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 375 毫秒
1.
Sponges (phylum Porifera) are the phylogenetic oldest Metazoa still extant. They can be considered as reference animals (Urmetazoa) for the understanding of the evolutionary processes resulting in the creation of Metazoa in general and also for the metazoan gene organization in particular. In the marine sponge Suberites domuncula, genes encoding p38 and JNK kinases contain nine and twelve introns, respectively. Eight introns in both genes share the same positions and the identical phases. One p38 intron slipped for six bases and the JNK gene has three more introns. However, the sequences of the introns are not conserved and the introns in JNK gene are generally much longer. Introns interrupt most of the conserved kinase subdomains I-XI and are found in all three phases (0, 1 and 2). We analyzed in details p38 and JNK genes from human, Caenorhabditis elegans and Drosophila melanogaster and found in most genes introns at the positions identical to those in sponge genes. The exceptions are two p38 genes from D. melanogaster that have lost all introns in the coding sequence. The positions of 11 introns in each of four human p38 genes are fully conserved and ten introns occupy identical positions as the introns in sponge p38 or JNK genes. The same is true for nine, out of ten introns in the human JNK-1 gene. The introns in human p38 and JNK genes are on average more than ten times longer than corresponding introns in sponges. It was proposed that yeast HOG1-like kinases (from i.e. Saccharomyces cerevisiae and Emericella nidulans) and metazoan p38 and JNK kinases are orthologues. p38 and JNK genes were created after the split from fungi by the duplication and diversification of the HOG1-like progenitor gene. Our results further support the common origin of p38 and JNK genes and speak in favor of a very early time of duplication. The ancestral gene contained at least ten introns, which are still present at the very conserved positions in p38 and JNK genes of extant animals. Four of these introns are present at the same positions in the HOG-like gene in the fungus E. nidulans. The others probably entered the ancestral gene after the split of fungi, but before the duplication of the gene and before the creation of the common, urmetazoan progenitor of all multicellular animals. A second gene coding for an immune molecule is described, the allograft inflammatory factor, which likewise showed a highly conserved exon/intron structure in S. domuncula and in human. These data show that the intron/exon borders are highly conserved in genes from sponges to human.  相似文献   

2.
Some of the principal transitions in the evolution of eukaryotes are characterized by engulfment of prokaryotes by primitive eukaryotic cells. In particular, approximately 1.6 billion years ago, engulfment of a cyanobacterium that became the ancestor of chloroplasts and other plastids gave rise to Plantae, the major branch of eukaryotes comprised of glaucophytes, red algae, green algae, and green plants. After endosymbiosis, there was large-scale migration of genes from the endosymbiont to the nuclear genome of the host such that approximately 18% of the nuclear genes in Arabidopsis appear to be of chloroplast origin. To gain insights into the process of evolution of gene structure in these, originally, intronless genes, we compared the properties and the evolutionary dynamics of introns in genes of plastid origin and ancestral eukaryotic genes in Arabidopsis, poplar, and rice genomes. We found that intron densities in plastid-derived genes were slightly but significantly lower than those in ancestral eukaryotic genes. Although most of the introns in both categories of genes were conserved between monocots (rice) and dicots (Arabidopsis and poplar), lineage-specific intron gain was more pronounced in plastid-derived genes than in ancestral genes, whereas there was no significant difference in the intron loss rates between the 2 classes of genes. Thus, after the transfer to the nuclear genome, the plastid-derived genes have undergone a massive intron invasion that, by the time of the divergence of dicots and monocots (150-200 MYA), yielded intron densities only slightly lower than those in ancestral genes. Nevertheless, the accumulation of introns in plastid-derived genes appears not to have reached saturation and continues to this time, albeit at a low rate. The overall pattern of intron gain and loss in the plastid-derived genes is shaped by this continuing gain and the more general tendency for loss that is characteristic of the recent evolution of plant genes.  相似文献   

3.
In order to understand the molecular evolution of catalase genes in higher plants, we compared the exon-intron structures of 12 genomic sequences from six plant species. It was assumed that the putative single primordial catalase gene had seven introns, because only those catalase genes having this structure are found in the monocotyledonae and dicotyledonae classes. After the evolutionary divergence of monocots from dicots, consecutive duplication of the primordial gene followed by the differential loss of introns occurred in each class to form three (or possibly four in dicots) diverse isozyme genes. In monocots, three ancestral isozyme genes were formed before the divergence of ancestral rice and maize. One of the rice genes, CatA, has an entirely new short intron which was not found in any other plant catalase gene examined. We have investigated the existence of the intron in the CatA homolog in other rice species by polymerase chain reaction (PCR) analysis. One major PCR product was found with the genomic DNAs from O. sativa (indica and japonica types), O. rufipogon and O. glaberrima. DNAs from several accessions of O. longistaminata showed variation in both the number and size of the DNA fragments amplified. PCR analyses and sequencing of the PCR products revealed that there are several CatA homologs having different sequences in some accessions of O. longistaminata. We have extended our study to other species in the Poaceae. The results suggest that the gain of the intron, most likely by insertion of a retroposon, took place in the ancestral genome of rice after its evolutionary divergence from other ancestral cereals such as barley, wheat and oat. Received: 20 November 1997 / Accepted: 5 January 1998  相似文献   

4.
Chromosomal linkage as well as sequence homologies provide unequivocal evidence that the genes for the alpha, beta and gamma chains of fibrinogen arose by successive duplication of a single ancestral gene. Yet, when the three fibrinogen chains are aligned by amino acid homology, the positions of intervening sequences coincide at only two positions for all three chains. While one additional intron occurs at a homologous site in the beta and gamma chains, none of the positions of the remaining 11 introns in the three genes is shared. This arrangement of introns in the three fibrinogen genes suggests that either introns were selectively lost, implying that there is essential information in the retained introns, or the common introns were present in the ancestral fibrinogen gene and introns have been randomly inserted since the triplication of the original gene. The more likely possibility of selective loss of introns implies that the ancestral gene, as it existed about one billion years ago, must have been composed of numerous small exons.  相似文献   

5.
We have found previously that the sequences important for recognition of pre-mRNA introns in dicot plants differ from those in the introns of vertebrates and yeast. Neither a conserved branch point nor a polypyrimidine tract, found in yeast and vertebrate introns respectively, are required. Instead, AU-rich sequences, a characteristic feature of dicot plant introns, are essential. Here we show that splicing in protoplasts of maize, a monocot, differs significantly from splicing in a dicot, Nicotiana plumbaginifolia. As in the case of dicots, a conserved branch point and a polypyrimidine tract are not required for intron processing in maize. However, unlike in dicots, AU-rich sequences are not essential, although their presence facilitates splicing if the splice site sequences are not optimal. The lack of an absolute requirement for AU-rich stretches in monocot introns in reflected in the occurrence of GC-rich introns in monocots but not in dicots. We also show that maize protoplasts are able to process a mammalian intron and short introns containing stem--loops, neither of which are spliced in N.plumbaginifolia protoplasts. The ability of maize, but not of N.plumbaginifolia to process stem--loop-containing or GC-rich introns suggests that one of the functions of AU-rich sequences during splicing of dicot plant pre-mRNAs may be to minimize secondary structure within the intron.  相似文献   

6.
W. Stephan  V. S. Rodriguez  B. Zhou    J. Parsch 《Genetics》1994,138(1):135-143
Three distinctly different alleles of the metallothionein gene Mtn have been identified in natural Drosophila melanogaster populations: Mtn(.3), Mtn(1), and Dp(Mtn(1)), where the latter designates a tandem duplication of Mtn(1). In Drosophila simulans, only Mtn(.3)-type alleles have been found. It has been suggested that Mtn(.3) is the ancestral allele and demonstrated that a presumed two-step transition from Mtn(.3) to Mtn(1) to Dp(Mtn(1)) is accompanied by an approximate 5-fold increase in RNA levels. We analyzed the evolutionary genetics of the Mtn locus of Drosophila ananassae, a distant relative of D. melanogaster and D. simulans within the melanogaster species group. The Mtn gene of D. ananassae is most similar to Mtn(.3). (i) it is identical with Mtn(.3) at the amino acid level, but differs from Mtn(1) in its terminal codon; (ii) its 3'' UTR contains a characteristic extra DNA segment of about 50 bp which is present in Mtn(.3), but lacking in Mtn(1); (iii) duplications of Mtn were not found in a worldwide sample of 110 wild D. ananassae chromosomes. However, the intron of the Mtn gene in D. ananassae is only 69 bp long, whereas the length of the Mtn(.3) and Mtn(1) introns is 265 bp; and it lacks a polypyrimidine stretch upstream of the 3'' splice site in contrast to the much greater pyrimidine-richness found in the Mtn(.3) and Mtn(1) introns. A short intron (67 bp) was also identified in a D. pseudoobscura Mtn allele, suggesting that the short intron is the ancestral form and that the transition from the short to the long intron occurred within the melanogaster species group. We discuss the significance of this observation with regard to the recently proposed classification of D. melanogaster introns into two groups: short introns (<90 bp) which tend to lack polypyrimidine stretches, and longer ones which have strong 3'' splice signals similar to mammalian introns. A database search revealed that this length dimorphism is an evolutionarily conserved feature of Drosophila introns; transitions from one size class to the other appear to be rare between closely related species (e.g., within the melanogaster subgroup).  相似文献   

7.
We isolated genes for hatching enzymes and their paralogs having two cysteine residues at their N-terminal regions in addition to four cysteines conserved in all the astacin family proteases. Genes for such six-cysteine-containing astacin proteases (C6AST) were searched out in the medaka genome database. Five genes for MC6AST1 to 5 were found in addition to embryo-specific hatching enzyme genes. RT-PCR and whole-mount in situ hybridization evidenced that MC6AST1 was expressed in embryos and epidermis of almost all adult tissues examined, while MC6AST2 and 3 were in mesenterium, intestine, and testis. MC6AST4 and 5 were specifically expressed in jaw. In addition, we cloned C6AST cDNA homologs from zebrafish, ayu, and fugu. The MC6AST1 to 5 genes were classified into three groups in the phylogenetic positions, and the expression patterns and hatching enzymes were clearly discriminated from other C6ASTs. Analysis of the exon–intron structures clarified that genes for hatching enzymes MHCE and MAHCE were intron-less, while other MC6AST genes were basically the same as the gene for another hatching enzyme MLCE. In the basal Teleost, the C6AST genes having the ancestral exon–intron structure (nine exon/eight intron structure) first appeared by duplication and chromosomal translocation. Thereafter, maintaining such ancestral exon–intron structure, the LCE gene was newly diversified in Euteleostei, and the MC6AST1 to 5 gene orthologs were duplicated and diversified independently in respective fish lineages. The HCE gene lost all introns in Euteleostei, whereas in the lineage to zebrafish, it was translocated from chromosome to chromosome and lost some of its introns.Electronic supplementary material Supplementary material is available in the online version of this article at and is accessible for authorized users.The nucleotide sequence data reported in the present paper will appear in the DDBJ/EMBL/GenBank nucleotide sequence databases with accession numbers from AB256940 to AB256952.  相似文献   

8.
Sequencing of eukaryotic genomes allows one to address major evolutionary problems, such as the evolution of gene structure. We compared the intron positions in 684 orthologous gene sets from 8 complete genomes of animals, plants, fungi, and protists and constructed parsimonious scenarios of evolution of the exon-intron structure for the respective genes. Approximately one-third of the introns in the malaria parasite Plasmodium falciparum are shared with at least one crown group eukaryote; this number indicates that these introns have been conserved through >1.5 billion years of evolution that separate Plasmodium from the crown group. Paradoxically, humans share many more introns with the plant Arabidopsis thaliana than with the fly or nematode. The inferred evolutionary scenario holds that the common ancestor of Plasmodium and the crown group and, especially, the common ancestor of animals, plants, and fungi had numerous introns. Most of these ancestral introns, which are retained in the genomes of vertebrates and plants, have been lost in fungi, nematodes, arthropods, and probably Plasmodium. In addition, numerous introns have been inserted into vertebrate and plant genes, whereas, in other lineages, intron gain was much less prominent.  相似文献   

9.
We examined the gene structure of a set of 2563 Arabidopsis thaliana paralogous pairs that were duplicated simultaneously 20-60 MYA by tetraploidy. Out of a total of 23,164 introns in these genes, we found that 10,004 pairs have been conserved and 578 introns have been inserted or deleted in the time since the duplication event. This intron insertion/deletion rate of 2.7 x 10(-3) to 9.1 x 10(-4) per site per million years is high in comparison to previous studies. At least 56 introns were gained and 39 lost based on parsimony analysis of the phylogenetic distribution of these introns. We found weak evidence that genes undergoing intron gain and loss are biased with respect to gene ontology terms. Gene pairs that experienced at least 2 intron insertions or deletions show evidence of enrichment for membrane location and transport and transporter activity function. We do not find any relationship of intron flux to expression level or G + C content of the gene. Detection of a bias in the location of intron gains and losses within a gene depends on the method of measurement: an intragene method indicates that events (specifically intron losses) are biased toward the 3' end of the gene. Despite the relatively recent acquisition of these introns, we found only one case where we could identify the mechanism of intron origin--the TOUCH3 gene has experienced 2 tandem, partial, internal gene duplications that duplicated a preexisting intron and also created a novel, alternatively spliced intron that makes use of a duplicated pair of cryptic splice sites.  相似文献   

10.
We have analyzed the evolution of fibroblast growth factor receptor (FGFR) tyrosine kinase genes throughout a wide range of animal phyla. No evidence for an FGFR gene was found in Porifera, but we tentatively identified an FGFR gene in the placozoan Trichoplax adhaerens. The gene encodes a protein with three immunoglobulin-like domains, a single-pass transmembrane, and a split tyrosine kinase domain. By superimposing intron positions of 20 FGFR genes from Placozoa, Cnidaria, Protostomia, and Deuterostomia over the respective protein domain structure, we identified ten ancestral introns and three conserved intron groups. Our analysis shows (1) that the position of ancestral introns correlates to the modular structure of FGFRs, (2) that the acidic domain very likely evolved in the last common ancestor of triploblasts, (3) that splicing of IgIII was enabled by a triploblast-specific insertion, and (4) that IgI is subject to substantial loss or duplication particularly in quickly evolving genomes. Moreover, intron positions in the catalytic domain of FGFRs map to the borders of protein subdomains highly conserved in other serine/threonine kinases. Nevertheless, these introns were introduced in metazoan receptor tyrosine kinases exclusively. Our data support the view that protein evolution dating back to the Cambrian explosion took place in such a short time window that only subtle changes in the domain structure are detectable in extant representatives of animal phyla. We propose that the first multidomain FGFR originated in the last common ancestor of Placozoa, Cnidaria, and Bilateria. Additional domains were introduced mainly in the ancestor of triploblasts and in the Ecdysozoa.  相似文献   

11.
Although spliceosomal introns are present in all characterized eukaryotes, intron numbers vary dramatically, from only a handful in the entire genomes of some species to nearly 10 introns per gene on average in vertebrates. For all previously studied intron-rich species, significant fractions of intron positions are shared with other widely diverged eukaryotes, indicating that 1) large numbers of the introns date to much earlier stages of eukaryotic evolution and 2) these lineages have not passed through a very intron-poor stage since early eukaryotic evolution. By the same token, among species that have lost nearly all of their ancestral introns, no species is known to harbor large numbers of more recently gained introns. These observations are consistent with the notion that intron-dense genomes have arisen only once over the course of eukaryotic evolution. Here, we report an exception to this pattern, in the intron-rich diatom Thalassiosira pseudonana. Only 8.1% of studied T. pseudonana intron positions are conserved with any of a variety of divergent eukaryotic species. This implies that T. pseudonana has both 1) lost nearly all of the numerous introns present in the diatom-apicomplexan ancestor and 2) gained a large number of new introns since that time. In addition, that so few apparently inserted T. pseudonana introns match the positions of introns in other species implies that insertion of multiple introns into homologous genic sites in eukaryotic evolution is less common than previously estimated. These results suggest the possibility that intron-rich genomes may have arisen multiple times in evolution. These results also provide evidence that multiple intron insertion into the same site is rare, further supporting the notion that early eukaryotic ancestors were very intron rich.  相似文献   

12.
Recently, a novel kinesin-like protein (KCBP) that is regulated by Ca2+/calmodulin was isolated from dicot plants. A homolog of KCBP has not been reported in monocots. To determine if this motor protein is present in phylogenetically divergent flowering plants, Arabidopsis KCBP cDNA was used as a probe to screen a genomic library of maize, an evolutionarily divergent species. This screening resulted in isolation of a KCBP homolog. Comparison of the predicted amino acid sequence of the KCBP from maize (ZmKCBP), a monocot, with the previously reported KCBP sequences from dicot species showed that the amino acid sequence, domain organization, and gene structure are highly conserved between monocots and dicots. The C-terminal region of ZmKCBP, containing the motor domain and the calmodulin-binding domain, and the N-terminal tail, with a myosin tail homology region (MyTH4) and talin-like region, showed strong sequence similarity to the KCBP homolog from dicots. However, the coiled-coil region is less conserved between monocots and dicots. The ZmKCBP gene contained 22 exons and 21 introns. The location of 19 of the 21 introns of ZmKCBP is also conserved. The ZmKCBP protein is encoded by a single gene and expressed in all tissues. Affinity-purified antibody to the calmodulin-binding domain of Arabidopsis KCBP detected a protein in both the soluble and the microsomal fractions. The C-terminal region of ZmKCBP, containing the motor and calmodulin-binding domains, bound calmodulin in the presence of calcium and failed to bind in the presence of EGTA. The ZmKCBP, along with other KCBPs from dicots, was grouped into a distinct group in the C-terminal subfamily of kinesin-like proteins. These data suggest that the KCBP is ubiquitous and highly conserved in all flowering plants and the origin of KCBP predated the divergence of monocots and dicots.  相似文献   

13.
14.

 

We report that the positions of minor, U12 introns are conserved in orthologous genes from human and Arabidopsis to an even greater extent than the positions of the major, U2 introns. The U12 introns, especially, conserved ones are concentrated in 5'-portions of plant and animal genes, where the U12 to U2 conversions occurs preferentially in the 3'-portions of genes. These results are compatible with the hypothesis that the high level of conservation of U12 intron positions and their persistence in genomes despite the unidirectional U12 to U2 conversion are explained by the role of the slowly excised U12 introns in down-regulation of gene expression.

Reviewers

This article was reviewed by John Logsdon and Manyuan Long. For the full reviews, please go to the Reviewers' Reports section.  相似文献   

15.
In this work, we investigated (1) the compositional distributions of all available nuclear coding sequences (and of their three codon positions) of six dicots and four Gramineae; this considerably expanded our knowledge about the differences previously seen between these two groups of plants; (2) the compositional correlations of homologous genes from dicots and from Gramineae, as well as from both groups; all correlations were characterized by very good coefficients, with slopes close to unity in the former two cases and very high in the last; (3) the compositional transition that accompanied the emergence of Gramineae from an ancestral monocot; (4) the compositional correlations between exons and introns, which were very good in Gramineae, but only poor to good in dicots; and (5) the compositional profiles of homologous genes from angiosperms, which were characterized by a series of peaks (exons) and valleys (introns) separated by 15–20% GC. The conservative and transitional modes of compositional evolution in plant genes and their general implications are discussed. Received: 24 June 1997 / Accepted: 20 August 1997  相似文献   

16.
The mechanisms and evolutionary dynamics of intron insertion and loss in eukaryotic genes remain poorly understood. Reconstruction of parsimonious scenarios of gene structure evolution in paralogous gene families in animals and plants revealed numerous gains and losses of introns. In all analyzed lineages, the number of acquired new introns was substantially greater than the number of lost ancestral introns. This trend held even for lineages in which vertical evolution of genes involved more intron losses than gains, suggesting that gene duplication boosts intron insertion. However, dating gene duplications and the associated intron gains and losses based on the molecular clock assumption showed that very few, if any, introns were gained during the last ~100 million years of animal and plant evolution, in agreement with previous conclusions reached through analysis of orthologous gene sets. These results are generally compatible with the emerging notion of intensive insertion and loss of introns during transitional epochs in contrast to the relative quiet of the intervening evolutionary spans.  相似文献   

17.
Are intron positions correlated with regions of high amino acid conservation? For a set of ancient conserved proteins, with intronless prokaryotic but intron-containing eukaryotic homologs, multiple sequence alignments identified residues invariant throughout evolution. Intron positions between codons show no preferences. However, introns lying after the first base of a codon prefer conserved regions, markedly in glycines. Because glycines are in excess in conserved regions, this behavior could reflect phase-one introns entering glycine residues randomly in the ancestral sequences. Examination of intron positions within codons of evolutionarily invariable amino acids showed that roughly 50% of these introns are bordered by guanines at both 5'- and 3'-ends, 25% have a G only before the intron, and 5% have a G only after the intron, whereas about 20% are bordered by nonguanine bases.  相似文献   

18.
The structural organization of the two closely related vitellogenin genes A1 and A2 has been determined and compared by electron microscopy. In both genes the mRNA-coding sequence of 6 kb is interrupted 33 times, leading to a total gene length of 21 kb for gene A1 and 16 kb for gene A2. Thus both genes have a mean exon length of 0.175 kb, while the mean intron length is 0.45 kb in gene A1 and 0.31 kb in gene A2. Because the introns interrupt the structural sequence at homologous positions in genes A1 and A2, we suggest that these two genes are the products of a duplication of an ancestral gene which had an intron-exon arrangement similar to that of the extant genes. Since the duplication event, the sequence and length of the analogous introns have changed rapidly, whereas homologous exons have diverged to an extent of only 5% of their sequences. The results suggest different mechanisms of evolution for exons and introns. While the exons evolved primarily by point mutations, such mutations, as well as deletion, insertion and duplication events, were important in the evolution of the introns.  相似文献   

19.
We identified the single gene for nuclear lamin in the genome draft of the sea anemone Nematostella vectensis, a member of the cnidaria, a very old metazoan phylum. The gene consists of 10 exons and 9 introns. Strikingly all 9 intron positions are conserved in the human lamin B genes, which have only 1 (lamin B1) or 2 (lamin B2) additional introns. Using the information on neighboring genes we propose that the human lamin B1 gene on chromosome 5 is the true homolog of the Nematostella lamin gene, while the lamin B2 gene on chromosome 19 arose during vertebrate evolution. In marked contrast to this conservation of gene structure are the results in the rapidly evolving genomes of Drosophila and Caenorhabditis elegans. Here the lamin genes have much fewer introns and these occur often at novel positions. In the single nematode lamin gene and the Drosophila lamin Dmo gene no intron position coincides with an intron in the sea anemone lamin gene.  相似文献   

20.
Two classes of genes in plants   总被引:19,自引:0,他引:19  
Carels N  Bernardi G 《Genetics》2000,154(4):1819-1825
Two classes of genes were identified in three Gramineae (maize, rice, barley) and six dicots (Arabidopsis, soybean, pea, tobacco, tomato, potato). One class, the GC-rich class, contained genes with no, or few, short introns. In contrast, the GC-poor class contained genes with numerous, long introns. The similarity of the properties of each class, as present in the genomes of maize and Arabidopsis, is particularly remarkable in view of the fact that these plants exhibit large differences in genome size, average intron size, and DNA base composition. The functional relevance of the two classes of genes is stressed by (1) the conservation in homologous genes from maize and Arabidopsis not only of the number of introns and of their positions, but also of the relative size of concatenated introns; and (2) the existence of two similar classes of genes in vertebrates; interestingly, the differences in intron sizes and numbers in genes from the GC-poor and GC-rich classes are much more striking in plants than in vertebrates.  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