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 15 毫秒
1.
Targeted gene silencing by RNA interference allows the study of gene function in plants and animals. In cell culture and small animal models, genetic screens can be performed—even tissue-specifically in Drosophila—with genome-wide RNAi libraries. However, a major problem with the use of RNAi approaches is the unavoidable false-positive error caused by off-target effects. Until now, this is minimized by computational RNAi design, comparing RNAi to the mutant phenotype if known, and rescue with a presumed ortholog. The ultimate proof of specificity would be to restore expression of the same gene product in vivo. Here, we present a simple and efficient method to rescue the RNAi-mediated knockdown of two independent genes in Drosophila. By exploiting the degenerate genetic code, we generated Drosophila RNAi Escape Strategy Construct (RESC) rescue proteins containing frequent silent mismatches in the complete RNAi target sequence. RESC products were no longer efficiently silenced by RNAi in cell culture and in vivo. As a proof of principle, we rescue the RNAi-induced loss of function phenotype of the eye color gene white and tracheal defects caused by the knockdown of the heparan sulfate proteoglycan syndecan. Our data suggest that RESC is widely applicable to rescue and validate ubiquitous or tissue-specific RNAi and to perform protein structure–function analysis.  相似文献   

2.
3.
Systematic perturbation screens provide comprehensive resources for the elucidation of cancer driver genes. The perturbation of many genes in relatively few cell lines in such functional screens necessitates the development of specialized computational tools with sufficient statistical power. Here we developed APSiC (Analysis of Perturbation Screens for identifying novel Cancer genes) to identify genetic drivers and effectors in perturbation screens even with few samples. Applying APSiC to the shRNA screen Project DRIVE, APSiC identified well-known and novel putative mutational and amplified cancer genes across all cancer types and in specific cancer types. Additionally, APSiC discovered tumor-promoting and tumor-suppressive effectors, respectively, for individual cancer types, including genes involved in cell cycle control, Wnt/β-catenin and hippo signalling pathways. We functionally demonstrated that LRRC4B, a putative novel tumor-suppressive effector, suppresses proliferation by delaying cell cycle and modulates apoptosis in breast cancer. We demonstrate APSiC is a robust statistical framework for discovery of novel cancer genes through analysis of large-scale perturbation screens. The analysis of DRIVE using APSiC is provided as a web portal and represents a valuable resource for the discovery of novel cancer genes.  相似文献   

4.
5.
Predicting gene functions by integrating large-scale biological data remains a challenge for systems biology. Here we present a resource for Drosophila melanogaster gene function predictions. We trained function-specific classifiers to optimize the influence of different biological datasets for each functional category. Our model predicted GO terms and KEGG pathway memberships for Drosophila melanogaster genes with high accuracy, as affirmed by cross-validation, supporting literature evidence, and large-scale RNAi screens. The resulting resource of prioritized associations between Drosophila genes and their potential functions offers a guide for experimental investigations.  相似文献   

6.
7.
8.
Interchromosomal duplications are especially important for the study of X-linked genes. Males inheriting a mutation in a vital X-linked gene cannot survive unless there is a wild-type copy of the gene duplicated elsewhere in the genome. Rescuing the lethality of an X-linked mutation with a duplication allows the mutation to be used experimentally in complementation tests and other genetic crosses and it maps the mutated gene to a defined chromosomal region. Duplications can also be used to screen for dosage-dependent enhancers and suppressors of mutant phenotypes as a way to identify genes involved in the same biological process. We describe an ongoing project in Drosophila melanogaster to generate comprehensive coverage and extensive breakpoint subdivision of the X chromosome with megabase-scale X segments borne on Y chromosomes. The in vivo method involves the creation of X inversions on attached-XY chromosomes by FLP-FRT site-specific recombination technology followed by irradiation to induce large internal X deletions. The resulting chromosomes consist of the X tip, a medial X segment placed near the tip by an inversion, and a full Y. A nested set of medial duplicated segments is derived from each inversion precursor. We have constructed a set of inversions on attached-XY chromosomes that enable us to isolate nested duplicated segments from all X regions. To date, our screens have provided a minimum of 78% X coverage with duplication breakpoints spaced a median of nine genes apart. These duplication chromosomes will be valuable resources for rescuing and mapping X-linked mutations and identifying dosage-dependent modifiers of mutant phenotypes.MANY eukaryotes of biomedical and agricultural importance—including humans, other mammals, birds, and Drosophila—are heterogametic. Their sex chromosomes differ drastically in size and genetic composition. In species with X and Y chromosomes, males carry only one copy of each X-linked gene. This poses a serious challenge for experimental geneticists, because males inheriting a mutation in a vital X-linked gene die before they can be used in genetic crosses. In fact, the hemizygosity of X-linked genes in males has been a significant barrier to the functional analysis of many X-linked genes and is largely responsible for the poor genetic characterization of X chromosomes relative to autosomes in most organisms.The lethality of X-linked mutations can be rescued by providing a wild-type copy of the mutated gene elsewhere in the genome. This can be accomplished with a transgenic construct if the molecular identity of the mutated gene is known. In many cases, however, the mutated gene has not been identified and it is necessary to provide wild-type function with a multigene interchromosomal duplication, i.e., a segment of the X inserted in another chromosome. If the proximal and distal extents of the duplicated segment are known, phenotypic rescue maps the mutated gene to the defined X chromosome region.Multigene deletions can also be used to map X-linked mutations by complementation, but crosses between individuals carrying deletions and X-linked lethal mutations are impossible without rescuing the lethality of either the deletion or the lethal mutation in males. Projects at the Bloomington Drosophila Stock Center and elsewhere (Parks et al. 2004; Ryder et al. 2007) have generated large collections of deletions with molecularly defined breakpoints in Drosophila melanogaster, but the utility of the X deletions is limited without duplications of the corresponding chromosomal regions.Duplications are potentially important for gene discovery. Identifying sets of genes involved in the same cellular process is a major focus of functional genomics research and this can be accomplished genetically by identifying dosage-sensitive modifiers of mutant phenotypes. Often, increasing or decreasing the copy number of a gene will enhance or suppress the phenotype associated with mutating another gene involved in the same process. Screening collections of deletions is a popular way to identify interacting genes in Drosophila (for examples, see Seher et al. 2007; Zhao et al. 2008; Aerts et al. 2009; Salzer et al. 2010) and was a major impetus for the assembly of the Bloomington Stock Center “Deficiency Kit,” which provides maximal coverage of the genome with the fewest deletions. Though dosage-sensitive modifiers could also be identified using increased gene dosage, the use of duplications in enhancer and suppressor screens remains largely unexplored. Assembling sets of duplications providing efficient genomic coverage would likely popularize this experimental approach.The size of duplicated segments determines how duplication chromosomes are used experimentally. Small duplicated segments allow high resolution gene mapping, but they are not suitable for other purposes. Only large duplicated segments are capable of rescuing the lethality of sizable multigene X deletions. Likewise, large duplicated segments provide efficiency in initially localizing mutations and identifying dosage-dependent modifiers. Despite their usefulness, interchromosomal duplications of large segments are among the hardest chromosomal rearrangements to isolate. In Drosophila, many existing duplications were recovered fortuitously as three-breakpoint aberrations following irradiation, but such rearrangements are rare and difficult to identify in screens. Other duplications were methodically constructed from preexisting rearranged chromosomes. This approach works well when it is possible, but it can be used only when progenitor aberrations with appropriate breakpoints are available. Because of these difficulties, the selection of duplication strains generated by Drosophila workers over the past several decades is not satisfactory for many purposes. The duplications are often difficult to use experimentally, their breakpoints are sparsely distributed along the X chromosome and only roughly mapped, and substantial gaps in coverage exist. Obviously, improved duplication resources are needed.Here we present the methodology and progress of a project at the Bloomington Drosophila Stock Center to construct interchromosomal duplications of large, megabase-scale X segments. Our approach builds on the long history of manipulating Drosophila chromosomes in vivo (Novitski and Childress 1976; Ashburner et al. 2005), but we have eliminated the need for preexisting aberrations by generating progenitor chromosomes using the FLP-FRT system. Indeed, this site-specific recombination system has had an enormous impact on the ability of fly geneticists to engineer many kinds of novel chromosomes (Golic and Golic 1996; Parks et al. 2004; Ryder et al. 2007). We will demonstrate how we have combined FLP-mediated recombination and other chromosome manipulation techniques to produce Y-linked duplications of large X segments. As we will show, appending X segments to Y chromosomes rather than autosomes has advantages both for the synthesis and experimental use of X duplications.To date, we have generated a minimum of 78% X coverage with duplication breakpoints spaced a median of nine genes apart. We anticipate completion of the project within the coming year. Using these duplications, mutations and genetic modifiers can be mapped first to large X intervals using a tiling set of the largest duplicated segments and then to small chromosome intervals using subsets of the duplications. These duplications will also facilitate deletion mapping. The creation of a set of stocks providing complete duplication coverage and extensive breakpoint subdivision of the X chromosome in a consistent genetic background will remove an impediment to investigating the functions of X-linked genes that has frustrated generations of Drosophila geneticists.  相似文献   

9.
While genetic screens have identified many genes essential for neurite outgrowth, they have been limited in their ability to identify neural genes that also have earlier critical roles in the gastrula, or neural genes for which maternally contributed RNA compensates for gene mutations in the zygote. To address this, we developed methods to screen the Drosophila genome using RNA-interference (RNAi) on primary neural cells and present the results of the first full-genome RNAi screen in neurons. We used live-cell imaging and quantitative image analysis to characterize the morphological phenotypes of fluorescently labelled primary neurons and glia in response to RNAi-mediated gene knockdown. From the full genome screen, we focused our analysis on 104 evolutionarily conserved genes that when downregulated by RNAi, have morphological defects such as reduced axon extension, excessive branching, loss of fasciculation, and blebbing. To assist in the phenotypic analysis of the large data sets, we generated image analysis algorithms that could assess the statistical significance of the mutant phenotypes. The algorithms were essential for the analysis of the thousands of images generated by the screening process and will become a valuable tool for future genome-wide screens in primary neurons. Our analysis revealed unexpected, essential roles in neurite outgrowth for genes representing a wide range of functional categories including signalling molecules, enzymes, channels, receptors, and cytoskeletal proteins. We also found that genes known to be involved in protein and vesicle trafficking showed similar RNAi phenotypes. We confirmed phenotypes of the protein trafficking genes Sec61alpha and Ran GTPase using Drosophila embryo and mouse embryonic cerebral cortical neurons, respectively. Collectively, our results showed that RNAi phenotypes in primary neural culture can parallel in vivo phenotypes, and the screening technique can be used to identify many new genes that have important functions in the nervous system.  相似文献   

10.
Previously, we successfully cloned a d-cycloserine (d-CS) biosynthetic gene cluster consisting of 10 open reading frames (designated dcsA to dcsJ) from d-CS-producing Streptomyces lavendulae ATCC 11924. In this study, we put four d-CS biosynthetic genes (dcsC, dcsD, dcsE, and dcsG) in tandem under the control of the T7 promoter in an Escherichia coli host. SDS-PAGE analysis demonstrated that the 4 gene products were simultaneously expressed in host cells. When l-serine and hydroxyurea (HU), the precursors of d-CS, were incubated together with the E. coli resting cell suspension, the cells produced significant amounts of d-CS (350 ± 20 μM). To increase the productivity of d-CS, the dcsJ gene, which might be responsible for the d-CS excretion, was connected downstream of the four genes. The E. coli resting cells harboring the five genes produced d-CS at 660 ± 31 μM. The dcsD gene product, DcsD, forms O-ureido-l-serine from O-acetyl-l-serine (OAS) and HU, which are intermediates in d-CS biosynthesis. DcsD also catalyzes the formation of l-cysteine from OAS and H2S. To repress the side catalytic activity of DcsD, the E. coli chromosomal cysJ and cysK genes, encoding the sulfite reductase α subunit and OAS sulfhydrylase, respectively, were disrupted. When resting cells of the double-knockout mutant harboring the four d-CS biosynthetic genes, together with dcsJ, were incubated with l-serine and HU, the d-CS production was 980 ± 57 μM, which is comparable to that of d-CS-producing S. lavendulae ATCC 11924 (930 ± 36 μM).  相似文献   

11.
12.
The immutability of the genetic code has been challenged with the successful reassignment of the UAG stop codon to non-natural amino acids in Escherichia coli. In the present study, we demonstrated the in vivo reassignment of the AGG sense codon from arginine to l-homoarginine. As the first step, we engineered a novel variant of the archaeal pyrrolysyl-tRNA synthetase (PylRS) able to recognize l-homoarginine and l-N6-(1-iminoethyl)lysine (l-NIL). When this PylRS variant or HarRS was expressed in E. coli, together with the AGG-reading tRNAPylCCU molecule, these arginine analogs were efficiently incorporated into proteins in response to AGG. Next, some or all of the AGG codons in the essential genes were eliminated by their synonymous replacements with other arginine codons, whereas the majority of the AGG codons remained in the genome. The bacterial host''s ability to translate AGG into arginine was then restricted in a temperature-dependent manner. The temperature sensitivity caused by this restriction was rescued by the translation of AGG to l-homoarginine or l-NIL. The assignment of AGG to l-homoarginine in the cells was confirmed by mass spectrometric analyses. The results showed the feasibility of breaking the degeneracy of sense codons to enhance the amino-acid diversity in the genetic code.  相似文献   

13.
14.
RNAi-mediated gene knockdown in Drosophila melanogaster is a powerful method to analyze loss-of-function phenotypes both in cell culture and in vivo. However, it has also become clear that false positives caused by off-target effects are prevalent, requiring careful validation of RNAi-induced phenotypes. The most rigorous proof that an RNAi-induced phenotype is due to loss of its intended target is to rescue the phenotype by a transgene impervious to RNAi. For large-scale validations in the mouse and Caenorhabditis elegans, this has been accomplished by using bacterial artificial chromosomes (BACs) of related species. However, in Drosophila, this approach is not feasible because transformation of large BACs is inefficient. We have therefore developed a general RNAi rescue approach for Drosophila that employs Cre/loxP-mediated recombination to rapidly retrofit existing fosmid clones into rescue constructs. Retrofitted fosmid clones carry a selection marker and a phiC31 attB site, which facilitates the production of transgenic animals. Here, we describe our approach and demonstrate proof-of-principle experiments showing that D. pseudoobscura fosmids can successfully rescue RNAi-induced phenotypes in D. melanogaster, both in cell culture and in vivo. Altogether, the tools and method that we have developed provide a gold standard for validation of Drosophila RNAi experiments.RNAi-mediated gene knockdown, whereby an exogenous double stranded RNA (dsRNA) is used to trigger homology-dependent suppression of the target gene, is an effective loss-of-function method to interrogate gene function. The RNAi technology in Drosophila melanogaster is widely used for genomewide RNAi screens in cell culture (see review by Perrimon and Mathey-Prevot 2007a), and more recently has been extended to large scale in vivo studies (Dietzl et al. 2007; Ni et al. 2009; Mummery-Widmer et al. 2009). Gene knockdown by RNAi is achieved by the introduction of dsRNAs into cultured cells or by inducible overexpression of “hairpin” dsRNAs in transgenic flies. In the context of in vivo RNAi screening, the combination of a tissue-specific GAL4 driver with a GAL4-responsive hairpin dsRNA transgene allows knockdown of the target gene only in the desired cells, thus providing a powerful way of probing biological processes that have been so far difficult to investigate.Analysis of the specificity of long dsRNAs in Drosophila cells has revealed that these reagents, depending on their sequences and levels of expression, can knock down genes others than the intended target (Kulkarni et al. 2006; Ma et al. 2006). This phenomenon is not specific to long dsRNAs and has also been commonly observed with 21-nt long siRNAs and shRNAs used in mammalian RNAi screens. In fact the rate of false positives associated with off-target effects observed in mammalian screens is usually higher than those observed with long dsRNAs (Echeverri and Perrimon 2006). Unwanted false positives created by off-target effects are a major problem in RNAi screens and require lengthy secondary validation tests (Echeverri and Perrimon 2006; Perrimon and Mathey-Prevot 2007b; Ramadan et al. 2007). Further, false positives associated with RNAi reagents are not limited to tissue culture experiments, as they have also been reported in the context of transgenic RNAi. For example, ∼25% of the hairpins targeting nonessential genes cause lethality when driven by the constitutively expressed Act5C-GAL4 driver (Dietzl et al. 2007; Ni et al. 2009).A number of approaches can be used to validate the specificity of RNAi-induced phenotypes (Echeverri and Perrimon 2006). These include validation by multiple dsRNAs that target the same gene but that do not overlap in sequence, comparison of knockdown efficiencies of multiple dsRNAs and the phenotypic strengths, and rescue of the phenotype by either cDNAs or genomic DNAs. Rescue of RNAi phenotypes constitutes the gold standard in the field as it provides unambiguous proof that the targeted gene is indeed responsible for the phenotype observed. In Drosophila cell culture experiments, cDNAs that lack the original 3′-untranslated region (UTR) have been used to rescue phenotypes induced by dsRNAs targeting the 3′-UTR (Yokokura et al. 2004; Stielow et al. 2008). In mammalian cell culture experiments, cDNAs that have a silent point mutation in the region targeted by an siRNA are commonly used (Lassus et al. 2002). The intrinsic problem of these approaches, however, is that overexpression of cDNAs alone can evoke abnormal cellular responses on their own, complicating interpretation of the results. A cleaner method is based on cross-species rescue that uses genomic DNA from a different species whose sequence is divergent enough from the host species to make it refractory to the RNAi reagent directed against the host gene. This approach effectively addresses the issue of overexpression artifact, as the rescue transgene is expressed from its endogenous promoter, ensuring proper levels and precise spatiotemporal regulation of gene expression. Cross-species rescue methods that use bacterial artificial chromosome (BACs) retrofitted with an appropriate selection marker have been described for mammals and C. elegans (Kittler et al. 2005; Sarov et al. 2006). However, the BAC strategies are not realistic for large-scale studies, because transformation of BACs, which are typically larger than 100 kb, is inefficient, albeit not impossible, in Drosophila (Venken et al. 2006).To provide a feasible way to validate large-scale RNAi screening results, we decided to develop a universal method for cross-species RNAi rescue in Drosophila. We chose to use fosmids, which are single-copy bacterial vectors with a cloning capacity of ∼40 kb, rather than BACs because (1) transformation of plasmids around this size is relatively efficient (Venken et al. 2006) and (2) end-sequenced fosmid clones for 11 different Drosophila species generated by the Drosophila species genome project are now publicly available (Richards et al. 2005; Drosophila 12 Genomes Consortium 2007).  相似文献   

15.
Whole-genome sequencing (WGS) of organisms displaying a specific mutant phenotype is a powerful approach to identify the genetic determinants of a plethora of biological processes. We have previously validated the feasibility of this approach by identifying a point-mutated locus responsible for a specific phenotype, observed in an ethyl methanesulfonate (EMS)-mutagenized Caenorhabditis elegans strain. Here we describe the genome-wide mutational profile of 17 EMS-mutagenized genomes as assessed with a bioinformatic pipeline, called MAQGene. Surprisingly, we find that while outcrossing mutagenized strains does reduce the total number of mutations, a striking mutational load is still observed even in outcrossed strains. Such genetic complexity has to be taken into account when establishing a causative relationship between genotype and phenotype. Even though unintentional, the 17 sequenced strains described here provide a resource of allelic variants in almost 1000 genes, including 62 premature stop codons, which represent candidate knockout alleles that will be of further use for the C. elegans community to study gene function.INDUCING molecular lesions in a genome is an effective approach to interrogate the genome for its functional elements. Molecular lesions can be induced using a variety of methods. Because of their efficiency and their ability to generate alleles with various different alterations in gene activity (e.g., amorphic, antimorphic, hypomorphic, and hypermorphic), chemical mutagens, such as ethyl methanesulfonate (EMS), are frequently used in genetic mutant screens (Anderson 1995). However, due to mutagen efficiency, a mutant animal selected for a single-locus phenotype invariably contains EMS-induced “background mutations” in its genome. Experimenters try to minimize the potential impact of background mutations through outcrossing to animals with a wild-type genome. Yet no full snapshots of genome sequences right after EMS mutagenesis and after outcrossing have so far been provided to illustrate the extent of background mutations and the extent to which they can indeed be eliminated.Another caveat of using base-changing chemical mutagens is the relative difficulty associated with identifying the phenotype-causing molecular lesion. In multicellular genetic model organisms, mutant identification involves time-consuming positional cloning approaches, usually involving breeding with genetically marked strains that allow pinpointing of the location of a molecular lesion. Even with rapid, SNP-based mapping approaches in animals with short generation times, such as Caenorhabditis elegans, substantial time hurdles, particularly in the final, fine-mapping stages, still exist. Conceptually similar problems in defining the location of a molecular lesion are encountered by human geneticists who attempt to identify disease-causing genetic lesions.Whole-genome sequencing (WGS) is beginning to emerge as an efficient and cost-effective tool to shortcut time-consuming mapping and positional cloning efforts (Hobert 2010). The sequencing of an entire genome and its ensuing comparison to a wild-type reference genome can potentially directly pinpoint the molecular lesion that results in the mutant phenotype the animal has been selected for. Proof-of-concept studies in bacteria, yeast, plants, worms, and flies have validated the applicability of this approach (Sarin et al. 2008; Smith et al. 2008; Srivatsan et al. 2008; Blumenstiel et al. 2009; Irvine et al. 2009; Flowers et al. 2010).Present-day deep sequencing platforms used for WGS generate relatively short sequence reads, thereby posing the bioinformatic challenge to align those reads to a reference genome. We previously described a software pipeline, MAQGene, which is based on the standard alignment program MAQ (Li et al. 2008) and facilitates this bioinformatic step by providing the end user with an extensively curated list of sequence variants from a WGS run of a mutated genome compared to a reference genome (Bigelow et al. 2009). This pipeline can be used for well-annotated, assembled genomes, such as C. elegans or Drosophila. In this article, we describe that this pipeline can identify not only point mutations but also deletions. We then use this pipeline to analyze a total of 17 EMS-mutagenized genomes. We find that EMS-mutagenized genomes carry a significant mutational load including presumptive loss-of-function alleles in several protein-coding genes that can lead to synthetic genetic interactions, one of which we describe here in more detail. We show that outcrossing to wild-type animals can lighten the mutational load; however, a substantial number of sequence variants are also introduced during outcrossing. Even though background mutations uncovered by WGS may complicate the interpretation of mutant phenotypes, they do provide a potentially useful source for functional studies of the affected genes.  相似文献   

16.
Lactobacillus casei strains 64H and BL23, but not ATCC 334, are able to ferment d-ribitol (also called d-adonitol). However, a BL23-derived ptsI mutant lacking enzyme I of the phosphoenolpyruvate:carbohydrate phosphotransferase system (PTS) was not able to utilize this pentitol, suggesting that strain BL23 transports and phosphorylates d-ribitol via a PTS. We identified an 11-kb region in the genome sequence of L. casei strain BL23 (LCABL_29160 to LCABL_29270) which is absent from strain ATCC 334 and which contains the genes for a GlpR/IolR-like repressor, the four components of a mannose-type PTS, and six metabolic enzymes potentially involved in d-ribitol metabolism. Deletion of the gene encoding the EIIB component of the presumed ribitol PTS indeed prevented d-ribitol fermentation. In addition, we overexpressed the six catabolic genes, purified the encoded enzymes, and determined the activities of four of them. They encode a d-ribitol-5-phosphate (d-ribitol-5-P) 2-dehydrogenase, a d-ribulose-5-P 3-epimerase, a d-ribose-5-P isomerase, and a d-xylulose-5-P phosphoketolase. In the first catabolic step, the protein d-ribitol-5-P 2-dehydrogenase uses NAD+ to oxidize d-ribitol-5-P formed during PTS-catalyzed transport to d-ribulose-5-P, which, in turn, is converted to d-xylulose-5-P by the enzyme d-ribulose-5-P 3-epimerase. Finally, the resulting d-xylulose-5-P is split by d-xylulose-5-P phosphoketolase in an inorganic phosphate-requiring reaction into acetylphosphate and the glycolytic intermediate d-glyceraldehyde-3-P. The three remaining enzymes, one of which was identified as d-ribose-5-P-isomerase, probably catalyze an alternative ribitol degradation pathway, which might be functional in L. casei strain 64H but not in BL23, because one of the BL23 genes carries a frameshift mutation.  相似文献   

17.
An l-glucose-utilizing bacterium, Paracoccus sp. 43P, was isolated from soil by enrichment cultivation in a minimal medium containing l-glucose as the sole carbon source. In cell-free extracts from this bacterium, NAD+-dependent l-glucose dehydrogenase was detected as having sole activity toward l-glucose. This enzyme, LgdA, was purified, and the lgdA gene was found to be located in a cluster of putative inositol catabolic genes. LgdA showed similar dehydrogenase activity toward scyllo- and myo-inositols. l-Gluconate dehydrogenase activity was also detected in cell-free extracts, which represents the reaction product of LgdA activity toward l-glucose. Enzyme purification and gene cloning revealed that the corresponding gene resides in a nine-gene cluster, the lgn cluster, which may participate in aldonate incorporation and assimilation. Kinetic and reaction product analysis of each gene product in the cluster indicated that they sequentially metabolize l-gluconate to glycolytic intermediates, d-glyceraldehyde-3-phosphate, and pyruvate through reactions of C-5 epimerization by dehydrogenase/reductase, dehydration, phosphorylation, and aldolase reaction, using a pathway similar to l-galactonate catabolism in Escherichia coli. Gene disruption studies indicated that the identified genes are responsible for l-glucose catabolism.  相似文献   

18.
The transfer of mitochondrial genes to the nucleus is a recurrent and consistent feature of eukaryotic genome evolution. Although many theories have been proposed to explain such transfers, little relevant data exist. The observation that clonal and self-fertilizing plants transfer more mitochondrial genes to their nuclei than do outcrossing plants contradicts predictions of major theories based on nuclear recombination and leaves a gap in our conceptual understanding how the observed pattern of gene transfer could arise. Here, with a series of deterministic and stochastic simulations, we show how epistatic selection and relative mutation rates of mitochondrial and nuclear genes influence mitochondrial-to-nuclear gene transfer. Specifically, we show that when there is a benefit to having a mitochondrial gene present in the nucleus, but absent in the mitochondria, self-fertilization dramatically increases both the rate and the probability of gene transfer. However, absent such a benefit, when mitochondrial mutation rates exceed those of the nucleus, self-fertilization decreases the rate and probability of transfer. This latter effect, however, is much weaker than the former. Our results are relevant to understanding the probabilities of fixation when loci in different genomes interact.GENOMIC investigations of obligate intracellular endosymbionts (Moran and Wernegreen 2000; Akman et al. 2002; Tamas et al. 2002; Wernegreen et al. 2002; Klasson and Andersson 2004; Foster et al. 2005) reveal a reduction in genome size and number of protein-coding genes compared to their free-living relatives (Charles et al. 1999; Gil et al. 2002; Wernegreen et al. 2002; Moran 2003; Van Ham et al. 2003; Klasson and Andersson 2004; Khachane et al. 2007). Similarly, mitochondria—ancient obligate intracellular symbionts of eukaryotes—have retained very few protein-coding genes (Boore 1999; Adams et al. 2002) [Reclinomonas americanas is at the extreme of retention of mitochondrial genes (Lang et al. 1997)]. Understanding the process of gene loss in mitochondria and other endosymbionts is a major research focus of mitochondrial and endosymbiont genomics (Moran 2003; Timmis et al. 2004; Khachane et al. 2007; Bock and Timmis 2008).The loss of endosymbiont genes can be complete, in which lost genes are absent from the host–endosymbiont complex, a substitution, in which a nuclear allele functions in place of the lost symbiont gene, or a functional transfer of an endosymbiont gene to the nucleus, followed by its loss (Adams and Palmer 2003). Such “functional transfer” involves the relocation of a mitochondrial gene to the nucleus, its acquisition of a promoter, successful targeting to the mitochondria for proper function, and the eventual loss of the gene from the mitochondrial genome altogether. Although this process is probably quite complex and requires numerous evolutionary modifications (Murcha et al. 2005), there is evidence that some mitochondrial genes are preadapted to functional transfer as they contain signals that target them to the mitochondria before functional transfer to the nucleus (Ueda et al. 2008a). The complex evolution of rps16 is an illuminating case of both functional gene transfer and substitution. In some lineages, the mitochondrial rps16 is functionally expressed in the nucleus but absent from the mitochondria (functional transfer) while in a subset of taxa, the chloroplast copy is also absent and the nuclear gene is also targeted to the chloroplast [substitution (Ueda et al. 2008b)].A number of evolutionary scenarios have been proposed to account for the massive loss of genes from endosymbionts. A subset of models argues that endosymbiont gene loss is a neutral or nearly neutral process. Since endosymbiosis reduces the strength of selection on genes that are unnecessary or redundant in an obligate intracellular environment, these genes may be quickly lost by the neutral fixation of a deletion or other loss-of-function mutations. Moreover, even when selection favors the retention of genes in endosymbionts, such selection may be ineffective because of reduction in effective population size due to recurrent bottlenecking (Rispe and Moran 2000). Additionally, frequent input of functional endosymbiont genes into the nucleus makes symbiont genes redundant, exacerbating gene loss via functional transfer (Berg and Kurland 2000).An alternative class of explanations views the loss of mitochondrial genes (be it complete loss, substitution, or functional transfer) as an adaptive process. The “mitochondrial competition theory” argues that mitochondrial genomes that either do not contain or do not express a given allele have a replicative advantage over other mitochondria, providing a within-host selective advantage to mitochondrial gene loss (Albert et al. 1996; Selosse et al. 2001; Yamauchi 2005). The “benefits of sex” model posits that the genomic diploid nuclear environment (diploid, sexual) is in some way preferable (e.g., as an escape from Muller''s ratchet or Hill–Robertson interference) to a haploid asexual mitochondrial environment (Blanchard and Lynch 2000). The epistatic model (Wade and Goodnight 2006) does not advance a specific or consistent benefit to transfer, but posits that transfer is explicitly a process of coevolution between mitochondrial and nuclear genomes, where fitness is a function of the gene combination rather than of either gene separately.Because few species are currently undergoing mitochondrial to nuclear gene transfer, these alternative hypotheses are difficult to distinguish with direct experimentation. However, the distribution of transferred genes across lineages allows for evaluation of the alternative hypotheses. For example, self-pollination reduces the rate of heteroplasmy and consequently the opportunity for competition among genetically distinct mitochondria. Thus, the mitochondrial competition theory predicts an excess of transfer events in sexual, outcrossing lineages, with high degrees of “paternal leakage.” Similarly, frequent self-fertilization diminishes the benefits of sex, and thus the benefits of sex hypothesis predicts fewer transfers in selfing and clonally reproducing plants than in outcrossing taxa. The epistatic model makes the opposite prediction. Selfing and clonal reproduction maintain cyto-nuclear gene combinations and increase the response to selection on epistatic combinations, potentially encouraging transfer. On the other hand, outcrossing tends to break apart adaptive cyto-nuclear gene combinations, potentially decreasing the amount of adaptive transfer in outcrossing lineages.Plant lineages with high levels of self-fertilization or asexual reproduction transfer more mitochondrial genes to their nuclei than predominantly sexual and outcrossing lineages (Brandvain et al. 2007). This result is consistent with predictions of the epistatic model and is contrary to predictions of the mitochondrial competition or benefits of sex models. More specific predictions allowing further empirical tests require more detailed theoretical investigations of the gene transfer process. Here, we investigate the roles of mutation, selection, and random drift in gene transfer using both deterministic models and stochastic simulations to refine and extend predictions of patterns of functional mitochondrial to nuclear gene transfer.  相似文献   

19.
Conditionally expressed genes have the property that every individual in a population carries and transmits the gene, but only a fraction, φ, expresses the gene and exposes it to natural selection. We show that a consequence of this pattern of inheritance and expression is a weakening of the strength of natural selection, allowing deleterious mutations to accumulate within and between species and inhibiting the spread of beneficial mutations. We extend previous theory to show that conditional expression in space and time have approximately equivalent effects on relaxing the strength of selection and that the effect holds in a spatially heterogeneous environment even with low migration rates among patches. We support our analytical approximations with computer simulations and delineate the parameter range under which the approximations fail. We model the effects of conditional expression on sequence polymorphism at mutation–selection–drift equilibrium, allowing for neutral sites, and show that sequence variation within and between species is inflated by conditional expression, with the effect being strongest in populations with large effective size. As φ decreases, more sites are recruited into neutrality, leading to pseudogenization and increased drift load. Mutation accumulation diminishes the degree of adaptation of conditionally expressed genes to rare environments, and the mutational cost of phenotypic plasticity, which we quantify as the plasticity load, is greater for more rarely expressed genes. Our theory connects gene-level relative polymorphism and divergence with the spatial and temporal frequency of environments inducing gene expression. Our theory suggests that null hypotheses for levels of standing genetic variation and sequence divergence must be corrected to account for the frequency of expression of the genes under study.IN genetically and ecologically subdivided populations, some individuals will experience a local environment very different from others, making it difficult to evolve a single adaptation adequate for all local conditions. Phenotypic plasticity allows organisms to respond adaptively to spatially and temporally varying environments by developing alternative phenotypes that enhance fitness under local conditions (Scheiner 1993; Via et al. 1995). Examples of alternative phenotypes, i.e., polyphenisms, include the defensive morphologies in Daphnia and algae induced by the presence of predators (e.g., Lively 1986; DeWitt 1998; Harvell 1998; Hazel et al. 2004); the winged and wingless morphs of bean beetles responding to resource variation (e.g., Abouheif and Wray 2002; Roff and Gelinas 2003; Lommen et al. 2005); and bacterial genes involved in traits such as quorum sensing, antibiotic production, biofilm formation, and virulence (Fuqua et al. 1996). The developmental basis of such alternative phenotypes often lies in the inducible expression of some genes in some individuals by environmental variables. That is, all individuals carry and transmit the conditionally expressed genes but only a fraction of individuals, φ, express them when environmental conditions are appropriate.The genes underlying plastic traits should experience relaxed selection due to conditional expression. Wade and co-workers have shown that genes hidden from natural selection in a fraction of individuals in the population by X-linked (Whitlock and Wade 1995; Linksvayer and Wade 2009) or sex-limited expression (Wade 1998; Demuth and Wade 2007) experience relaxed selective constraint. In Drosophila spp., sequence data for genes with maternally limited expression quantitatively support the theoretical predictions both for within-species polymorphism (Barker et al. 2005; Cruickshank and Wade 2008) and for between-species divergence (Barker Et Al 2005; Demuth and Wade 2007; Cruickshank and Wade 2008). Furthermore, male-specific genes in the facultatively sexual pea aphid have been shown to have elevated levels of sequence variation due to relaxed selection (Brisson and Nuzhdin 2008). Genes with spatially restricted expression in a heterogeneous environment should likewise experience relaxed selection. Adaptation to the most common environment in an ecologically subdivided population (Rosenzweig 1987; Holt and Gaines 1992; Holt 1996) allows deleterious mutations to accumulate in traits expressed in rare environments (Kawecki 1994; Whitlock 1996).Here we extend these results by quantifying the consequences of relaxed selection on conditionally expressed genes. Specifically, we show that, with weak selection, spatial and temporal fluctuations in selection intensity generate approximately equivalent effects on mean trait fitness, even with low rates of migration between habitats, resulting in a great simplification of analytical results. Our analytical approximations are supported with deterministic and stochastic simulations, and we note the conditions under which the approximations fail. We then derive general expressions for (1) the expected level of sequence polymorphism within populations under mutation, migration, drift, and purifying selection with conditional gene expression; (2) the rate of sequence divergence among populations, for dominant and recessive mutations; and (3) the reduction in mean population fitness due to accumulation of deleterious mutations at conditionally expressed loci. We find that the rate of accumulation of deleterious mutations for conditionally expressed genes is accelerated and the probability of fixation of beneficial mutations is reduced, causing a reduction in the fitness of conditional traits and an inflation in sequence variation within and between species. Our results suggest that evolutionary null hypotheses must be adjusted to account for the frequency of expression of genes under study, such that signatures of elevated within- or between-species sequence variation are not necessarily evidence of the action of diversifying natural selection. Furthermore, if conditional expression is due to spatial heterogeneity, we show that the level of genetic variation in a sample will often depend on whether or not genotypes were sampled from the selective habitat, the neutral habitat, or both. In the discussion we address the scope and limitations of our theory, as well as its implications for the maintenance of genetic variation, adaptive divergence between species, constraints on phenotypic plasticity, and evolutionary inference from sequence data.  相似文献   

20.
This study quantifies the effects of naturally occurring X-linked variation on immune response in Drosophila melanogaster to assess associations between immunity genotypes and innate immune response. We constructed a set of 168 X-chromosomal extraction lines, incorporating X chromosomes from a natural population into co-isogenic autosomal backgrounds, and genotyped the lines at 88 SNPs in 20 X-linked immune genes. We find that genetic variation in many of the genes is associated with immune response phenotypes, including bacterial load and immune gene expression. Many of the associations act in a sex-specific or sexually antagonistic manner, supporting the theory that with the selective pressures facing genes on the X chromosome, sexually antagonistic variation may be more easily maintained.THE deep evolutionary conservation of many specific genes in innate immunity underscores the potent forces of natural selection maintaining this vital function. While it is widely accepted as the ancestral form of immune response, its role in the activation of adaptive immune response further motivates investigation into variation in its function (Medzhitov and Janeway 1997). Drosophila has been used as a valuable model organism to identify and characterize functions of the components of innate immune pathways as well as the evolutionary patterns present among the genes comprising these pathways (reviewed in Brennan and Anderson 2004; Irving et al. 2004; Ferrandon et al. 2007). The humoral response, resulting in the production of antimicrobial peptides in response to bacterial or fungal infection, relies mainly on Toll and imd signal transduction pathways, both of which are highly homologous to pathways in mammalian immunity (reviewed in Kimbrell and Beutler 2001). The cellular component, on the other hand, incorporates phagocytic engulfment as well as melanization and encapsulation of infecting particles. While less well defined in the Drosophila model, portions of other systems also appear to affect the effectiveness of immune response, including JAK/STAT and JNK signaling pathways, hematopoesis, and iron metabolism.Population genetic analysis can be used to determine whether sequence polymorphism and divergence patterns among Drosophila genes in innate immune pathways are consistent with signatures of selection acting within and between species of flies. If, for example, the innate immune pathways are involved in an evolutionary “arms race” with pathogenic organisms, genes in these pathways would be expected to show signs of positive selection driven by evolutionary pressure to counter virulence mechanisms of invading microbes. When signs of selection (as inferred from sequence comparisons within Drosophila simulans populations and between D. simulans and D. melanogaster) in immune genes and nonimmune genes were evaluated, immune genes as a group were found to have higher KA/KS ratios than nonimmune genes, providing evidence for elevated adaptive evolution (Schlenke and Begun 2003). Since receptor, effector, and signaling proteins function in different portions of the immune response pathways, these may be exposed to differing levels of contact with invading microbes and may display nonuniform levels of functional redundancy or pleiotropy. Thus, genes from different functional groups may be exposed to distinct selective pressures. Antimicrobial peptides, which might be expected to encounter unique selective pressures due to their direct interactions with invading microbes, have shown little sign of positive selection, bearing low levels of amino acid divergence (Clark and Wang 1997; Date et al. 1998; Ramos-Onsins and Aguadé 1998; Lazzaro and Clark 2003). Furthermore, sequence analyses of immune-related receptors have shown evidence for purifying selection in peptidoglycan recognition proteins (PGRPs), while others, including some scavenger receptors (SRs), appear to be rapidly evolving under pressures consistent with positive selection (Jiggins and Hurst 2003; Lazzaro 2005). On a deeper evolutionary timescale, sequence comparisons between immune genes in multiple Drosophila species (based on full-genome sequence data) have shown striking differences among functional groups of immune genes, with recognition molecules showing much more positive selection than either signaling or effector genes (Sackton et al. 2007).Beyond using sequence data and the analysis of polymorphism and divergence to infer levels and modes of selection that have previously acted on immune genes (either individually or in functional groups), other studies have investigated correlations between autosomal variation in genotype and immune response phenotype in natural populations of Drosophila (Lazzaro et al. 2004, 2006). These experiments tested associations between naturally occurring genetic variation in immune-related genes and postinfection bacterial load. In these studies, genetic variation in many of the immune genes was found to associate significantly with one or more of the bacterial load phenotypes. Specifically, polymorphisms in autosomal genes encoding recognition and signaling proteins (but not antimicrobial peptides) associate consistently with bacterial load phenotypes, suggesting that not all functional classes of immune-related genes harbor equally influential genetic variation.The focus of this study is X-linked immune genes, which may be under unique regulatory and selective pressures simply because they are hemizygous in males, are dosage compensated, and face elevated influence of random genetic drift due to their smaller effective population size. As a consequence, the X chromosome should favor the more rapid fixation of beneficial recessive alleles and more rapid loss of harmful recessive alleles compared to the autosomes (Charlesworth et al. 1987; Singh et al. 2008). Thus, with different selective pressures compared to autosomal genes, X-linked immunity genes are expected to bear different standing levels of variation, and segregating polymorphisms in these genes may have different impacts on phenotype.Different exposures of X-linked genes to selection in males and females can also contribute to sexual dimorphism. Rice (1984) suggested that X-linked sexually antagonistic alleles may more freely influence sexually dimorphic traits than can those on autosomes. In fact, the X chromosome appears to favor the maintenance of sexually antagonistic variation (Gibson et al. 2002); if a given allele is slightly deleterious in one sex, it may be maintained in the population by being beneficial to the other sex. Immune-related genes may be particularly prone to bearing sexual dimorphism in Drosophila, since males and females have been shown to have different evolutionary optima for energetic expenditure on immune response, and thus their respective immune responses may differ on the basis of conditions such as food or reproductive resource availability (McKean and Nunney 2001, 2005). If sexually antagonistic traits are responsible for some of the observed sexual dimorphism, variation in X-linked genes could contribute to phenotypic differences, and so X-linked variation in immune genes could face unique selective pressures.In this report we investigate the standing levels of variation in X-linked immune genes in natural populations of D. melanogaster and quantify the impacts of that variation on immune response phenotypes. We genotyped 168 lines at single-nucleotide polymorphisms (SNPs) across 20 X-linked immunity loci and quantified postinfection bacterial load and immune gene expression phenotypes. We found significant variation across the lines for bacterial load after infection, and we were able to identify polymorphisms in immune-related genes that associate with immune response phenotypes individually and in interacting pairs of SNPs. Additionally, some of the genetic variation was found to associate with a sex difference in immune competence, with alleles acting in either a sex-specific or a sexually antagonistic manner. This provides evidence for X-linked genetic variation in immune-related loci associating with both phenotypic variation among lines and sex differences in these phenotypes.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号