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 31 毫秒
1.
2.

Background  

Compactness of highly/broadly expressed genes in human has been explained as selection for efficiency, regional mutation biases or genomic design. However, highly expressed genes in flowering plants were shown to be less compact than lowly expressed ones. On the other hand, opposite facts have also been documented that pollen-expressed Arabidopsis genes tend to contain shorter introns and highly expressed moss genes are compact. This issue is important because it provides a chance to compare the selectionism and the neutralism views about genome evolution. Furthermore, this issue also helps to understand the fates of introns, from the angle of gene expression.  相似文献   

3.
4.
In plants, highly expressed genes are the least compact   总被引:1,自引:0,他引:1  
  相似文献   

5.
The Tact1 and Tact2 genes, each of which encodes an actin-like protein, are exclusively expressed and translated in haploid germ cells in testis. To characterize the haploid germ cell-specific gene structure, a mouse genomic library was screened with a Tact1 cDNA as a probe, and four independent phage clones containing the Tact1 gene were isolated. Southern hybridization and sequencing analyses revealed that Tact1 and Tact2 were single copy genes contained on a common fragment in a head-to-head orientation, and that the distance between these genes was less than 2 kb. Comparison of the nucleotide sequences of genomic DNA and cDNA demonstrated that Tact1 and Tact2 lack introns, although all known actin or actin-related genes in mammals contain introns. Human Tact orthologues also lack introns and are located within 6.4 kb in a head-to-head orientation. These findings indicate that Tact1 and Tact2 or one of these genes arose by retroposition of a spliced mRNA transcribed from an actin progenitor gene prior to the divergence of rodents and primates. The Tact1 and Tact2 genes are unusual retroposons in that they have retained an open reading frame and are expressed in testicular germ cells, because almost all retroposons become pseudogenes. It was revealed that a 2kb sequence between the two genes bidirectionally controls haploid germ-cell specific expression by analyzing transgenic mice. Comparison of the murine Tact genes with their human orthologues showed a high level of identity between the two species in the 5'-upstream and non-coding sequences as well as in the coding region, indicating that conserved elements in these regions may be involved in the regulation of haploid germ cell-specific expression. The promoter region contains no TATA-, CCAAT- or GC-boxes, although there are potential cAMP response element (CRE)-like motifs in the 5'-upstream region and the 5'-untranslated region in Tact1 and Tact2, respectively. Transient promoter analyses indicate that CREMtau may activate Tact1 and Tact2 expression in germ cells.  相似文献   

6.
The AKin10 gene from Arabidopsis thaliana encoding a putative Ser/Thr protein kinase (PK) has been isolated and characterized. The AKin10-encoding gene is located on a genomic 5.4-kb BamHI fragment and contains ten introns, one being located in the 5' untranslated region. The deduced amino acid sequence of AKin10 is 65% identical over the catalytic domain to the yeast PK (SNF1). SNF1 is essential for the derepression of many glucose-repressible genes, including Suc2 which encodes invertase. Southern blot hybridization experiments suggested the presence of one copy of the gene per haploid genome of A. thaliana. Northern hybridization experiments indicated that this gene is expressed in roots, shoots and leaves. AKin10 may play an important role in a signal transduction cascade regulating gene expression and carbohydrate metabolism in higher plants.  相似文献   

7.
By cross-hybridization with a cDNA probe for the Xenopus laevis ribosomal protein L1 we have been able to isolate the homologous genes from a Saccharomyces cerevisiae genomic library. We have shown that these genes code for a ribosomal protein which was previously named L2. In yeast, like in X. laevis, these genes are present in two copies per haploid genome and, unlike the vertebrate counterpart, they do not contain introns. Amino acid comparison of the X. laevis L1 and S. cerevisiae L2 proteins has shown the presence of a highly conserved protein domain embedded in very divergent sequences. Although these sequences are very poorly homologous, they confer an overall secondary structure and folding highly conserved in the two species.  相似文献   

8.
9.
10.
11.
12.
The relative rate of evolution for sex‐biased genes has often been used as a measure of the strength of sex‐specific selection. In contrast to studies in a wide variety of animals, far less is known about the molecular evolution of sex‐biased genes in plants, particularly in dioecious angiosperms. Here, we investigate the gene expression patterns and evolution of sex‐biased genes in the dioecious plant Salix viminalis. We observe lower rates of sequence evolution for male‐biased genes expressed in the reproductive tissue compared to unbiased and female‐biased genes. These results could be partially explained by the lower codon usage bias for male‐biased genes leading to elevated rates of synonymous substitutions compared to unbiased genes. However, the stronger haploid selection in the reproductive tissue of plants, together with pollen competition, would also lead to higher levels of purifying selection acting to remove deleterious variation. Future work should focus on the differential evolution of haploid‐ and diploid‐specific genes to understand the selective dynamics acting on these loci.  相似文献   

13.
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   

14.
15.
Abstract

In recent years a number of experimental findings have indicated that in higher plants the gametophytic phase is able to express its own genetic information, a large part of which it shares with the sporophytic generation. Quantitative estimates of haploid and haplodiploid gene expression have been obtained by mRNA and isozyme analysis in several plant species: 60-70% of the genes are expressed in both pollen and plant, about 10% are pollen-specific, and 20% represent the sporophytic domain. Moreover, it has been demonstrated that stage-specific genes are expressed in the gametophytic generation: at least two sets of genes are activated during pollen development, others are expressed only in the postshedding period, during germination and tube growth. Studies have been made to ascertain the role played by gametophyte-expressed genes in pollen development; the in vivo and in vitro pollen tube growth rate has been revealed to be controlled by the gametophyte genome itself. Differential effects of specific chromosomal deficiencies on the development of maize pollen grains have indicated that components of normal microspore development are controlled by genes located in specific parts of the genome. For single gene analysis, gene transfer can be used; on the contrary, for traits with a multifactorial genetic control, direct proof of gene expression both in the gametophytic and the sporophytic generation can be obtained when selection is applied to the pollen population of a hybrid plant, and response to selection is observed in the resulting sporophytic progeny. Response to selection, applied at different stages of the gametophytic phase, has been described in the sporophytic progeny and this with regard to many adaptive traits; thus the phenomenon can have an important bearing on the genetic structure of natural populations and on higher plant evolution, it can also be used as a breeding tool to increase the efficiency of conventional selection methods.  相似文献   

16.
The ribosomal protein genes are present in two to four copies per haploid genome of Xenopus laevis. Using cloned complementary DNA probes, we have isolated, from a genomic library of X. laevis, several clones containing genes for two different ribosomal proteins (L1 and L14). These genes contain intervening sequences. In the case of the L1 gene, the exons are 100 to 200 base-pairs long and the introns, on average, 400 base-pairs. Along the genomic fragments, two different classes of repetitive DNA are present: highly and middle repetitive DNA. Both are evolutionarily unstable as shown by hybridization to Xenopus tropicalis DNA. Several introns of the gene coding for protein L1 contain middle repetitive sequences. Hybridization and hybrid-released translation experiments have shown that sequences inside the two genes hybridize to several poly(A) messenger RNAs. Some of the products encoded by these mRNA have electrophoretic properties of ribosomal proteins.  相似文献   

17.
18.
19.
The primary structures of a cDNA and the genomic DNA of a gene selectively expressed in chronic lymphocytic leukemia were determined. A computer search of the nucleotide sequence data bank identified this gene as the invariant gamma-chain associated with class II histocompatibility antigens. The invariant gamma-chain genomic sequence spans about 11 kilobases, with eight exons and seven introns. Three of the introns contain members of the Alu repeat family. A putative cap site and promoter sequence were identified at the 5' end of the gene. One or two copies of the gene is present in each haploid genome, and no evidence for amplification or polymorphism was obtained.  相似文献   

20.
Analysis of evolution of exon-intron structure of eukaryotic genes   总被引:10,自引:0,他引:10  
The availability of multiple, complete eukaryotic genome sequences allows one to address many fundamental evolutionary questions on genome scale. One such important, long-standing problem is evolution of exon-intron structure of eukaryotic genes. Analysis of orthologous genes from completely sequenced genomes revealed numerous shared intron positions in orthologous genes from animals and plants and even between animals, plants and protists. The data on shared and lineage-specific intron positions were used as the starting point for evolutionary reconstruction with parsimony and maximum-likelihood approaches. Parsimony methods produce reconstructions with intron-rich ancestors but also infer lineage-specific, in many cases, high levels of intron loss and gain. Different probabilistic models gave opposite results, apparently depending on model parameters and assumptions, from domination of intron loss, with extremely intron-rich ancestors, to dramatic excess of gains, to the point of denying any true conservation of intron positions among deep eukaryotic lineages. Development of models with adequate, realistic parameters and assumptions seems to be crucial for obtaining more definitive estimates of intron gain and loss in different eukaryotic lineages. Many shared intron positions were detected in ancestral eukaryotic paralogues which evolved by duplication prior to the divergence of extant eukaryotic lineages. These findings indicate that numerous introns were present in eukaryotic genes already at the earliest stages of evolution of eukaryotes and are compatible with the hypothesis that the original, catastrophic intron invasion accompanied the emergence of the eukaryotic cells. Comparison of various features of old and younger introns starts shedding light on probable mechanisms of intron insertion, indicating that propagation of old introns is unlikely to be a major mechanism for origin of new ones. The existence and structure of ancestral protosplice sites were addressed by examining the context of introns inserted within codons that encode amino acids conserved in all eukaryotes and, accordingly, are not subject to selection for splicing efficiency. It was shown that introns indeed predominantly insert into or are fixed in specific protosplice sites which have the consensus sequence (A/C)AG|Gt.  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