for flowering time (Gouesnard et al. 2002; Camus-Kulandaivelu et al. 2006), for which numerous QTL have been identified (Chardon et al. 2004). In contrast, only a few flowering time mutants are known and only a handful of flowering time genes, including DWARF8 (D8), DELAYED FLOWERING1 (DLF1), VEGETATIVE TO GENERATIVE TRANSITION1 (VGT1), and INDETERMINATE GROWTH1 (ID1), have been cloned in maize (Thornsberry et al. 2001; Colasanti et al. 2006; Muszynski et al. 2006; Salvi et al. 2007; Colasanti and Coneva 2009). Variation at or near D8 and VGT1 is related to latitudinal adaptation, but these genes do not appear to regulate photoperiod responses and account for only a limited proportion of the standing flowering time variation in maize (Camus-Kulandaivelu et al. 2006, 2008; Ducrocq et al. 2008; Buckler et al. 2009).Quantitative trait loci (QTL) mapping was a key first step to identifying the genes underlying natural variation for flowering time in Arabidopsis (Koornneef et al. 2004). Photoperiodic QTL have been mapped previously in individual biparental maize mapping populations (Koester et al. 1993; Moutiq et al. 2002; Wang et al. 2008; Ducrocq et al. 2009). Such studies are informative with respect to the parents from which the populations were derived, but often do not reflect the genetic heterogeneity of broader genetic reference populations (Holland 2007).Association mapping (Thornsberry et al. 2001; Ersoz et al. 2007) and combined analysis of multiple biparental crosses (Rebaï et al. 1997; Rebaï and Goffinet 2000; Blanc et al. 2006; Verhoeven et al. 2006; Yu et al. 2008) represent alternative approaches to understanding the variation in genetic control for complex traits among diverse germplasm. Association mapping has limited power to identify genes that affect traits closely associated with population structure, such as flowering time in maize (Camus-Kulandaivelu et al. 2006; Ersoz et al. 2007). In contrast, joint QTL analysis of multiple populations is not hindered by the associations between causal genes and population structure. Combined QTL analysis of multiple mapping populations provides improved power to detect QTL, more precise estimation of their effects and positions, and better understanding of their functional allelic variation and distribution across more diverse germplasm compared to single-population mapping (Rebaï et al. 1997; Wu and Jannink 2004; Jourjon et al. 2005; Blanc et al. 2006; Verhoeven et al. 2006; Yu et al. 2008; Buckler et al. 2009). Joint analysis also provides a direct test of the importance of higher-order epistatic interactions between founder alleles at individual loci with genetic backgrounds (Jannink and Jansen 2001; Blanc et al. 2006). In this study, joint analysis of multiple populations was used to test directly the hypothesis that diverse tropical maize lines carry functionally similar alleles at key photoperiod loci, which would imply genetic homogeneity for a common set of mutations and a shared evolutionary pathway for photoperiod insensitivity.The objective of this study was to integrate candidate gene analyses with photoperiod QTL mapping across multiple maize populations. We tested candidate floral regulators known from other species for associations with natural variation for photoperiod response in maize. We analyzed flowering time in four interrelated recombinant inbred line (RIL) populations, each derived from crosses between temperate and tropical maize parents (Figure 1), in both long- and short-day environments to characterize their responses to distinct photoperiods. Joint population analysis provided high resolution of many QTL positions, permitting robust testing of underlying candidate genes. We directly and indirectly mapped homologs of flowering time candidates genes from Arabidopsis, rice, and barley on a dense consensus genetic map of these four populations, permitting identification of homologs that colocalize with genome regions associated with variation for photoperiod response. These mapping families are being integrated into the maize nested association mapping (NAM) population (Buckler et al. 2009; McMullen et al. 2009) because they were genotyped with the maize NAM map SNP markers, they involve the common parent B73, and their seed and genotypic information (File S1 cont.) are publicly available. Their availability further expands the genetic diversity represented by the maize NAM population and enhances this valuable public community resource.Open in a separate windowFigure 1.—Factorial mating of two temperate (B73 and B97) and two tropical (CML254 and Ki14) inbred maize lines to create four related recombinant inbred line mapping populations.  相似文献   

15.
16.
17.
18.
19.
Autophagy is an evolutionarily conserved degradative pathway that has been implicated in a number of physiological events important for human health. This process was originally identified as a response to nutrient deprivation and is thought to serve in a recycling capacity during periods of nutritional stress. Autophagy activity appears to be highly regulated and multiple signaling pathways are known to target a complex of proteins that contains the Atg1 protein kinase. The data here extend these observations and identify a particular phosphorylation event on Atg1 as a potential control point within the autophagy pathway in Saccharomyces cerevisiae. This phosphorylation occurs at a threonine residue, T226, within the Atg1 activation loop that is conserved in all Atg1 orthologs. Replacing this threonine with a nonphosphorylatable residue resulted in a loss of Atg1 protein kinase activity and a failure to induce autophagy. This phosphorylation required the presence of a functional Atg1 kinase domain and two known regulators of Atg1 activity, Atg13 and Atg17. Interestingly, the levels of this modification were found to increase dramatically upon exposure to conditions that induce autophagy. In addition, T226 phosphorylation was associated with an autophosphorylated form of Atg1 that was found specifically in cells undergoing the autophagy process. In all, these data suggest that autophosphorylation within the Atg1 activation loop may represent a point of regulatory control for this degradative process.MACROAUTOPHAGY (hereafter referred to as autophagy) is a highly conserved process of self-degradation that is essential for cell survival during periods of nutrient limitation (Tsukada and Ohsumi 1993). During autophagy, a double membrane grows out from a specific nucleation site, known as the pre-autophagosomal structure, or PAS, in Saccharomyces cerevisiae and the phagophore assembly site in mammals (Suzuki and Ohsumi 2007). This membrane encapsulates bulk protein and other constituents of the cytoplasm and ultimately targets this material to the vacuole/lysosome for degradation (Xie and Klionsky 2007). Recent studies have linked this pathway to a number of processes important for human health, including tumor suppression, innate immunity, and neurological disorders, like Huntington''s disease (Rubinsztein et al. 2007; Levine and Kroemer 2008). Determining how this pathway is regulated is therefore important for our understanding of these processes and our attempts to manipulate autophagy in clinically beneficial ways.Most of the molecular components of the autophagy pathway were initially characterized in the budding yeast, S. cerevisiae, but orthologs of many of these Atg proteins have since been found in other eukaryotes (Tsukada and Ohsumi 1993; Meijer et al. 2007). A complex of proteins that contains the Atg1 protein kinase is of special interest and appears to be a key point of regulatory control within this pathway (Kamada et al. 2000; Budovskaya et al. 2005; He and Klionsky 2009; Stephan et al. 2009). In S. cerevisiae, genetic and biochemical data indicate that this complex is targeted by at least three different signaling pathways. Two of these pathways, involving the Tor and cAMP-dependent protein kinases, inhibit this process, whereas the AMP-activated protein kinase is needed for the full induction of autophagy (Noda and Ohsumi 1998; Wang et al. 2001; Budovskaya et al. 2004; Stephan and Herman 2006; Kamada et al. 2010). The manner in which these signaling pathways regulate Atg1 activity and the precise role of this kinase in the autophagy process are presently matters of intense scrutiny.Although Atg1 kinase activity is required for the induction of autophagy, relatively little is known about how this enzyme is regulated in vivo. Two proteins associated with Atg1, Atg13 and Atg17, have been shown to be required for full Atg1 kinase activity both in vitro and in vivo (Kamada et al. 2000; Stephan et al. 2009). The roles of these proteins appear to be conserved through evolution as functional homologs of both have been identified in fruit flies and/or mammals (Hara et al. 2008; Chan et al. 2009; Chang and Neufeld 2009; Ganley et al. 2009; Hosokawa et al. 2009; Jung et al. 2009; Mercer et al. 2009). However, it is not yet clear precisely how these proteins stimulate Atg1 activity. In this study, we show that Atg1 is autophosphorylated within the activation loop and that this phosphorylation is required for both Atg1 kinase activity and the induction of autophagy. The activation loop is a structurally conserved element within the kinase domain and phosphorylation within this loop is often a necessary prerequisite for efficient substrate binding and/or phosphotransfer in the catalytic site (Johnson et al. 1996; Nolen et al. 2004). This loop generally corresponds to the sequence between two signature elements within the core kinase domain, the DFG and APE motifs (Hanks and Hunter 1995). Phosphorylation within this loop tends to result in a more ordered structure for this region and the proper positioning of key elements within the catalytic core of the kinase domain (Knighton et al. 1991; Johnson and O''reilly 1996; Huse and Kuriyan 2002). We found that Atg1 activation loop phosphorylation was correlated with the onset of autophagy and that replacing the site of phosphorylation with a phosphomimetic residue led to constitutive Atg1 autophosphorylation in vivo. In all, the data here suggest that Atg1 phosphorylation within its activation loop may be an important point of regulation within the autophagy pathway and models that discuss these data are presented.  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