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 203 毫秒
1.
Floral bilateral symmetry (zygomorphy) has evolved several times independently in angiosperms from radially symmetrical (actinomorphic) ancestral states. Homologs of the Antirrhinum majus Cycloidea gene (Cyc) have been shown to control floral symmetry in diverse groups in core eudicots. In the basal eudicot family Ranunculaceae, there is a single evolutionary transition from actinomorphy to zygomorphy in the stem lineage of the tribe Delphinieae. We characterized Cyc homologs in 18 genera of Ranunculaceae, including the four genera of Delphinieae, in a sampling that represents the floral morphological diversity of this tribe, and reconstructed the evolutionary history of this gene family in Ranunculaceae. Within each of the two RanaCyL (Ranunculaceae Cycloidea-like) lineages previously identified, an additional duplication possibly predating the emergence of the Delphinieae was found, resulting in up to four gene copies in zygomorphic species. Expression analyses indicate that the RanaCyL paralogs are expressed early in floral buds and that the duration of their expression varies between species and paralog class. At most one RanaCyL paralog was expressed during the late stages of floral development in the actinomorphic species studied whereas all paralogs from the zygomorphic species were expressed, composing a species-specific identity code for perianth organs. The contrasted asymmetric patterns of expression observed in the two zygomorphic species is discussed in relation to their distinct perianth architecture.  相似文献   

2.
Somatic cycloheximide-resistant mutants of syngen 1 of Tetrahymena pyriformis were isolated and genetically characterized. Two properties of the mutants were independently examined: (a) The transmission of the mutant phenotype during conjugation and (b) the kinetics of phenotypic assortment during vegetative propagation. The results of both studies strongly support the idea that these somatic mutations have a macronuclear location. The kinetics of assortment are consistent with the idea that the syngen 1 macronucleus contains about 45 assorting genetic units. The sib-selection method employed here, used in conjunction with the analysis of assortment kinetics and a previously described test for randomness of distribution, provides a probe of macronuclear genetics applicable to many ciliates, including those in which conjugation is not known to occur or is not under experimental control.  相似文献   

3.
As part of a study of protein folding, we have constructed a fine-structure map of 9 existing and 29 newly isolated UV- and hydroxylamine-induced temperature-sensitive (ts) mutations in gene 9 of Salmonella bacteriophage P22. Gene 9 specifies the polypeptide chain of the multimeric tail spikes, six of which form the cell attachment organelle of the phage. The 38 ts mutants were mapped against deletion lysogens with endpoints in gene 9. They mapped in 10 of the 15 deletion intervals. Two- and three-factor crosses between mutants within each interval indicated that at least 31 ts sites are represented among the 38 mutants. To determine the distribution of ts sites within the physical map, we identified the protein fragments from infection of su- hosts with 10 gene 9 amber mutants. Their molecular weights, ranging from 13,900 to 55,000 daltons, were combined with the genetic data to yield a composite map of gene 9. The 31 ts sites were distributed through most of the gene, but were most densely clustered in the central third.—None of the ts mutant pairs tested exhibited intragenic complementation. Studies of the defective phenotypes of the ts mutants (Goldenberg and King 1981; Smith and King 1981) revealed that most do not affect the thermostability of the mature protein, but instead prevent the folding or subunit assembly of the mutant chains synthesized at restrictive temperature. Thus, many of thes ts mutations identify sites in the polypeptide chain that are critical for the folding or maturation of the tail-spike protein.  相似文献   

4.
Bayesian inference methods are extensively used to detect the presence of population structure given genetic data. The primary output of software implementing these methods are ancestry profiles of sampled individuals. While these profiles robustly partition the data into subgroups, currently there is no objective method to determine whether the fixed factor of interest (e.g. geographic origin) correlates with inferred subgroups or not, and if so, which populations are driving this correlation. We present ObStruct, a novel tool to objectively analyse the nature of structure revealed in Bayesian ancestry profiles using established statistical methods. ObStruct evaluates the extent of structural similarity between sampled and inferred populations, tests the significance of population differentiation, provides information on the contribution of sampled and inferred populations to the observed structure and crucially determines whether the predetermined factor of interest correlates with inferred population structure. Analyses of simulated and experimental data highlight ObStruct''s ability to objectively assess the nature of structure in populations. We show the method is capable of capturing an increase in the level of structure with increasing time since divergence between simulated populations. Further, we applied the method to a highly structured dataset of 1,484 humans from seven continents and a less structured dataset of 179 Saccharomyces cerevisiae from three regions in New Zealand. Our results show that ObStruct provides an objective metric to classify the degree, drivers and significance of inferred structure, as well as providing novel insights into the relationships between sampled populations, and adds a final step to the pipeline for population structure analyses.  相似文献   

5.
The soil bacterium Bacillus subtilis forms biofilms on surfaces and at air-liquid interfaces. It was previously reported that these biofilms disassemble late in their life cycle and that conditioned medium from late-stage biofilms inhibits biofilm formation. Such medium contained a mixture of d-leucine, d-methionine, d-tryptophan, and d-tyrosine and was reported to inhibit biofilm formation via the incorporation of these d-amino acids into the cell wall. Here, we show that l-amino acids were able to specifically reverse the inhibitory effects of their cognate d-amino acids. We also show that d-amino acids inhibited growth and the expression of biofilm matrix genes at concentrations that inhibit biofilm formation. Finally, we report that the strain routinely used to study biofilm formation has a mutation in the gene (dtd) encoding d-tyrosyl-tRNA deacylase, an enzyme that prevents the misincorporation of d-amino acids into protein in B. subtilis. When we repaired the dtd gene, B. subtilis became resistant to the biofilm-inhibitory effects of d-amino acids without losing the ability to incorporate at least one noncanonical d-amino acid, d-tryptophan, into the peptidoglycan peptide side chain. We conclude that the susceptibility of B. subtilis to the biofilm-inhibitory effects of d-amino acids is largely, if not entirely, due to their toxic effects on protein synthesis.  相似文献   

6.
7.
Sex determination in plants involves a variety of mechanisms. Here, we report the map-based cloning and characterization of the unisexual-flower-controlling gene M. M was identified as a previously characterized putative 1-aminocyclopropane-1-carboxylic acid synthase gene, while the m allele that mutated at a conserved site (Gly33Cys) lost activity in the original enzymatically active allele.SEX determination in angiosperms, including crop plants, evolves a variety of mechanisms that involve a number of different genetic and epigenetic factors (Tanurdzic and Banks 2004). Due to its diversity in sex types and to the extensive physiological and genetic studies conducted on it, cucumber (Cucumis sativus L.; 2n = 2x = 14) is becoming a model plant for sex-determination research (Atsmon 1968; Tsao 1988; Perl-Treves 1999; Tanurdzic and Banks 2004). In cucumber plants, male and female flowers are generally produced separately in the same individual; however, certain lines also produce bisexual flowers. Preliminary genetic studies have indicated that three major genes are responsible for sex expression and segregation in the cucumber plantF/f, M/m, and A/a. The F gene may promote femaleness, while the m gene regulates the appearance of hermaphroditic flowers on the plant. Furthermore, in combination with the homozygous recessive f gene, the recessive a gene can intensify the androecious nature (Galun 1961; Robinson et al. 1976).Sex expression in cucumber plants can also be modified by various environmental factors and plant hormones such as ethylene (Atsmon 1968; Takahashi et al. 1983; Takahashi and Jaffe 1984; Perl-Treves 1999; Yamasaki et al. 2005). A series of studies (Kamachi et al. 1997, 2000; Trebitsh et al. 1997; Yamasaki et al. 2003a; Mibus and Tatlioglu 2004; Knopf and Trebitsh 2006) have been conducted to investigate the F/f gene. These studies have shown that CsACS1G, which encodes a key enzyme of the ethylene-synthesis pathway, is the candidate gene for the F/f locus. However, the M/m gene has not been studied in as much detail as the F/f gene. Here, we report the map-based cloning and characterization of the unisexual-flower-controlling M gene.  相似文献   

8.
The first enzyme in the pathway for l-arabinose catabolism in eukaryotic microorganisms is a reductase, reducing l-arabinose to l-arabitol. The enzymes catalyzing this reduction are in general nonspecific and would also reduce d-xylose to xylitol, the first step in eukaryotic d-xylose catabolism. It is not clear whether microorganisms use different enzymes depending on the carbon source. Here we show that Aspergillus niger makes use of two different enzymes. We identified, cloned, and characterized an l-arabinose reductase, larA, that is different from the d-xylose reductase, xyrA. The larA is up-regulated on l-arabinose, while the xyrA is up-regulated on d-xylose. There is however an initial up-regulation of larA also on d-xylose but that fades away after about 4 h. The deletion of the larA gene in A. niger results in a slow growth phenotype on l-arabinose, whereas the growth on d-xylose is unaffected. The l-arabinose reductase can convert l-arabinose and d-xylose to their corresponding sugar alcohols but has a higher affinity for l-arabinose. The Km for l-arabinose is 54 ± 6 mm and for d-xylose 155 ± 15 mm.  相似文献   

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

10.
We developed a novel process for efficient synthesis of l-threo-3-hydroxyaspartic acid (l-THA) using microbial hydroxylase and hydrolase. A well-characterized mutant of asparagine hydroxylase (AsnO-D241N) and its homologous enzyme (SCO2693-D246N) were adaptable to the direct hydroxylation of l-aspartic acid; however, the yields were strictly low. Therefore, the highly stable and efficient wild-type asparagine hydroxylases AsnO and SCO2693 were employed to synthesize l-THA. By using these recombinant enzymes, l-THA was obtained by l-asparagine hydroxylation by AsnO followed by amide hydrolysis by asparaginase via 3-hydroxyasparagine. Subsequently, the two-step reaction was adapted to one-pot bioconversion in a test tube. l-THA was obtained in a small amount with a molar yield of 0.076% by using intact Escherichia coli expressing the asnO gene, and thus, two asparaginase-deficient mutants of E. coli were investigated. A remarkably increased l-THA yield of 8.2% was obtained with the asparaginase I-deficient mutant. When the expression level of the asnO gene was enhanced by using the T7 promoter in E. coli instead of the lac promoter, the l-THA yield was significantly increased to 92%. By using a combination of the E. coli asparaginase I-deficient mutant and the T7 expression system, a whole-cell reaction in a jar fermentor was conducted, and consequently, l-THA was successfully obtained from l-asparagine with a maximum yield of 96% in less time than with test tube-scale production. These results indicate that asparagine hydroxylation followed by hydrolysis would be applicable to the efficient production of l-THA.  相似文献   

11.
Shewanella oneidensis couples oxidation of lactate to respiration of many substrates. Here we report that llpR (l-lactate-positive regulator, SO_3460) encodes a positive regulator of l-lactate utilization distinct from previously studied regulators. We also demonstrate d-lactate inhibition of l-lactate utilization in S. oneidensis, resulting in preferential utilization of the d isomer.  相似文献   

12.
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   

13.
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   

14.
15.
Weilong Hao  G. Brian Golding 《Genetics》2009,182(4):1365-1375
Lateral gene transfer (LGT) and gene rearrangement are essential for shaping bacterial genomes during evolution. Separate attention has been focused on understanding the process of lateral gene transfer and the process of gene translocation. However, little is known about how gene translocation affects laterally transferred genes. Here we have examined gene translocations and lateral gene transfers in closely related genome pairs. The results reveal that translocated genes undergo elevated rates of evolution and gene translocation tends to take place preferentially in recently acquired genes. Translocated genes have a high probability to be truncated, suggesting that translocation followed by truncation/deletion might play an important role in the fast turnover of laterally transferred genes. Furthermore, more recently acquired genes have a higher proportion of genes on the leading strand, suggesting a strong strand bias of lateral gene transfer.GENE insertions and deletions, together with gene translocations play important roles in bacterial genome evolution (Garcia-Vallvé et al. 2000; Ochman and Jones 2000; Tillier and Collins 2000a; Fraser-Liggett 2005). Gene insertions and deletions, as the essential driving forces in influencing gene content (Kunin and Ouzounis 2003), have received a great deal of attention. Various methods have been employed to study gene insertions and deletions previously; for instance, there are studies of population dynamics (Nielsen and Townsend 2004), such as a birth-and-death model of evolution (Berg and Kurland 2002; Novozhilov et al. 2005), phylogeny-dependent studies including parsimony methods (Daubin et al. 2003a,b; Mirkin et al. 2003; Hao and Golding 2004), and maximum-likelihood methods (Hao and Golding 2006b, 2008b). It has been shown that recently laterally transferred genes have high evolutionary rates and high rates of gene turnover (Daubin et al. 2003b; Hao and Golding 2004, 2006b).Gene rearrangement has also been commonly studied as another important driving force that shapes bacterial genomes (for a review, see Rocha 2004). Gene order changes in genomes are history dependent; for instance, fewer gene rearrangements are expected among more closely related species. Gene order within genomes has therefore been used to reconstruct phylogeny (Sankoff et al. 2000; Tamames 2001; Rogozin et al. 2004; Belda et al. 2005). Previous studies have focused mainly on lateral gene transfer (LGT) and gene rearrangement individually, but little is known about any association between laterally transferred genes and gene rearrangements. The study of gene order of laterally acquired genes might shed some light on the understanding of the LGT process.In this study, we have examined gene translocations and lateral gene transfers in closely related genome pairs. It is shown that the proportion of translocated genes among recently acquired genes is always high, while the proportion of translocated genes is always low in ancient genes, suggesting that gene translocation tends to take place in recently transferred genes. The results also reveal that translocated genes have elevated rates of evolution compared with positionally conserved genes and gene truncation is more prevalent in translocated genes. These findings suggest that gene translocation might accelerate the gene turnover of recently transferred genes and/or that genes likely to undergo translocation are those genes more likely to be laterally transferred and dispensable for the genome. Furthermore, the proportion of recently acquired genes is higher on the leading strand, suggesting that laterally transferred genes are biased toward being on the leading strand. After lateral transfer, some genes could be translocated to the lagging strand and some translocated genes are likely to be eliminated during evolution.  相似文献   

16.
17.
18.
With the most recent releases of the Drosophila melanogaster genome sequences, much of the previously absent heterochromatic sequences have now been annotated. We undertook an extensive genetic analysis of existing lethal mutations, as well as molecular mapping and sequence analysis (using a candidate gene approach) to identify as many essential genes as possible in the centromeric heterochromatin on the right arm of the second chromosome (2Rh) of D. melanogaster. We also utilized available RNA interference lines to knock down the expression of genes in 2Rh as another approach to identifying essential genes. In total, we verified the existence of eight novel essential loci in 2Rh: CG17665, CG17683, CG17684, CG17883, CG40127, CG41265, CG42595, and Atf6. Two of these essential loci, CG41265 and CG42595, are synonymous with the previously characterized loci l(2)41Ab and unextended, respectively. The genetic and molecular analysis of the previously reported locus, l(2)41Ae, revealed that this is not a single locus, but rather it is a large region of 2Rh that extends from unextended (CG42595) to CG17665 and includes four of the novel loci uncovered here.THE term “heterochromatin” was introduced by Heitz (1928) to describe regions of mitotic chromosomes that remain condensed throughout the cell cycle, in contrast to regions of euchromatin, which condense only during cell division. Heterochromatin was later divided into two classes: constitutive and facultative heterochromatin (Brown 1966). Constitutive heterochromatin is found in large blocks near centromeres and telomeres, while facultative heterochromatin can be described as silenced euchromatin that undergoes heterochromatization at specific developmental stages. Other properties of constitutive heterochromatin include late replication in S phase, low gene density, strikingly reduced level of meiotic recombination, enrichment in transposable element sequences and highly repetitive satellite DNA sequences, and the ability to silence euchromatic gene expression in a phenomenon called position effect variegation.Approximately 30% of the Drosophila melanogaster genome consists of constitutive heterochromatin (Gatti and Pimpinelli 1992). Centromeric heterochromatin in D. melanogaster is composed of mainly middle-repeat satellite DNA sequences and clusters of transposable element sequences (Lohe et al. 1993; Pimpinelli et al. 1995). Genes that reside in the heterochromatin are scattered like islands between the satellites and clusters of transposable elements. On average, heterochromatic genes are larger than euchromatic genes, primarily due to the prevalent accumulation of transposable element sequences in their introns (Devlin et al. 1990; Biggs et al. 1994; Dimitri et al. 2003a,b; Hoskins et al. 2007). Heterochromatic genes also tend to be AT-rich compared to their euchromatic counterparts; there is some evidence suggesting that the coding sequences of heterochromatic genes evolve toward AT richness in response to being located in heterochromatin (Yasuhara et al. 2005; Díaz-Castillo and Golic 2007).Drosophila heterochromatin is vastly under-replicated in polytene chromosomes, so heterochromatic genes cannot easily be mapped through polytene analysis. However, by using Hoechst 33258 and N-chromosome banding techniques, Dimitri (1991) was successful in dividing heterochromatin in mitotic chromosomes into distinct cytological bands; this was an important step in mapping the precise location of heterochromatic genes because before this time heterochromatic genes could be mapped only relative to one another. Here we focus on further refining the previous mapping work on essential genes in the proximal heterochromatin of the right arm of the second chromosome (2Rh) in cytological region h41–h46 of D. melanogaster (Hilliker 1976; Hilliker et al. 1980; Coulthard et al. 2003; Myster et al. 2004).Early mapping studies in D. melanogaster putatively placed the light (lt) and rolled (rl) genes in, or near, chromosome 2 heterochromatin (Schultz 1936; Hannah 1951; Hessler 1958). The first large-scale mutagenesis specifically directed at finding vital loci in second chromosome heterochromatin was conducted by Hilliker (1976). Using heterochromatic deletions created by Hilliker and Holm (1975), Hilliker (1976) set out to map vital loci using the mutagen ethyl methanesulfonate (EMS). He identified seven individual lethal complementation groups in 2Rh that were interpreted as representing seven vital loci. One of these heterochromatic loci was identified as the previously described rl gene. Two of the remaining vital loci have since been identified: Nipped-A is synonymous with the l(2) 41Ah complementation group (Rollins et al. 1999) and RpL38 is synonymous with Minute(2)41A and Hilliker''s (1976) l(2)41Af complementation group (Marygold et al. 2005; also referred to as l(2)Ag in FlyBase). In addition, Rollins et al. (1999) found the Nipped-B gene to be located in 2Rh, but how this locus fit into the data from Hilliker (1976) was unclear.With the limited release of some of the more distal heterochromatic sequences (Hoskins et al. 2002), a more recent mutagenesis screen focusing on distal 2Rh was conducted by Myster et al. (2004). In the region defined by the overlap between Df(2R)41A8 and Df(2R)41A10 (the latter was previously shown to be deficient for most of 2Rh; Hilliker and Holm 1975), Myster et al. (2004) reported the existence of 15 vital loci, considerably more than the 4 essential loci predicted by Hilliker (1976). The discrepancy between these two studies was the catalyst for this current work. Each group used the same mutagen, EMS, yet each group came up with very different interpretations of the number of vital loci.Hilliker''s interpretation relied on earlier evidence that EMS preferentially produced point mutations and not large-scale aberrations (Lim and Snyder 1974). Assuming that the mutants isolated in his study were point mutations, or small aberrations limited to one locus, Hilliker found that some of the loci that he identified exhibited complex interallelic complementation; the most complex complementation pattern was observed with locus l(2)41Ae. On the other hand, the interpretation of Myster et al. (2004) was that heterochromatin was more sensitive to EMS and that EMS could produce large heterochromatic deletions; they proposed that the complex interallelic complementation in l(2)41Ae was due to the presence of deletions and that l(2)41Ae represented a region of 2Rh containing many genes, rather than being a single locus.To resolve these different interpretations of the genomic segment containing l(2)41Ae (i.e., is it a single locus or a region of 2Rh), we set out to map l(2)41Ae and the region surrounding the presumed location of l(2)41Ae (as in Myster et al. 2004) by performing a large-scale inter se complementation analysis between all available mutant lines that were previously mapped to l(2)41Ae (including Nipped-B). In addition, we undertook a molecular mapping and sequence analysis, using a candidate gene approach with the most recent annotation of 2Rh (Hoskins et al. 2007), to characterize the region and identify as many essential genes as possible. We also used these approaches to map l(2)41Ab and unextended [two of the more proximal complementation groups identified by Hilliker (1976)]. Finally, we also utilized available RNA interference (RNAi) lines to knock down the expression of 12 genes in 2Rh in an attempt to identify essential genes.  相似文献   

19.
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