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 406 毫秒
1.
Eukaryotic translation initiation factor 2 (eIF2) is a G protein that delivers the methionyl initiator tRNA to the small ribosomal subunit and releases it upon GTP hydrolysis after the recognition of the initiation codon. eIF2 is composed of three subunits, alpha, beta, and gamma. Subunit gamma shows the strongest conservation, and it confers both tRNA and GTP/GDP binding. Using intron positioning and protein sequence alignment, here we show that eIF2gamma is a suitable phylogenetic marker for eukaryotes. We determined or completed the sequences of 13 arthropod eIF2gamma genes. Analyzing the phylogenetic distribution of 52 different intron positions in 55 distantly related eIF2gamma genes, we identified ancient ones and shared derived introns in our data set. Obviously, intron positioning in eIF2gamma is evolutionarily conserved. However, there were episodes of complete and partial intron losses followed by intron gains. We identified 17 clusters of intron positions based on their distribution. The evolution of these clusters appears to be connected with preferred exon length and can be used to estimate the relative timing of intron gain because nearby precursor introns had to be erased from the gene before the new introns could be inserted. Moreover, we identified a putative case of intron sliding that constitutes a synapomorphic character state supporting monophyly of Coleoptera, Lepidoptera, and Diptera excluding Hymenoptera. We also performed tree reconstructions using the eIF2gamma protein sequences and intron positioning as phylogenetic information. Our results support the monophyly of Viridoplantae, Ascomycota, Homobasidiomyceta, and Apicomplexa.  相似文献   

2.
Due to the prevailing importance of the tracheal system for insect respiration, hemoglobins had been considered rare exceptions in this arthropod subphylum. Here we report the identification, cloning and expression analysis of a true hemoglobin gene in the honeybee Apis mellifera (Hymenoptera). The deduced amino acid sequence covers 171 residues (19.5kDa) and harbors all globin-typical features, including the proximal and the distal histidines. The protein has no signal peptide for transmembrane transport and was predicted to localize in the cytoplasm. The honeybee hemoglobin gene shows an ancient structure, with introns in positions B12.2 and G7.0, while most other insect globins have divergent intron positions. In situ hybridization studies showed that hemoglobin expression in the honeybee is mainly associated with the tracheal system. We also observe hemoglobin expression in the Malpighi tubes and testis. We further demonstrated that hemoglobins occur in other insect orders (Hemiptera, Coleoptera, Lepidoptera), suggesting that such genes belong to the standard repertoire of an insect genome. Phylogenetic analyses show that globins evolved along with the accepted insect systematics, with a remarkable diversification within the Diptera. Although insect hemoglobins may be in fact involved in oxygen metabolism, it remains uncertain whether they carry out a myoglobin-like function in oxygen storage and delivery.  相似文献   

3.
Sensory neuron membrane proteins (SNMPs) are membrane bound proteins initially identified in olfactory receptor neurons of Lepidoptera and are thought to play a role in odor detection; SNMPs belong to a larger gene family characterized by the human protein CD36. We have identified 12-14 candidate SNMP/CD36 homologs from each of the genomes of Drosophila melanogaster, D. pseudoobscura, Anopheles gambiae and Aedes aegypti (Diptera), eight candidate homologs from Apis mellifera (Hymenoptera), and 15 from Tribolium castaneum (Coleoptera). Analysis (sequence similarity and intron locations) suggests that the insect SNMP/CD36 genes fall into three major groups. Group 1 includes the previously characterized D. melanogaster emp (epithelial membrane protein). Group 2 includes the previously characterized D. melanogaster croquemort, ninaD, santa maria, and peste. Group 3 genes include the SNMPs, which fall into two subgroups referred to as SNMP1 and SNMP2. D. melanogaster SNMP1 (CG7000) shares both significant sequence similarity and five of its six intron insertion sites with the lepidopteran Bombyx mori SNMP1. The topological conservation of this gene family within the three major holometabolous lineages indicates that it predates the coleopteran and hymenoptera/dipera/lepidoptera split 300+ million years ago. The current state of knowledge of the characterized insect members of this gene family is discussed.  相似文献   

4.
Analysis of DNA sequences of 132 introns and 140 exons from 42 pairs of orthologous genes of mouse and rat was used to compare patterns of evolutionary change between introns and exons. The mean of the absolute difference in length (measured in base pairs) between the two species was nearly five times as high in the case of introns as in the case of exons. The average rate of nucleotide substitution in introns was very similar to the rate of synonymous substitution in exons, and both were about three times the rate of substitution at nonsynonymous sites in exons. G+C content of introns and exons of the same gene were correlated; but mean G+C content at the third positions of exons was significantly higher than that of introns or positions 1–2 of exons from the same gene. G+C content was conserved over evolutionary time, as indicated by strong correlations between mouse and rat; but the change in G+C content was greatest at position 3 of exons, intermediate in introns, and lowest at positions 1–2 in introns. Received: 23 December 1996 / Accepted: 1 April 1997  相似文献   

5.
Can Codon Usage Bias Explain Intron Phase Distributions and Exon Symmetry?   总被引:1,自引:0,他引:1  
More introns exist between codons (phase 0) than between the first and the second bases (phase 1) or between the second and the third base (phase 2) within the codon. Many explanations have been suggested for this excess of phase 0. It has, for example, been argued to reflect an ancient utility for introns in separating exons that code for separate protein modules. There may, however, be a simple, alternative explanation. Introns typically require, for correct splicing, particular nucleotides immediately 5 in exons (typically a G) and immediately 3 in the following exon (also often a G). Introns therefore tend to be found between particular nucleotide pairs (e.g., G|G pairs) in the coding sequence. If, owing to bias in usage of different codons, these pairs are especially common at phase 0, then intron phase biases may have a trivial explanation. Here we take codon usage frequencies for a variety of eukaryotes and use these to generate random sequences. We then ask about the phase of putative intron insertion sites. Importantly, in all simulated data sets intron phase distribution is biased in favor of phase 0. In many cases the bias is of the magnitude observed in real data and can be attributed to codon usage bias. It is also known that exons may carry either the same phase (symmetric) or different phases (asymmetric) at the opposite ends. We simulated a distribution of different types of exons using frequencies of introns observed in real genes assuming random combination of intron phases at the opposite sides of exons. Surprisingly the simulated pattern was quite similar to that observed. In the simulants we typically observe a prevalence of symmetric exons carrying phase 0 at both ends, which is common for eukaryotic genes. However, at least in some species, the extent of the bias in favor of symmetric (0,0) exons is not as great in simulants as in real genes. These results emphasize the need to construct a biologically relevant null model of successful intron insertion.Reviewing Editor: Dr. Manyuan Long  相似文献   

6.
7.
Yuasa HJ  Takagi T 《Gene》2001,268(1-2):17-22
Troponin C (TnC) superfamily genes essentially possess five introns, the positions of all but the fourth being highly conserved. The fourth intron is frequently absent from protostomian invertebrate genes, such as calmodulin or TnC. We previously proposed that the common ancestor of TnC superfamily genes never possessed an intron corresponding to today's fourth introns, and that members of the superfamily independently gained a fourth intron in the evolutionary pathway of each lineage. In the present study, we isolated the TnC cDNA from the sandworm, Perinereis vancaurica tetradentata and determined its genomic structure. Sandworm TnC appears to exist as a single copy gene consisting of six exons and five introns. The positions of the first, second, third and fifth introns are identical to other TnCs, but that of the fourth intron is unique. This is in good agreement with the above-mentioned scheme, i.e. the gain of the fourth intron of sandworm TnC might have occurred within the annelid lineage after annelida/mollusca divergence.  相似文献   

8.
Statistics of the exon-intron structure and splicing sites of several diverse eukaryotes was studied. The yeast exon-intron structures have a number of unique features. A yeast gene usually have at most one intron. The branch site is strongly conserved, whereas the polypirimidine tract is short. Long yeast introns tend to have stronger acceptor sites. In other species the branch site is less conserved and often cannot be determined. In non-yeast samples there is an almost universal correlation between lengths of neighboring exons (all samples excluding protists) and correlation between lengths of neighboring introns (human, drosophila, protists). On the average first introns are longer, and anomalously long introns are usually first introns in a gene. There is a universal preference for exons and exon pairs with the (total) length divisible by 3. Introns positioned between codons are preferred, whereas those positioned between the first and second positions in codon are avoided. The choice of A or G at the third position of intron (the donor splice sites generally prefer purines at this position) is correlated with the overall GC-composition of the gene. In all samples dinucleotide AG is avoided in the region preceding the acceptor site.  相似文献   

9.
Yuasa HJ  Yamamoto H  Takagi T 《Gene》1999,229(1-2):163-169
Two distinct calmodulin (CaM) genes are isolated from the ascidian, Halocynthia roretzi, (Hr-CaM A and Hr-CaM B) and those structures are determined. There are three nucleotide substitutions, producing two amino acid differences between Hr-CaM A and Hr-CaM B, and those are corresponding to two of the known eight variable residues among metazoan CaMs. Both Hr-CaM A and Hr-CaM B are constructed from six exons and five introns, and the positions of introns are identical. The positions of introns of Hr-CaMs are also identical with those of vertebrate CaMs, except third introns. The third introns of Hr-CaMs are inserted at 28bp upstream when compared with vertebrate CaMs. Thus, sliding of the third intron might have occurred in only the ascidian lineage prior to the gene duplication that also occurred only in that lineage. In addition, with the comparison of the intron positions, we attempt to investigate the vicissitude of introns during the evolution of metazoan CaMs.  相似文献   

10.
11.
We report the complete sequence of a paralogous copy of elongation factor-1 alpha (EF-1 alpha) in the honeybee, Apis mellifera (Hymenoptera: Apidae). This copy differs from a previously described copy in the positions of five introns and in 25% of the nucleotide sites in the coding regions. The existence of two paralogous copies of EF-1 alpha in Drosophila and Apis suggests that two copies of EF-1 alpha may be widespread in the holometabolous insect orders. To distinguish between a single, ancient gene duplication and parallel, independent fly and bee gene duplications, we performed a phylogenetic analysis of hexapod EF-1 alpha sequences. Unweighted parsimony analysis of nucleotide sequences suggests an ancient gene duplication event, whereas weighted parsimony analysis of nucleotides and unweighted parsimony analysis of amino acids suggests the contrary: that EF-1 alpha underwent parallel gene duplications in the Diptera and the Hymenoptera. The hypothesis of parallel gene duplication is supported both by congruence among nucleotide and amino acid data sets and by topology-dependent permutation tail probability (T-PTP) tests. The resulting tree topologies are also congruent with current views on the relationships among the holometabolous orders included in this study (Diptera, Hymenoptera, and Lepidoptera). More sequences, from diverse orders of holometabolous insects, will be needed to more accurately assess the historical patterns of gene duplication in EF-1 alpha.   相似文献   

12.
Unusual two-domain arginine kinases (AKs) arose independently at least two times during molecular evolution of phosphagen kinases: AKs from the primitive sea anemone Anthopleura japonicus and from the clam Pseudocardium sachalinensis. To elucidate its unusual evolution, the structures of Anthopleura and Pseudocardium AK genes have been determined. The Anthopleura gene consisted of 4 exons and 3 introns: two domains are linked by a bridge intron, and each domain contains one intron in different positions. On the other hand, the Pseudocardium gene consisted of 10 exons and 9 introns: two domains are also linked by a bridge intron, and domains 1 and 2 contains 3 and 5 introns, respectively, of which 3 introns are located in exactly same positions. Since the two domains of Pseudocardium AK are estimated to have diverged about 290 million years ago, the 3 introns have been conserved at least for this long. Comparison of intron positions in Anthopleura, Pseudocardium and C. elegans AK genes indicates that there is no intron conserved through the three AK lineages, in sharp contrast to relatively conservative intron positions in creatine kinase (CK) gene family.  相似文献   

13.
Nuclear protein-coding genes of euglenids (Discoba, Euglenozoa, Euglenida) contain conventional (spliceosomal) and nonconventional introns. The latter have been found only in euglenozoans. A unique feature of nonconventional introns is the ability to form a stable and slightly conserved RNA secondary structure bringing together intron ends and placing adjacent exons in proximity. To date, little is known about the mechanism of their excision (e.g. whether it involves the spliceosome or not). The tubA gene of Euglena gracilis harbors three conventional and three nonconventional introns. While the conventional introns are excised as lariats, nonconventional introns are present in the cell solely as circular RNAs with full-length ends. Based on this discovery as well as on previous observations indicating that nonconventional introns are observed frequently at unique positions of genes, we suggest that this new type of intronic circRNA might play a role in intron mobility.  相似文献   

14.
A database called eukaryotic intron database (EID) was developed based on the data from GenBank.Studies on the statistical characteristics of EID show that there were 103,848 genes,478,484 introns,and 582,332 exons,with an average of 4.61 introns and 5.61 exons per gene.Introns of 40-120 nt in length were abundant in the database.Results of the statistical analysis on the data from nine model species showed that in eukaryotes,higher species do not necessarily have more introns or exons in a gene than lower species.Furthermore,characteristics of EID,such as intron phase,distribution of different splice sites,and the relationship between genome size and intron proportion or intron density,have been studied.  相似文献   

15.
Gene structure data can substantially advance our understanding of metazoan evolution and deliver an independent approach to resolve conflicts among existing hypotheses. Here, we used changes of spliceosomal intron positions as novel phylogenetic marker to reconstruct the animal tree. This kind of data is inferred from orthologous genes containing mutually exclusive introns at pairs of sequence positions in close proximity, so-called near intron pairs (NIPs). NIP data were collected for 48 species and utilized as binary genome-level characters in maximum parsimony (MP) analyses to reconstruct deep metazoan phylogeny. All groupings that were obtained with more than 80% bootstrap support are consistent with currently supported phylogenetic hypotheses. This includes monophyletic Chordata, Vertebrata, Nematoda, Platyhelminthes and Trochozoa. Several other clades such as Deuterostomia, Protostomia, Arthropoda, Ecdysozoa, Spiralia, and Eumetazoa, however, failed to be recovered due to a few problematic taxa such as the mite Ixodes and the warty comb jelly Mnemiopsis. The corresponding unexpected branchings can be explained by the paucity of synapomorphic changes of intron positions shared between some genomes, by the sensitivity of MP analyses to long-branch attraction (LBA), and by the very unequal evolutionary rates of intron loss and intron gain during evolution of the different subclades of metazoans. In addition, we obtained an assemblage of Cnidaria, Porifera, and Placozoa as sister group of Bilateria + Ctenophora with medium support, a disputable, but remarkable result. We conclude that NIPs can be used as phylogenetic characters also within a broader phylogenetic context, given that they have emerged regularly during evolution irrespective of the large variation of intron density across metazoan genomes.  相似文献   

16.
A genome-wide analysis of innate immunity-related genes and gene families was conducted using the silkworm, Bombyx mori. We identified orthologs for a large number of genes involved in insect immunity that have been reported from Drosophila melanogaster (Diptera), Anopheles gambiae (Diptera), Apis mellifera (Hymenoptera) and Tribolium castaneum (Coleoptera). B. mori has a unique recognition gene and antimicrobial peptide genes that are not present in the Drosophila, Anopheles, Apis and Tribolium genomes, suggesting a lineage-specific gene evolution for lepidopteran insects. The comparative analysis of the insect immune repertoires indicated a dynamic and flexible gene expansion in recognition, modulation and effector mechanisms due to different selection pressures. Differential gene regulation by different bacterial species was found in PGRP and Serpin genes, suggesting that Bombyx has a highly selective gene regulation system depending on bacterial species.  相似文献   

17.
Many spliceosomal introns exist in the eukaryotic nuclear genome. Despite much research, the evolution of spliceosomal introns remains poorly understood. In this paper, we tried to gain insights into intron evolution from a novel perspective by comparing the gene structures of cytoplasmic ribosomal proteins (CRPs) and mitochondrial ribosomal proteins (MRPs), which are held to be of archaeal and bacterial origin, respectively. We analyzed 25 homologous pairs of CRP and MRP genes that together had a total of 527 intron positions. We found that all 12 of the intron positions shared by CRP and MRP genes resulted from parallel intron gains and none could be considered to be “conserved,” i.e., descendants of the same ancestor. This was supported further by the high frequency of proto-splice sites at these shared positions; proto-splice sites are proposed to be sites for intron insertion. Although we could not definitively disprove that spliceosomal introns were already present in the last universal common ancestor, our results lend more support to the idea that introns were gained late. At least, our results show that MRP genes were intronless at the time of endosymbiosis. The parallel intron gains between CRP and MRP genes accounted for 2.3% of total intron positions, which should provide a reliable estimate for future inferences of intron evolution.  相似文献   

18.
The evolution of spliceosomal introns remains poorly understood. Although many approaches have been used to infer intron evolution from the patterns of intron position conservation, the results to date have been contradictory. In this paper, we address the problem using a novel maximum likelihood method, which allows estimation of the frequency of intron insertion target sites, together with the rates of intron gain and loss. We analyzed the pattern of 10,044 introns (7,221 intron positions) in the conserved regions of 684 sets of orthologs from seven eukaryotes. We determined that there is an average of one target site per 11.86 base pairs (bp) (95% confidence interval, 9.27 to 14.39 bp). In addition, our results showed that: (i) overall intron gains are ~25% greater than intron losses, although specific patterns vary with time and lineage; (ii) parallel gains account for ~18.5% of shared intron positions; and (iii) reacquisition following loss accounts for ~0.5% of all intron positions. Our results should assist in resolving the long-standing problem of inferring the evolution of spliceosomal introns.  相似文献   

19.
20.
《环境昆虫学报》2014,(5):705-710
本研究选择西方蜜蜂、黑腹果蝇、致倦库蚊和家蚕四种昆虫为例,以降解保幼激素的特异性酯酶(juvenile hormone esterase,JHE)为研究对象,首先运用NCBI的Spidey在线软件将各昆虫的JHE基因与其相应的基因组序列作比对,分别确定各昆虫JHE基因的上游2000bp的具体序列和外显子及内含子区域。发现西方蜜蜂JHE基因含有7个外显子和6个内含子,而其余三种昆虫的JHE基因均含有6个外显子和5个内含子。接着,分别统计基因各区域的CpG,G,C位点的占有量,并计算出CpG O/E值(CpG位点的实际值与期望值之间的比值),发现各区域CpG O/E平均值的大小依次为:基因上游2000bp区>内含子区>外显子区,基因上游2000bp区的CpG O/E平均值大于1.0,而第1到第4外显子的CpG O/E平均值均只有0.8左右。表明在昆虫JHE基因中,若有CpG甲基化位点发生,应主要发生在第1到第4外显子区域。  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