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.
While globin genes ctt-2β and ctt-9.1 in Chironomus thummi thummi each have a single intron, all of the other insect globin genes reported so far are intronless. We analyzed four globin genes linked to the two intron-bearing genes in C. th. thummi. Three have a single intron at the same position as ctt-2β and ctt-9.1; the fourth is intronless and lies between intron bearing genes. Finally, in addition to its intron, one gene (ctt-13RT) was recently interrupted by retrotransposition. Phylogenetic analyses show that the six genes in C. th. thummi share common ancestry with five globin genes in the distantly related species C. tentans, and that a 5-gene ancestral cluster predates the divergence of the two species. One gene in the ancestral cluster gave rise to ctn-ORFB in C. tentans, and duplicated in C. th. thummi to create ctt-11 and ctt-12. From parsimonious calculations of evolutionary distances since speciation, ctt-11, ctt-12, and ctn-ORFB evolved rapidly, while ctn-ORFE in C. tentans evolved slowly compared to other globin genes in the clusters. While these four globins are under selective pressure, we suggest that most chironomid globin genes were not selected for their unique function. Instead, we propose that high gene copy number itself was selected because conditions favored organisms that could synthesize more hemoglobin. High gene copy number selection to produce more of a useful product may be the basis of forming multigene families, all of whose members initially accumulate neutral substitutions while retaining essential function. Maintenance of a large family of globin genes not only ensured high levels of hemoglobin production, but may have facilitated the extensive divergence of chironomids into as many as 5000 species. Received: 31 December 1996 / Accepted: 16 May 1997  相似文献   

2.
Base composition is not uniform across the genome of Drosophila melanogaster. Earlier analyses have suggested that there is variation in composition in D. melanogaster on both a large scale and a much smaller, within-gene, scale. Here we present analyses on 117 genes which have reliable intron/exon boundaries and no known alternative splicing. We detect significant heterogeneity in G+C content among intron segments from the same gene, as well as a significant positive correlation between the intron and the third codon position G+C content within genes. Both of these observations appear to be due, in part, to an overall decline in intron and third codon position G+C content along Drosophila genes with introns. However, there is also evidence of an increase in third codon position G+C content at the start of genes; this is particularly evident in genes without introns. This is consistent with selection acting against preferred codons at the start of genes. Received: 24 February 1997 / Accepted: 10 November 1997  相似文献   

3.
The extracellular hemoglobins of cladocerans derive from the aggregation of 12 two-domain globin subunits that are apparently encoded by four genes. This study establishes that at least some of these genes occur as a tandem array in both Daphnia magna and Daphnia exilis. The genes share a uniform structure; a bridge intron separates two globin domains which each include three exons and two introns. Introns are small, averaging just 77 bp, but a longer sequence (2.2–3.2 kb) separates adjacent globin genes. A survey of structural diversity in globin genes from other daphniids revealed three independent cases of intron loss, but exon lengths were identical, excepting a 3-bp insertion in exon 5 of Simocephalus. Heterogeneity in the extent of nucleotide divergence was marked among exons, largely as a result of the pronounced diversification of the terminal exon. This variation reflected, in part, varying exposure to concerted evolution. Conversion events were frequent in exons 1–4 but were absent from exons 5 and 6. Because of this difference, the results of phylogenetic analyses were strongly affected by the sequences employed in this construction. Phylogenies based on total nucleotide divergence in exons 1–4 revealed affinities among all genes isolated from a single species, reflecting the impact of gene conversion events. In contrast, phylogenies based on total nucleotide divergence in exons 5 and 6 revealed affinities among orthologous genes from different taxa. Received: 8 March 1999 / Accepted: 14 July 1999  相似文献   

4.
Introns are generally believed to evolve too rapidly and too erratically to be of much use in phylogenetic reconstructions. Few phylogenetically informative intron sequences are available, however, to ascertain the validity of this supposition. In the present study the supposition was tested on the example of the mammalian class II major histocompatibility complex (Mhc) genes of the DRB family. Since the Mhc genes evolve under balancing selection and are believed to recombine or rearrange frequently, the evolution of their introns could be expected to be particularly rapid and subject to scrambling. Sequences of intron 4 and 5 DRB genes were obtained from polymerase chain reaction-amplified fragments of genomic DNA from representatives of six eutherian orders—Primates, Scandentia, Chiroptera, Dermoptera, Lagomorpha, and Insectivora. Although short stretches of the introns have indeed proved to be unalignable, the bulk of the intron sequences from all six orders, spanning >85 million years (my) of evolution, could be aligned and used in a study of the tempo and mode of intron evolution. The analysis has revealed the Mhc introns to evolve at a rate similar to that of other genes and of synonymous sites of non-Mhc genes. No evidence of homogenization or large-scale scrambling of the intron sequences could be found. The Mhc introns apparently evolve largely by point mutations and insertions/deletions. The phylogenetic signals contained in the intron sequences could be used to identify Scandentia as the sister group of Primates, to support the existence of the Archonta superorder, and to confirm the monophyly of the Chiroptera. Received: 26 October 1998 / Accepted: 21 December 1998  相似文献   

5.
Vertebrate and many invertebrate globin genes have a three-exon/two-intron organization, with introns in highly conserved positions. According to the ``intron early' hypothesis, introns are the vestigial segments which flank previously independent coding sequences, thus providing evidence for the assembly of the ancient proteins by ``exon shuffling.' In this paper, we report the analysis of the genes of the bivalve mollusk Scapharca inaequivalvis tetrameric hemoglobin (HbII), which support this hypothesis, at least for the hemoglobin genes. We show the existence of ``minigenes' in the IIA and IIB globin genes, spanning part of the first and second introns, ``in frame' with the heme-binding domain coded by the second exon. Further support for the exon shuffling hypothesis can be found in the degree of identity of the ``new' translated sequences with those flanking the central protein domain of some invertebrate hemoglobins. Received: 31 July 1997 / Accepted: 12 December 1997  相似文献   

6.
TheArtemia hemoglobin contains two sub-units that are similar or different chains of nine globin domains. The domains are ancestrally related and are presumed to be derived from copies of an original single-domain parent gene. Since the gene copies have remained in the same environment for several hundred million years they provide an excellent model for the investigation of intron stability. The cDNA for one of the two types of nine-domain subunit (domains T1–T9) has been sequenced. Comparison with the corresponding genomic DNA reveals a total of 17 intradomain introns. Fourteen of the introns are in locations on the protein that are conventional in globins of other species. In eight of the nine domains an intron corresponds to the B helix, amino acid B12, following the second nucleotide (phase 2), and in six domains a G-helix intron is located between G6 and G7 (phase 0). The consistency of this pattern is supportive of the introns having been inherited from a single-domain parent gene. The remaining three introns are in unconventional locations. Two occur in the F helix, either in amino acid F3 (phase 1) in domain T3, or between F2 and F3 (phase 0) in domain T6. The two F introns strengthen an interpretation of intron inheritance since globin F introns are rare, and in domains T3 and T6 they replace rather than supplement the conventional G introns, as though displacement from G to F occurred before that part of the gene became duplicated. It is inferred that one of the F introns subsequently moved by one nucleotide. Similarly, the third unconventional intron location is the G intron in domain T4 which is in G6, phase 2, one nucleotide earlier than the other G introns. Domain T4 is also unusual in lacking a B intron. The pattern of introns in theArtemia globin gene supports a concept of general positional stability but the exceptions, where introns have moved out of reading frame, or have moved by several codons, or have been deleted, suggest that intron displacements can occur after inheritance from an ancient source. Correspondence to: C.N.A. Trotman  相似文献   

7.
Studies of the distribution of the three group I introns (intron A, intron T, and intron AT) in the 26S rDNA of Gaeumannomyces graminis had suggested that they were transferred to a common ancestor of G. graminis var. avenae and var. tritici after it had branched off from var. graminis. Intron AT and intron A exhibited vertical inheritance and coevolved in concert with their hosts. Intron loss could occur after its acquisition. Loss of any one of the three introns could occur in var. tritici whereas only loss of intron T had been found in the majority of var. avenae isolates. The existence of isolates of var. tritici and var. avenae with three introns suggested that intron loss could be reversed by intron acquisition and that the whole process is a dynamic one. This process of intron acquisition and intron loss reached different equilibrium points for different varieties and subgroups, which explained the irregular distribution of these introns in G. graminis. Each of the three group I introns was more closely related to other intron sequences that share the same insertion point in the 26S rDNA than to each other. These introns in distantly related organisms appeared to have a common ancestry. This system had provided a good model for studies on both the lateral transfer and common ancestry of group I introns in the 26S rRNA genes. Received: 17 May 1996 / Accepted: 14 January 1997  相似文献   

8.
The sequences of the entire blue opsin gene in the squirrel monkey (Saimiri boliviensis) and the five introns of the human blue opsin gene were obtained. Intron 3 of these genes contains an Alu sequence and intron 4 contains a partial mer13 sequence. A comparison of the squirrel monkey opsin sequence with published mammalian opsin sequences shows that features believed to be functionally critical are all conserved. However, the blue opsin has evolved twice as fast as rhodopsin and is only as conservative as the β globin, which has evolved at the average rate of mammalian proteins. Interestingly, the interhelical loops are, on average, actually more conservative than the transmembrane α helical regions. The introns of the blue opsin gene have evolved at the average rate of introns in primate genes. Received: 5 August 1996 / Accepted: 2 October 1996  相似文献   

9.
While the two amylase genes of Drosophila melanogaster are intronless, the three genes of D. pseudoobscura harbor a short intron. This raises the question of the common structure of the Amy gene in Drosophila species. We have investigated the presence or absence of an intron in the amylase genes of 150 species of Drosophilids. Using polymerase chain reaction (PCR), we have amplified a region that surrounds the intron site reported in D. pseudoobscura and a few other species. The results revealed that most species contain an intron, with a variable size ranging from 50 to 750 bp, although the very majoritary size was around 60–80 bp. Several species belonging to different lineages were found to lack an intron. This loss of intervening sequence was likely due to evolutionarily independent and rather frequent events. Some other species had both types of genes: In the obscura group, and to a lesser extent in the ananassae subgroup, intronless copies had much diverged from intron-containing genes. Base composition of short introns was found to be variable and correlated with that of the surrounding exons, whereas long introns were all A-T rich. We have extended our study to non-Drosophilid insects. In species from other orders of Holometaboles, Lepidoptera and Hymenoptera, an intron was found at an identical position in the Amy gene, suggesting that the intron was ancestral. Received: 23 October 1995 / Accepted: 5 March 1996  相似文献   

10.
The 22,704-bp circular mitochondrial DNA (mtDNA) of the chlamydomonad alga Chlorogonium elongatum was completely cloned and sequenced. The genome encodes seven proteins of the respiratory electron transport chain, subunit 1 of the cytochrome oxidase complex (cox1), apocytochrome b (cob), five subunits of the NADH dehydrogenase complex (nad1, nad2, nad4, nad5, and nad6), a set of three tRNAs (Q, W, M), and the large (LSU)- and small (SSU)-subunit ribosomal RNAs. Six group-I introns were found, two each in the cox1, cob, and nad5 genes. In each intron an open reading frame (ORF) related to maturases or endonucleases was identified. Both the LSU and the SSU rRNA genes are split into fragments intermingled with each other and with other genes. Although the average A + T content is 62.2%, GC-rich clusters were detected in intergenic regions, in variable domains of the rRNA genes, and in introns and intron-encoded ORFs. A comparison of the genome maps reveals that C. elongatum and Chlamydomonas eugametos mtDNAs are more closely related to one another than either is to Chlamydomonas reinhardtii mtDNA. Received: 3 November 1997 / Accepted: 12 January 1998  相似文献   

11.
Ferritin, a protein widespread in nature, concentrates iron ∼1011–1012-fold above the solubility within a spherical shell of 24 subunits; it derives in plants and animals from a common ancestor (based on sequence) but displays a cytoplasmic location in animals compared to the plastid in contemporary plants. Ferritin gene regulation in plants and animals is altered by development, hormones, and excess iron; iron signals target DNA in plants but mRNA in animals. Evolution has thus conserved the two end points of ferritin gene expression, the physiological signals and the protein structure, while allowing some divergence of the genetic mechanisms. Comparison of ferritin gene organization in plants and animals, made possible by the cloning of a dicot (soybean) ferritin gene presented here and the recent cloning of two monocot (maize) ferritin genes, shows evolutionary divergence in ferritin gene organization between plants and animals but conservation among plants or among animals; divergence in the genetic mechanism for iron regulation is reflected by the absence in all three plant genes of the IRE, a highly conserved, noncoding sequence in vertebrate animal ferritin mRNA. In plant ferritin genes, the number of introns (n= 7) is higher than in animals (n= 3). Second, no intron positions are conserved when ferritin genes of plants and animals are compared, although all ferritin gene introns are in the coding region; within kingdoms, the intron positions in ferritin genes are conserved. Finally, secondary protein structure has no apparent relationship to intron/exon boundaries in plant ferritin genes, whereas in animal ferritin genes the correspondence is high. The structural differences in introns/exons among phylogenetically related ferritin coding sequences and the high conservation of the gene structure within plant or animal kingdoms suggest that kingdom-specific functional constraints may exist to maintain a particular intron/exon pattern within ferritin genes. In the case of plants, where ferritin gene intron placement is unrelated to triplet codons or protein structure, and where ferritin is targeted to the plastid, the selection pressure on gene organization may relate to RNA function and plastid/nuclear signaling. Received: 25 July 1995 / Accepted: 3 October 1995  相似文献   

12.
We report sequences for nuclear lamins from the teleost fish Danio and six invertebrates. These include two cnidarians (Hydra and Tealia), one priapulid, two echinoderms, and the cephalochordate Branchiostoma. Combining these results with earlier data on Drosophila, Caenorhabditis elegans, and various vertebrates, the following conclusions on lamin evolution can be drawn. First, all invertebrate lamins resemble in size the vertebrate B-type lamin. Second, all lamins described previously for amphibia, birds and mammals as well as the first lamin of a fish, characterized here, show a cluster of 7 to 12 acidic residues in the tail domain. Since this acidic cluster is absent from all invertebrate lamins including that of the cephalochordate Branchiostoma, it was acquired with the vertebrate lineage. The larger A-type lamin of differentiated cells must have arisen subsequently by gene duplication and insertion of an extra exon. This extra exon of the vertebrate A-lamins is the only major change in domain organization in metazoan lamin evolution. Third, the three introns of the Hydra and Priapulus genes correspond in position to the last three introns of vertebrate B-type lamin genes. Thus the entirely different gene organization of the C. elegans and Drosophila Dmo genes seems to reflect evolutionary drift, which probably also accounts for the fact that C. elegans has the most diverse lamin sequence. Finally we discuss the possibility that two lamin types, a constitutively expressed one and a developmentally regulated one, arose independently on the arthropod and vertebrate lineages. Received: 4 February 1999 / Accepted: 1 April 1999  相似文献   

13.
Many genes for calmodulin-like domain protein kinases (CDPKs) have been identified in plants and Alveolate protists. To study the molecular evolution of the CDPK gene family, we performed a phylogenetic analysis of CDPK genomic sequences. Analysis of introns supports the phylogenetic analysis; CDPK genes with similar intron/exon structure are grouped together on the phylogenetic tree. Conserved introns support a monophyletic origin for plant CDPKs, CDPK-related kinases, and phosphoenolpyruvate carboxylase kinases. Plant CDPKs divide into two major branches. Plant CDPK genes on one branch share common intron positions with protist CDPK genes. The introns shared between protist and plant CDPKs presumably originated before the divergence of plants from Alveolates. Additionally, the calmodulin-like domains of protist CDPKs have intron positions in common with animal and fungal calmodulin genes. These results, together with the presence of a highly conserved phase zero intron located precisely at the beginning of the calmodulin-like domain, suggest that the ancestral CDPK gene could have originated from the fusion of protein kinase and calmodulin genes facilitated by recombination of ancient introns. Received: 11 July 2000 / Accepted: 18 April 2001  相似文献   

14.
The ascomycetous fungus Cryptendoxyla hypophloia contains an insertion of 433 base pairs in the genes encoding nuclear small subunit ribosomal RNA. Secondary structure analyses of the insert reveal characteristics indicative of a Group I intron, including elements P, Q, R, and S; however, the sequences of these conserved regions deviate significantly from recognized consensus sequences for Group I introns. Principal-components analysis, based on 79 nucleotide positions from the conserved core sequences of 93 Group I introns, identified 17 introns similar to that of C. hypophloia. This grouping, which includes inserts from phylogenetically diverse organisms, cannot readily be classified in any previously recognized major group of Group I introns. We propose the creation of a new group, IE, to accommodate these sequences, and discuss the evolutionary relationships between group IE and other major groups of Group I introns. Received: 11 January 1998 / Accepted: 12 October 1998  相似文献   

15.
We have determined the genomic structure of an integrin β-subunit gene from the coral, Acropora millepora. The coding region of the gene contains 26 introns, spaced relatively uniformly, and this is significantly more than have been found in any integrin β-subunit genes from higher animals. Twenty-five of the 26 coral introns are also found in a β-subunit gene from at least one other phylum, indicating that the coral introns are ancestral. While there are some suggestions of intron gain or sliding, the predominant theme seen in the homologues from higher animals is extensive intron loss. The coral baseline allows one to infer that a number of introns found in only one phylum of higher animals result from frequent intron loss, as opposed to the seemingly more parsimonious alternative of isolated intron gain. The patterns of intron loss confirm results from protein sequences that most of the vertebrate genes, with the exception of β4, belong to one of two β subunit families. The similarity of the patterns within each of the β1,2,7 and β3,5,6,8 groups indicates that these gene structures have been very stable since early vertebrate evolution. Intron loss has been more extensive in the invertebrate genes, and obvious patterns have yet to emerge in this more limited data set. Received: 5 March 2001 / Accepted: 17 May 2001  相似文献   

16.
Large dsDNA-containing chlorella viruses encode a pyrimidine dimer-specific glycosylase (PDG) that initiates repair of UV-induced pyrimidine dimers. The PDG enzyme is a homologue of the bacteriophage T4-encoded endonuclease V. The pdg gene was cloned and sequenced from 42 chlorella viruses isolated over a 12-year period from diverse geographic regions. Surprisingly, the pdg gene from 15 of these 42 viruses contain a 98-nucleotide intron that is 100% conserved among the viruses and another 4 viruses contain an 81-nucleotide intron, in the same position, that is nearly 100% identical (one virus differed by one base). In contrast, the nucleotides in the pdg coding regions (exons) from the intron-containing viruses are 84 to 100% identical. The introns in the pdg gene have 5′-AG/GTATGT and 3′-TTGCAG/AA splice site sequences which are characteristic of nuclear-located, spliceosomal processed pre-mRNA introns. The 100% identity of the 98-nucleotide intron sequence in the 15 viruses and the near-perfect identity of an 81-nucleotide intron sequence in another 4 viruses imply strong selective pressure to maintain the DNA sequence of the intron when it is in the pdg gene. However, the ability of intron-plus and intron-minus viruses to repair UV-damaged DNA in the dark was nearly identical. These findings contradict the widely accepted dogma that intron sequences are more variable than exon sequences. Received: 13 May 1999 / Accepted: 20 August 1999  相似文献   

17.
The intron–genome size relationship was studied across a wide evolutionary range (from slime mold and yeast to human and maize), as well as the relationship between genome size and the ratio of intervening/coding sequence size. The average intron size is scaled to genome size with a slope of about one-fourth for the log-transformed values; i.e., on the global scale its increase in evolution is lower than the increase in genome size by four orders of magnitude. There are exceptions to the general trend. In baker's yeast introns are extraordinarily long for its genome size. Tetrapods also have longer introns than expected for their genome sizes. In teleost fish the mean intron size does not differ significantly, notwithstanding the differences in genome size. In contrast to previous reports, avian introns were not found to be significantly shorter than introns of mammals, although avian genomes are smaller than genomes of mammals on average by about a factor of 2.5. The extra-/intragenic ratio of noncoding DNA can be higher in fungi than in animals, notwithstanding the smaller fungal genomes. In vertebrates and invertebrates taken separately, this ratio is increasing as the increase in genome size. Two hypotheses are proposed to explain the variation in the extra-/intragenic ratio of noncoding DNA in organisms with similar numbers of genes: transition (dynamic) and equilibrium (static). According to the transition model, this variation arises with the rapid shift of genome size because the bulk of extragenic DNA can be changed more rapidly than the finely interspersed intron sequences. The equilibrium model assumes that this variation is a result of selective adjustment of genome size with constraints imposed on the intron size due to its putative link to chromatin structure (and constraints of the splicing machinery). Received: 23 October 1997 / Accepted: 14 April 1999  相似文献   

18.
19.
20.
Invertebrates, tetrapod vertebrates, and fish might be expected to differ in their number of gene copies, possibly due the occurrence of genome duplication events during animal evolution. Reggie (flotillin) genes code for membrane-associated proteins involved in growth signaling in developing and regenerating axons. Until now, there appeared to be only two reggie genes in fruitflies, mammals, and fish. The aim of this research was to search for additional copies of reggie genes in fishes, since a genome duplication might have increased the gene copy number in this group. We report the presence of up to four distinct reggie genes (two reggie-1 and two reggie-2 genes) in the genomes of zebrafish and goldfish. Phylogenetic analyses show that the zebrafish and goldfish sequence pairs are orthologous, and that the additional copies could have arisen through a genome duplication in a common ancestor of bony fish. The presence of novel reggie mRNAs in fish embryos indicates that the newly discovered gene copies are transcribed and possibly expressed in the developing and regenerating nervous system. The intron/exon boundaries of the new fish genes characterized here correspond with those of human genes, both in location and phase. An evolutionary scenario for the evolution of reggie intron-exon structure, where loss of introns appears to be a distinctive trait in invertebrate reggie genes, is presented. Received: 24 January 2001 / Accepted: 27 July 2001  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