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.
Recombinations are known to disrupt bifurcating tree structure of gene genealogies. Although recently occurred recombinations are easily detectable by using conventional methods, recombinations may have occurred at any time. We devised a new method for detecting ancient recombinations through phylogenetic network analysis, and detected five ancient recombinations in gibbon ABO blood group genes [Kitano et al., 2009. Mol. Phylogenet. Evol., 51, 465–471]. We present applications of this method, now named as “PNarec”, to various virus sequences as well as HLA genes.  相似文献   

2.
H Vahidi  A Purac  J M LeBlanc  B M Honda 《Gene》1991,108(2):281-284
In the plant-parasitic nematode Meloidogyne arenaria, isolated rDNA repeats show length heterogeneity, and are unusual in that they contain putative 5S ribosomal RNA pseudogenes [Vahidi et al., J. Mol. Evol. 27 (1988) 222-227]. Potentially functional 5S rRNA-encoding genes can also be identified in various rDNA repeats, which appear to be tandemly organized in the genome.  相似文献   

3.
Horizontal gene transfer in microbial genome evolution   总被引:1,自引:0,他引:1  
Horizontal gene transfer is the collective name for processes that permit the exchange of DNA among organisms of different species. Only recently has it been recognized as a significant contribution to inter-organismal gene exchange. Traditionally, it was thought that microorganisms evolved clonally, passing genes from mother to daughter cells with little or no exchange of DNA among diverse species. Studies of microbial genomes, however, have shown that genomes contain genes that are closely related to a number of different prokaryotes, sometimes to phylogenetically very distantly related ones. (Doolittle et al., 1990, J. Mol. Evol. 31, 383-388; Karlin et al., 1997, J. Bacteriol. 179, 3899-3913; Karlin et al., 1998, Annu. Rev. Genet. 32, 185-225; Lawrence and Ochman, 1998, Proc. Natl. Acad. Sci. USA 95, 9413-9417; Rivera et al., 1998, Proc. Natl. Acad. Sci. USA 95, 6239-6244; Campbell, 2000, Theor. Popul. Biol. 57 71-77; Doolittle, 2000, Sci. Am. 282, 90-95; Ochman and Jones, 2000, Embo. J. 19, 6637-6643; Boucher et al. 2001, Curr. Opin., Microbiol. 4, 285-289; Wang et al., 2001, Mol. Biol. Evol. 18, 792-800). Whereas prokaryotic and eukaryotic evolution was once reconstructed from a single 16S ribosomal RNA (rRNA) gene, the analysis of complete genomes is beginning to yield a different picture of microbial evolution, one that is wrought with the lateral movement of genes across vast phylogenetic distances. (Lane et al., 1988, Methods Enzymol. 167, 138-144; Lake and Rivera, 1996, Proc. Natl. Acad. Sci. USA 91, 2880-2881; Lake et al., 1999, Science 283, 2027-2028).  相似文献   

4.
The complete nucleotide sequence of the mitochondrial cytochrome oxidase II (COII) gene was determined for five species of the honeybee (Genus: Apis): A. andreniformis, A. cerana, A. dorsata, A. florea, and A. koschevnikovi; these were then compared to the known sequence of the A. millifera gene from Crozier et al. (1989, Mol. Biol. Evol., 6: 399-411) and the wasp Excristes roborator (Liu and Beckenbach, 1992, Mol. Phylogenet. Evol., 1:41-52). Phylogenetic relationships were derived using the parasimony methods DNAPARS and PROTPARS of Felsenstein ("PHYLIP Manual Version 3.4, "University Herbarium, Univ. of California, Berkeley). The results suggest that A. dorsata is the most ancestral species, followed by the branching of A. florea/A. andreniformis and A. koschevnikovi, and then A. mellifera and A. cerana. This inference differs from the currently accepted view that considers the A. florea/A. andreniformis line to be the most ancestral.  相似文献   

5.
Human U1 small nuclear RNA is encoded by approximately 30 gene copies. All of the U1 genes share several kilobases of essentially perfect flanking homology both upstream and downstream from the U1 coding region, but remarkably, for many U1 genes excellent flanking homology extends at least 24 kilobases upstream and 20 kilobases downstream. Class I U1 RNA pseudogenes are abundant in the human genome. These pseudogenes contain a complete but imperfect U1 coding region and possess extensive flanking homology to the true U1 genes. We mapped four class I pseudogenes by in situ hybridization to the long arm of chromosome 1, bands q12-q22, a region distinct from the site on the distal short arm of chromosome 1 to which the U1 genes have been previously mapped (Lund et al., Mol. Cell. Biol. 3:2211-2220, 1983; Naylor et al., Somat. Cell Mol. Genet. 10:307-313, 1984). We confirmed our in situ hybridization results by genomic blotting experiments with somatic cell hybrid lines with translocation products of human chromosome 1. These experiments provide further evidence that class I U1 pseudogenes and the true U1 genes are not interspersed. The results, along with those published elsewhere (Bernstein et al., Mol. Cell. Biol. 5:2159-2171, 1985), suggest that gene amplification may be responsible for the sequence homogeneity of the human U1 gene family.  相似文献   

6.
Identifying the extant sister group to the remaining angiosperms has been a subject of long debate, for which the primary currently competing hypotheses are that Amborella alone is sister or that the clade (Amborella, Nymphaeales) is sister. Both Xi et al. (Syst. Biol., 2014, 63, 919) and Goremykin et al. (Syst. Biol., 2015, 64, 879) identified Amborella as sister in concatenation‐based phylogenetic analyses of their 310 nuclear genes and 78 plastid genes, respectively. But after application of Observed Variability‐based character subsampling, both papers reported the clade (Amborella, Nymphaeales) as sister. Hence alternative character‐sampling strategies may produce highly supported yet mutually exclusive phylogenetic inferences when applied to nuclear and plastid genomic data sets. Edwards et al. (Mol. Phylogenet. Evol., 2016, 94, 447) defended Observed Variability and the (Amborella, Nymphaeales) hypothesis. In this study I respond to Edwards et al.'s (Mol. Phylogenet. Evol., 2016, 94, 447) criticisms of Simmons and Gatesy (Mol. Phylogenet. Evol., 2015, 91, 98) and use Edwards et al.'s (Mol. Phylogenet. Evol., 2016, 94, 447) and Goremykin et al.'s (Syst. Biol., 2015, 64, 879) own data to demonstrate that the best‐supported phylogenetic hypothesis is that Amborella alone is sister and that the competing evidence in favour of the (Amborella, Nymphaeales) hypothesis is caused primarily by methodological artifacts (biased character deletion by Observed Variability, MP‐EST and STAR generally not being robust to the highly divergent and mis‐rooted gene trees that were used).  相似文献   

7.
The sequence of carbamoyl phosphate synthetase I (CPSase I) cDNA and expression of the enzyme in liver of the toad Xenopus laevis are reported. CPSase I mRNA increases 6-fold when toads are exposed to high salinity for extended periods of time. The deduced 1,494-amino acid sequence of the CPSase I is homologous to other CPSases and reveals a domain structure and conserved amino acids common to other CPSases. A serine residue (S287) is present where there is a cysteine residue required for glutamine-dependent activity in CPSase Types III and II (Type I CPSases utilize only ammonia as nitrogen-donating substrate). A sequence of DNA 964 bases upstream from the ATG start codon for the CPSase I gene is also reported. Phylogenetic analysis for 30 CPSase isoforms, including X. laevis CPSase I, across a wide spectrum of phyla is reported and discussed. The results are consistent with the views that eukaryotic CPSase II as a multifunctional complex evolved from prokaryotic CPSase II and that CPSase I in terrestrial vertebrates and CPSase III in fishes arose from eukaryotic CPSase II by independent events after the divergence of plants in eukaryotic evolution.  相似文献   

8.
Jeyaprakash A  Hoy MA 《Gene》2007,391(1-2):264-274
The complete mitochondrial genome of the phytoseiid Metaseiulus occidentalis (Arthropoda: Chelicerata: Acari: Phytoseiidae) has been sequenced. It is 24,961 bp in length and contains a 14,695-bp unique region, a 345-bp triplicated region, and a 9921-bp duplicated region, in that order. The A+T content of the unique region is 76.9% and contains 11 protein coding (COI-III; ATP6-8; CytB; ND1, 2, 4, 5 and 4L), two ribosomal RNA (srRNA and lrRNA), 22 transfer RNA (tRNA) genes, and two copies of D-loop control sequence. Two genes (ND3 and 6) appear to be missing, but there is a large intergenic spacer (390 bp) present, which could contain ND3 if a different codon usage is employed. The gene order is completely different from the pattern in all other known chelicerates, including the horseshoe crab Limulus polyphemus [Lavrov et al., Mol. Biol. Evol., 2000; 17:813-824]. All the inferred tRNA genes are missing the TPsiC arm, but this arm has fused with the variable arm to generate a TV replacement loop. The duplicated region (9921 bp) contains 18 genes in the same order as in the unique region from CytB to tRNA-His, plus one copy of D-loop control sequence (311 bp) and a partial tRNA-Leu2 sequence (34 bp). The small triplicated region (345 bp) contains a D-loop control sequence (311 bp) and a partial tRNA-Leu2 sequence (34 bp). Because of these anomalies, amplifying sequences posed technical difficulties, but were accomplished by using a primer-walking strategy and increasing the AT content to 75% in the high-fidelity PCR dNTP mix. This is the first phytoseiid mitochondrial genome to be completely sequenced and the largest (25 kb) detected from the Chelicerata.  相似文献   

9.
Despite mitochondria and chloroplasts having their own genome, 99% of mitochondrial proteins (Rehling et al., Nat Rev Mol Cell Biol 5:519–530, 2004) and more than 95% of chloroplast proteins (Soll, Curr Opin Plant Biol 5:529–535, 2002) are encoded by nuclear DNA, synthesised in the cytosol and imported post-translationally. Protein targeting to these organelles depends on cytosolic targeting factors, which bind to the precursor, and then interact with membrane receptors to deliver the precursor into a translocase. The molecular chaperones Hsp70 and Hsp90 have been widely implicated in protein targeting to mitochondria and chloroplasts, and receptors capable of recognising these chaperones have been identified at the surface of both these organelles (Schlegel et al., Mol Biol Evol 24:2763–2774, 2007). The role of these chaperone receptors is not fully understood, but they have been shown to increase the efficiency of protein targeting (Young et al., Cell 112:41–50, 2003; Qbadou et al., EMBO J 25:1836–1847, 2006). Whether these receptors contribute to the specificity of targeting is less clear. A class of chaperone receptors bearing tetratricopeptide repeat domains is able to specifically bind the highly conserved C terminus of Hsp70 and/or Hsp90. Interestingly, at least of one these chaperone receptors can be found on each organelle (Schlegel et al., Mol Biol Evol 24:2763–2774, 2007), which suggests a universal role in protein targeting for these chaperone receptors. This review will investigate the role that chaperone receptors play in targeting efficiency and specificity, as well as examining recent in silico approaches to find novel chaperone receptors.  相似文献   

10.
We previously reported the sequence of a 9260-bp fragment of mitochondrial (mt) DNA of the cephalopod Loligo bleekeri [J. Sasuga et al. (1999) J. Mol. Evol. 48:692–702]. To clarify further the characteristics of Loligo mtDNA, we have sequenced an 8148-bp fragment to reveal the complete mt genome sequence. Loligo mtDNA is 17,211 bp long and possesses a standard set of metazoan mt genes. Its gene arrangement is not identical to any other metazoan mt gene arrangement reported so far. Three of the 19 noncoding regions longer than 10 bp are 515, 507, and 509 bp long, and their sequences are nearly identical, suggesting that multiplication of these noncoding regions occurred in an ancestral Loligo mt genome. Comparison of the gene arrangements of Loligo, Katharina tunicata, and Littorina saxatilis mt genomes revealed that 17 tRNA genes of the Loligo mt genome are adjacent to noncoding regions. A majority (15 tRNA genes) of their counterparts is found in two tRNA gene clusters of the Katharina mt genome. Therefore, the Loligo mt genome (17 tRNA genes) may have spread over the genome, and this may have been coupled with the multiplication of the noncoding regions. Maximum likelihood analysis of mt protein genes supports the clade Mollusca + Annelida + Brachiopoda but fails to infer the relationships among Katharina, Loligo, and three gastropod species. Received: 9 May 2001 / Accepted: 3 October 2001  相似文献   

11.
Estimation of evolutionary distances between nucleotide sequences   总被引:11,自引:0,他引:11  
A formal mathematical analysis of the substitution process in nucleotide sequence evolution was done in terms of the Markov process. By using matrix algebra theory, the theoretical foundation of Barry and Hartigan's (Stat. Sci. 2:191–210, 1987) and Lanave et al.'s (J. Mol. Evol. 20:86–93, 1984) methods was provided. Extensive computer simulation was used to compare the accuracy and effectiveness of various methods for estimating the evolutionary distance between two nucleotide sequences. It was shown that the multiparameter methods of Lanave et al.'s (J. Mol. Evol. 20:86–93, 1984), Gojobori et al.'s (J. Mol. Evol. 18:414–422, 1982), and Barry and Hartigan's (Stat. Sci. 2:191–210, 1987) are preferable to others for the purpose of phylogenetic analysis when the sequences are long. However, when sequences are short and the evolutionary distance is large, Tajima and Nei's (Mol. Biol. Evol. 1:269–285, 1984) method is superior to others.  相似文献   

12.
All three genes encoding histone H3 proteins were cloned and sequenced from Tetrahymena thermophila. Two of these genes encode a major H3 protein identical to that of T. pyriformis and 87% identical to the major H3 of vertebrates. The third gene encodes hv2, a quantitatively minor replication independent (replacement) variant. The sequence of hv2 is only 85% identical to the animal replacement variant H3.3 and is the most divergent H3 replacement variant described. Phylogenetic analysis of 73 H3 protein sequences suggests that hv2, H3.3, and the plant replacement variant H3.III evolved independently, and that H3.3 is not the ancestral H3 gene, as was previously suggested (Wells, D., Bains, W., and Kedes, L. 1986, J. Mol. Evol., 23: 224-241). These results suggest it is the replication independence and not the particular protein sequence that is important in the function of H3 replacement variants.  相似文献   

13.
The neuronal calcium sensor (NCS) proteins belong to a subfamily of the EF-hand calcium binding proteins. These proteins are primarily expressed in the nervous system and currently include more than 20 members across species [Nakayama et al., J Mol Evol 34:416-448, 1992]. Two homologues of the ncs genes, Ce-ncs-1 and Ce-ncs-2, have recently been identified in the nematode C. elegans. Here we report the cDNA sequence of a third C. elegans ncs homologue, Ce-ncs-3. We demonstrate that a null mutation in this gene caused by a large deletion in the locus does not confer a visible phenotype in C. elegans. This, in addition to the strong homology between Ce-NCS-3 and the other C. elegans NCS proteins, may indicate functional redundancy between the three genes.  相似文献   

14.
The genetic code is implemented by aminoacyl-tRNA synthetases (aaRS). These 20 enzymes are divided into two classes that, despite performing same functions, have nothing common in structure. The mystery of this striking partition of aaRSs might have been concealed in their sterically complementary modes of tRNA recognition that, as we have found recently, protect the tRNAs with complementary anticodons from confusion in translation. This finding implies that, in the beginning, life increased its coding repertoire by the pairs of complementary codons (rather than one-by-one) and used both complementary strands of genes as templates for translation. The class I and class II aaRSs may represent one of the most important examples of such primordial sense–antisense (SAS) coding (Rodin and Ohno, Orig Life Evol Biosph 25:565–589, 1995). In this report, we address the issue of SAS coding in a wider scope. We suggest a variety of advantages that such coding would have had in exploring a wider sequence space before translation became highly specific. In particular, we confirm that in Achlya klebsiana a single gene might have originally coded for an HSP70 chaperonin (class II aaRS homolog) and an NAD-specific GDH-like enzyme (class I aaRS homolog) via its sense and antisense strands. Thus, in contrast to the conclusions in Williams et al. (Mol Biol Evol 26:445–450, 2009), this could indeed be a “Rosetta stone” gene (Carter and Duax, Mol Cell 10:705–708, 2002) (eroded somewhat, though) for the SAS origin of the two aaRS classes.  相似文献   

15.
Invasive species trigger biodiversity losses and alter ecosystem functioning, with life history shaping invasiveness (Sakai et al., Annu Rev Ecol Syst 32:305–332, 2001). However, pinpointing the relation of a specific life history to invasion success is difficult. One approach may be comparing congeners. The two Palearctic pavement ants, Tetramorium sp.E (widely known as T. caespitum, Schlick-Steiner et al., Mol Phylogenet Evol 40:259–273, 2006) and T. tsushimae have invaded North America (Steiner et al., Biol Invasions 8:117–123, 2006). Their life histories differ in that T. sp.E has separate single-queened colonies but T. tsushimae multi-queened colonies scattered over large areas (Sanada-Morimura et al., Insect Soc 53:141–148, 2006; Schlick-Steiner et al., Mol Phylogenet Evol 40:259–273, 2006; Steiner et al., Biol Invasions 8:117–123, 2006). Comparison of the genetic diversity in the entire native and non-native ranges will elucidate the invasion histories. Here, we present 13 and 11 microsatellites, developed for T. sp.E and T. tsushimae, respectively, and characterize all for both species. Florian M. Steiner, Wolfgang Arthofer and Birgit C. Schlick-Steiner contributed equally to this work.  相似文献   

16.
17.
H Yokouchi  A Horii  M Emi  N Tomita  S Doi  M Ogawa  T Mori  K Matsubara 《Gene》1990,90(2):281-286
We have previously reported concerning the existence of a third type of human alpha-amylase gene, AMY3 [Emi et al., Gene 62 (1988) 229-235; Tomita et al., Gene 76 (1989) 11-18], which is expressed in a lung carcinoid tissue, and differs in nucleotide sequence from the two previously characterized human alpha-amylase genes coding for salivary and pancreatic isozymes, termed AMY1 and AMY2, respectively. Here, we rename this gene AMY2B to coincide with the designation by Gumucio et al. [Mol. Cell Biol. 8 (1988) 1197-1205] and describe its genetic properties as revealed by sequencing studies. It consists of ten major exons whose sequences are highly homologous to those of AMY1 and AMY2. Not only the exons, but also most of the introns seem to be highly conserved, as judged from physical mapping data. The AMY2B gene identified from mRNA in a lung carcinoid tissue has at least two additional untranslated exons in its 5' region; hence the promoter lies far upstream relative to the other two AMY genes.  相似文献   

18.
A series of overlapping recombinant clones, which cover the vitellogenin gene, has been isolated from a phage-lambda linked chicken gene library. The DNA of the overlapping clones spans 28 kb of contiguous DNA sequences in the chicken genome. Electron microscopic analysis of hybrids between vitellogenin mRNA and the genomic clones indicates that the chicken vitellogenin gene has a length of approximately 22 kb, about 3.8 times the size of the mRNA. The mRNA sequence is interrupted by at least 33 intervening sequences (introns). Comparison with the vitellogenin gene A2 from Xenopus laevis (Wahli et al., 1980, Cell 20: 107-117) indicates conservation of the number and length of the exons during evolution. Heteroduplex analysis reveals a short stretch of sequence homology between the genes from chicken and frog.  相似文献   

19.
Maisonhaute C  Ogereau D  Hua-Van A  Capy P 《Gene》2007,393(1-2):116-126
Transposable elements (TEs), represent a large fraction of the eukaryotic genome. In Drosophila melanogaster, about 20% of the genome corresponds to such middle repetitive DNA dispersed sequences. A fraction of TEs is composed of elements showing a retrovirus-like structure, the LTR-retrotransposons, the first TEs to be described in the Drosophila genome. Interestingly, in D. melanogaster embryonic immortal cell culture genomes the copy number of these LTR-retrotransposons was revealed to be higher than the copy number in the Drosophila genome, presumably as the result of transposition of some copies to new genomic locations [Potter, S.S., Brorein Jr., W.J., Dunsmuir, P., Rubin, G.M., 1979. Transposition of elements of the 412, copia and 297 dispersed repeated gene families in Drosophila. Cell 17, 415-427; Junakovic, N., Di Franco, C., Best-Belpomme, M., Echalier, G., 1988. On the transposition of copia-like nomadic elements in cultured Drosophila cells. Chromosoma 97, 212-218]. This suggests that so many transpositions modified the genome organisation and consequently the expression of targeted genes. To understand what has directed the transposition of TEs in Drosophila cell culture genomes, a search to identify the newly transposed copies was undertaken using 1731, a LTR-retrotransposon. A comparison between 1731 full-length elements found in the fly sequenced genome (y(1); cn(1)bw(1), sp(1) stock) and 1731 full-length elements amplified by PCR in the two cell line was done. The resulting data provide evidence that all 1731 neocopies were derived from a single copy slightly active in the Drosophila genome and subsequently strongly activated in cultured cells; and that this active copy is related to a newly evolved genomic variant (Kalmykova, A.I., et al., 2004. Selective expansion of the newly evolved genomic variants of retrotransposon 1731 in the Drosophila genomes. Mol. Biol. Evol. 21, 2281-2289). Moreover, neocopies are shown to be inserted in different sets of genes in the two cell lines suggesting they might be involved in the biological and physiological differences observed between Kc and S2 cell lines.  相似文献   

20.
We have constructed isotype-specific subclones from the 3' untranslated regions of alpha-skeletal, alpha-cardiac, beta-cytoskeletal, and gamma-cytoskeletal actin cDNAs. These clones have been used as hybridization probes to assay the number and organization of these actin isotypes in the human genome. Hybridization of these probes to human genomic actin clones (Engel et al., Proc. Natl. Acad. Sci. U.S.A. 78:4674-4678, 1981; Engel et al., Mol. Cell. Biol. 2:674-684, 1982) has allowed the unambiguous assignment of the genomic clones to isotypically defined actin subfamilies. In addition, only one isotype-specific probe hybridizes to each actin-containing gene, with a single exception. This result suggests that the multiple actin genes in the human genome are not closely linked. Genomic DNA blots probed with these subclones under stringent conditions demonstrate that the alpha-skeletal and alpha-cardiac muscle actin genes are single copy, whereas the cytoskeletal actins, beta and gamma, are present in multiple copies in the human genome. Most of the actin genes of other mammals are cytoplasmic as well. These observations have important implications for the evolution of multigene families.  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