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.
We sequenced to completion the circular plastid genome of the red alga Gracilaria tenuistipitata var. liui. This is the first plastid genome sequence from the subclass Florideophycidae (Rhodophyta). The genome is composed of 183,883 bp and contains 238 predicted genes, including a single copy of the ribosomal RNA operon. Comparisons with the plastid genome of Porphyra pupurea reveal strong conservation of gene content and order, but we found major genomic rearrangements and the presence of coding regions that are specific to Gracilaria. Phylogenetic analysis of a data set of 41 concatenated proteins from 23 plastid and two cyanobacterial genomes support red algal plastid monophyly and a specific evolutionary relationship between the Florideophycidae and the Bangiales. Gracilaria maintains a surprisingly ancient gene content in its plastid genome and, together with other Rhodophyta, contains the most complete repertoire of plastid genes known in photosynthetic eukaryotes.Supplementary material () is available for this article.[Reviewing Editor: Dr. W. Ford Doolittle]  相似文献   

2.
A C-to-U RNA editing event creates a functional initiation codon for translation of the psbL mRNA in tobacco plastids. Small trans-acting guide RNAs (gRNAs) have been shown to be involved in editing site selection in kinetoplastid mitochondria. A computer search of the tobacco plastid genome (ptDNA) identified such a putative gRNA, a 14-nucleotide sequence motif that is complementary to the psbL mRNA, including the A nucleotide required to direct the C-to-U change. The critical A nucleotide of the putative gRNA gene was changed to G by plastid transformation. We report here that the introduced mutation did not abolish psbL editing. Since no other region of the plastid genome contains significant complementarity to the psbL editing site we suggest that, if gRNAs serve as trans-acting factors for plastid psbL mRNA editing, they either have only a limited complementarity to the editing site, or are encoded in the nuclear genome.  相似文献   

3.
Jo YD  Park J  Kim J  Song W  Hur CG  Lee YH  Kang BC 《Plant cell reports》2011,30(2):217-229
Plants in the family Solanaceae are used as model systems in comparative and evolutionary genomics. The complete chloroplast genomes of seven solanaceous species have been sequenced, including tobacco, potato and tomato, but not peppers. We analyzed the complete chloroplast genome sequence of the hot pepper, Capsicum annuum. The pepper chloroplast genome was 156,781 bp in length, including a pair of inverted repeats (IR) of 25,783 bp. The content and the order of 133 genes in the pepper chloroplast genome were identical to those of other solanaceous plastomes. To characterize pepper plastome sequence, we performed comparative analysis using complete plastome sequences of pepper and seven solanaceous plastomes. Frequency and contents of large indels and tandem repeat sequences and distribution pattern of genome-wide sequence variations were investigated. In addition, a phylogenetic analysis using concatenated alignments of coding sequences was performed to determine evolutionary position of pepper in Solanaceae. Our results revealed two distinct features of pepper plastome compared to other solanaceous plastomes. Firstly, large indels, including insertions on accD and rpl20 gene sequences, were predominantly detected in the pepper plastome compared to other solanaceous plastomes. Secondly, tandem repeat sequences were particularly frequent in the pepper plastome. Taken together, our study represents unique features of evolution of pepper plastome among solanaceous plastomes.  相似文献   

4.
The transfer and integration of tRNA genes from organellar genomes to the nuclear genome and between organellar genomes occur extensively in flowering plants. The routes of the genetic materials flowing from one genome to another are biased, limited largely by compatibility of DNA replication and repair systems differing among the organelles and nucleus. After thoroughly surveying the tRNA gene transfer among organellar genomes and the nuclear genome of a domesticated rice (Oryza sativa L. ssp. indica), we found that (i) 15 mitochondrial tRNA genes originate from the plastid; (ii) 43 and 80 nuclear tRNA genes are mitochondrion-like and plastid-like, respectively; and (iii) 32 nuclear tRNA genes have both mitochondrial and plastid counterparts. Besides the native (or genuine) tRNA gene sets, the nuclear genome contains organelle-like tRNA genes that make up a complete set of tRNA species capable of transferring all amino acids. More than 97% of these organelle-like nuclear tRNA genes flank organelle-like sequences over 20 bp. Nearly 40% of them colocalize with two or more other organelle-like tRNA genes. Twelve of the 15 plastid-like mitochondrial tRNA genes possess 5′- and 3′-flanking sequences over 20 bp, and they are highly similar to their plastid counterparts. Phylogenetic analyses of the migrated tRNA genes and their original copies suggest that intergenomic tRNA gene transfer is an ongoing process with noticeable discriminatory routes among genomes in flowering plants. Electronic Supplementary Material Electronic Supplementary material is available for this article at and accessible for authorised users. Reviewing Editor: Dr. David Guttman  相似文献   

5.
The plastid genome of Trifolium subterraneum is 144,763 bp, about 20 kb longer than those of closely related legumes, which also lost one copy of the large inverted repeat (IR). The genome has undergone extensive genomic reconfiguration, including the loss of six genes (accD, infA, rpl22, rps16, rps18, and ycf1) and two introns (clpP and rps12) and numerous gene order changes, attributable to 14–18 inversions. All endpoints of rearranged gene clusters are flanked by repeated sequences, tRNAs, or pseudogenes. One unusual feature of the Trifolium subterraneum genome is the large number of dispersed repeats, which comprise 19.5% (ca. 28 kb) of the genome (versus about 4% for other angiosperms) and account for part of the increase in genome size. Nine genes (psbT, rbcL, clpP, rps3, rpl23, atpB, psbN, trnI-cau, and ycf3) have also been duplicated either partially or completely. rpl23 is the most highly duplicated gene, with portions of this gene duplicated six times. Comparisons of the Trifolium plastid genome with the Plant Repeat Database and searches for flanking inverted repeats suggest that the high incidence of dispersed repeats and rearrangements is not likely the result of transposition. Trifolium has 19.5 kb of unique DNA distributed among 160 fragments ranging in size from 30 to 494 bp, greatly surpassing the other five sequenced legume plastid genomes in novel DNA content. At least some of this unique DNA may represent horizontal transfer from bacterial genomes. These unusual features provide direction for the development of more complex models of plastid genome evolution. Electronic supplementary material  The online version of this article (doi:) contains supplementary material, which is available to authorized users.  相似文献   

6.
Cyanidioschyzon merolae andCyanidium caldarium are representative species among of the most primitive algae, although the two species are distinctly different in various morphological traits. We determined the nucleotide sequence of therbcL gene and a flanking 8-kb region in the plastid genome of each of these algae. In both algae, 12 genes were identified in this region, in an identical order. This gene order is not conserved in the plastid genomes of other species of the kingdom Plantae that have been sequenced to data. An additional unidentified open reading frame was also found in the two algae that we analyzed, which has not been described in any other species of algae includingPorphyra purpurea. Comparison of the amino acid sequences of selected genes also supported the conclusion thatCyanidioschyzon merolae andCyanidium caldarium are closely related and that they are distinct from other rhodophytes. The nucleotide sequence data reported in this paper will appear in the DDBJ, EMBL and GenBank Nucleotide Sequence Databases under the accession numbers D63675 and D63676.  相似文献   

7.
The complete nucleotide sequence of mulberry (Morus indica cv. K2) chloroplast genome (158,484 bp) has been determined using a combination of long PCR and shotgun-based approaches. This is the third angiosperm tree species whose plastome sequence has been completely deciphered. The circular double-stranded molecule comprises of two identical inverted repeats (25,678 bp each) separating a large and a small single-copy region of 87,386 bp and 19,742 bp, respectively. A total of 83 protein-coding genes including five genes duplicated in the inverted repeat regions, eight ribosomal RNA genes and 37 tRNA genes (30 gene species) representing 20 amino acids, were assigned on the basis of homology to predicted genes from other chloroplast genomes. The mulberry plastome lacks the genes infA, sprA, and rpl21 and contains two pseudogenes ycf15 and ycf68. Comparative analysis, based on sequence similarity, both at the gene and genome level, indicates Morus to be closer to Cucumis and Lotus, phylogenetically. However, at genome level, inclusion of non-coding regions brings it closer to Eucalyptus, followed by Cucumis. This may reflect differential selection pressure operating on the genic and intergenic regions of the chloroplast genome.Electronic supplementary material Supplementary material is available in the online version of this article at and is accessible for authorized users.Communicated by Y. Tsumura  相似文献   

8.
Peridinin-pigmented dinoflagellates contain secondary plastids that seem to have undergone more nearly complete plastid genome reduction than other eukaryotes. Many typically plastid-encoded genes appear to have been transferred to the nucleus, with a few remaining genes found on minicircles. To understand better the evolution of the dinoflagellate plastid, four categories of plastid-associated genes in dinoflagellates were defined based on their history of transfer and evaluated for rate of sequence evolution, including minicircle genes (presumably plastid-encoded), genes probably transferred from the plastid to the nucleus (plastid-transferred), and genes that were likely acquired directly from the nucleus of the previous plastid host (nuclear-transferred). The fourth category, lateral-transferred genes, are plastid-associated genes that do not appear to have a cyanobacterial origin. The evolutionary rates of these gene categories were compared using relative rate tests and likelihood ratio tests. For comparison with other secondary plastid-containing organisms, rates were calculated for the homologous sequences from the haptophyte Emiliania huxleyi. The evolutionary rate of minicircle and plastid-transferred genes in the dinoflagellate was strikingly higher than that of nuclear-transferred and lateral-transferred genes and, also, substantially higher than that of all plastid-associated genes in the haptophyte. Plastid-transferred genes in the dinoflagellate had an accelerated rate of evolution that was variable but, in most cases, not as extreme as the minicircle genes. Furthermore, the nuclear-transferred and lateral-transferred genes showed rates of evolution that are similar to those of other taxa. Thus, nucleus-to-nucleus transferred genes have a more typical rate of sequence evolution, while those whose history was wholly or partially within the dinoflagellate plastid genome have a markedly accelerated rate of evolution. Electronic Supplementary Material Electronic Supplementary material is available for this article at and accessible for authorised users. [Reviewing Editor: Dr. Debashish Battacharya]  相似文献   

9.
Summary Partial sequence analysis of the plastid DNA (ptDNA) from a red alga, Antithamnion sp., revealed the presence of a homologue to the Escherichia coli SecA gene as well as two open reading frames (ORF 510, ORF 179). In addition a sec Y homologue has been detected on the plastid genome by heterologous hybridization. None of these genes has been found in completely sequenced chlorophytic plastid genomes. SecA and secY gene copies were also detected in the ptDNA of a chromophytic alga, indicating that secAY may be ubiquitous in rhodophytes and chromophytes. The significance of these findings for the evolution of plastid genomes and the thylakoid protein import mechanism is discussed.  相似文献   

10.
Plastid transformation vectors are E. coli plasmids carrying a plastid marker gene for selection, adjacent cloning sites and flanking plastid DNA to target insertions in the plastid genome by homologous recombination. We report here on a family of next generation plastid vectors carrying synthetic DNA vector arms targeting insertions in the rbcL-accD intergenic region of the tobacco (Nicotiana tabacum) plastid genome. The pSS22 plasmid carries only synthetic vector arms from which the undesirable restriction sites have been removed by point mutations. The pSS24 vector carries a c-Myc tagged spectinomycin resistance (aadA) marker gene whereas in vector pSS30 aadA is flanked with loxP sequences for post-transformation marker excision. The synthetic vectors will enable direct manipulation of passenger genes in the transformation vector targeting insertions in the rbcL-accD intergenic region that contains many commonly used restriction sites. Electronic supplementary material  The online version of this article (doi:) contains supplementary material, which is available to authorized users.  相似文献   

11.
Plastid marker gene excision by the phiC31 phage site-specific recombinase   总被引:5,自引:0,他引:5  
Marker genes are essential for selective amplification of rare transformed plastid genome copies to obtain genetically stable transplastomic plants. However, the marker gene becomes dispensable when homoplastomic plants are obtained. Here we report excision of plastid marker genes by the phiC31 phage site-specific integrase (Int) that mediates recombination between bacterial (attB) and phage (attP) attachment sites. We tested marker gene excision in a two-step process. First we transformed the tobacco plastid genome with the pCK2 vector in which the spectinomycin resistance (aadA) marker gene is flanked with suitably oriented attB and attP sites. The transformed plastid genomes were stable in the absence of Int. We then transformed the nucleus with a gene encoding a plastid-targeted Int that led to efficient marker gene excision. The aadA marker free Nt-pCK2-Int plants were resistant to phosphinothricin herbicides since the pCK2 plastid vector also carried a bar herbicide resistance gene that, due to the choice of its promoter, causes a yellowish-golden (aurea) phenotype. Int-mediated marker excision reported here is an alternative to the currently used CRE/loxP plastid marker excision system and expands the repertoire of the tools available for the manipulation of the plastid genome.  相似文献   

12.
Apyrases have been suggested to play important roles in plant nutrition, photomorphogenesis, and nodulation. To help trace the evolution of these genes in the legumes—and possibly, the acquisition of new functions for nodulation—apyrase-containing BACs were sequenced from three legume genomes. Genomic sequences from Medicago truncatula, Glycine max and Lotus japonicus were compared to one another and to corresponding regions in Arabidopsis thaliana. A phylogenetic analysis of apyrase homologs from these regions and sequences from other legume species, as well as other plant families, identified a potentially legume-specific clade that contains a well-characterized soybean ( G. soja) apyrase, Gs52, as well as homologs from Dolichos, Lotus , Medicago and Pisum. Sister clades contain homologs from members of Brassicaceae, Solanaceae, Poaceae and Fabaceae. Comparisons of rates of change at synonymous and nonsynonymous sites in the Gs52 and sister clades show rapid evolution in the potentially legume-specific Gs52 clade. The genomic organization of the apyrase-containing BACs shows evidence of gene duplication, genomic rearrangement, and gene conversion among Gs52 homologs. Taken together, these results suggest a scenario of local apyrase gene duplication in an ancestor of the legumes, followed by functional diversification and increased rates of change in the new genes, and further duplications in the Galegae (which include the genera Medicago and Pisum). The study also provides a detailed comparison of genomic regions between two model genomes which are now being sequenced ( M. truncatula and L. japonicus), and a genome from an economically important legume species ( G. max).Electronic Supplementary Material Supplementary material is available in the online version of this article at Communicated by A. Kondorosi  相似文献   

13.
The E8 gene is related to ethylene biosynthesis in plants. To explore the effect of the expression pattern of the E8 gene on different E8 promoters, the molecular evolution of E8 promoters was investigated. A total of 16 E8 promoters were cloned from 16 accessions of seven tomato species, and were further analysed. The results from 19 E8 promoters including three previously cloned E8 promoters (X13437, DQ317599 and AF515784) showed that the size of the E8 promoters varied from 2101 bp (LA2150) to 2256 bp (LA2192); their sequences shared 69.9% homology and the average A/T content was 74.9%. Slide-window analysis divided E8 promoters into three regions — A, B and C — and the sequence identity in these regions was 72.5%, 41.2% and 70.8%, respectively. By searching the cis-elements of E8 promoters in the PLACE database, mutant nucleotides were found in some functional elements, and deletions or insertions were also found in regions responsible for ethylene biosysnthesis (−1702 to −1274) and the negative effect region (−1253 to −936). Our results indicate that the size of the functional region for ethylene biosynthesis in the E8 promoter could be shortened from 429 bp to 113 bp (−1612 to −1500). The results of molecular evolution analysis showed that the 19 E8 promoters could be classified into four clade groups, which is basically consistent with evolution of the tomato genome. Southern blot analysis results showed that the copy number of E8 promoters in tomato and some other wild species changed from 1 to 4. Taken together, our study provides important information for further elucidating the E8 gene expression pattern in tomato, analysing functional elements in the E8 promoter and reconstructing the potent E8 promoter. Electronic Supplementary Material  Supplementary material is available for this article at and is accessible for authorized users. Supplementary material pertaining to this article is available on the Journal of Biosciences Website at  相似文献   

14.
15.
Tomato plastid transformants were obtained using two vectors containing cloned plastid DNA of either Nicotiana tabacum or Solanum nigrum and including point mutations conferring resistance to spectinomycin and streptomycin. Transformants were recovered after PEG-mediated direct DNA uptake into protoplasts, followed by selection on spectinomycin-containing medium. Sixteen lines contained the point mutation, as confirmed by mapping restriction enzyme sites. One line obtained with each vector was analysed in more detail, in comparison with a spontaneous spectinomycin-resistant mutant. Integration of the cloned Solanum or Nicotiana plastid DNA, by multiple recombination events, into the tomato plastome was confirmed by sequence analysis of the targeted region of plastid DNA in the inverted repeat region. Maternal inheritance of spectinomycin and streptomycin resistances or sensitivity in seedlings also confirmed the transplastomic status of the two transformants. The results demonstrate the efficacy in tomato of a selection strategy which avoids the integration of a dominant bacterial antibiotic resistance gene.  相似文献   

16.
Summary The comparison of EcoRI patterns of chloroplast DNAs (ctDNAs) from five species of the genus Pelargonium and from 16 cultivars and varieties of Pelargonium zonale hort. demonstrates a remarkable inter- and intraspecific ctDNA (plastome) variation. The plastome of the P. zonale varieties could be differentiated into groups I, II and III. Reasons for this variation seem to be: occurrence of numerous spontaneous plastome mutations, intense hybridisation by gardeners and breeders, and biparental plastid inheritance.Crosses of P. zonale varieties with different ctDNA types lead to the direct evidence on the molecular level of biparental plastid inheritance and plastid sorting-out in F1-hybrids.  相似文献   

17.
Plastids from Nicotiana benthamiana were transformed with the vector for dicistronic expression of two genes—aminoglycoside 3'-adenyltransferase (aadA) and green fluorescent protein (gfp)—in the plastids of Nicotiana tabacum. Transplastomic shoots exhibited green fluorescence under UV light. Transformation efficiencies were similar between species. Although the border sequence (trnI and trnA) for homologous recombination to transform the plastid genome of N. benthamiana was identical to that sequence of N. tabacum, the exception was a 9-bp addition in the intron of trnI. This indicated that the N. tabacum sequence used as a border region for recombination was sufficient to insert the foreign gene into the target site between the trnI and trnA of N. benthamiana with similar efficiency. Southern blot analysis detected the presence of aadA and gfp between trnI and trnA in the plastid genome of N. benthamiana. Northern and western blot analyses revealed high expression of gfp in the plastids from petals and leaves. Our results suggest that the plastid transformation system established here is applicable to investigations of the interactions between plastid and nucleus in N. benthamiana.  相似文献   

18.
Cuscuta (dodder) is the only parasitic genus found in Convolvulaceae (morning-glory family). We used long PCR approach to obtain large portions of plastid genome sequence from Cuscuta sandwichiana in order to determine the size, structure, gene content, and synteny in the plastid genome of this Cuscuta species belonging to the poorly investigated holoparasitic subgenus Grammica. These new sequences are compared with the tobacco chloroplast genome, and, where data are available, with corresponding regions from taxa in the other Cuscuta subgenera. When all known plastid genome structural rearrangements in parasitic and nonparasitic Convolvulaceae are considered in a molecular phylogenetic framework, three categories of rearrangements in Cuscuta are revealed: plesiomorphic, autapomorphic, and synapomorphic. Many of the changes in Cuscuta, previously attributed to its parasitic mode of life, are better explained either as plesiomorphic conditions within the family, i.e., conditions shared with the rest of the Convolvulaceae, or, in most cases, autapomorphies of particular Cuscuta taxa, not shared with the rest of the species in the genus. The synapomorphic rearrangements are most likely to correlate with the parasitic lifestyle, because they represent changes found in Cuscuta exclusively. However, it appears that most of the affected regions, belonging to all of these three categories, have probably no function (e.g., introns) or are of unknown function (a number of open reading frames, the function of which, if any, has yet to be discovered).[Reviewing Editor: Dr. Debashish Bhattacharya]  相似文献   

19.
A series of strains of the homozygous speciesOenothera grandiflora (characterized by the genome BB and plastome III) were combined with plastome IV fromO. parviflora (BC-IV) by means of appropriate crosses. An incompatibility between genome B and plastome IV is expressed in the haplo- and diplophase: (1) B-IV pollen, though normally developed, is largely inactive. The extent of the inactivation varies between different strains and shows a seasonal fluctuation as determined by seed set in outcrossing and selfing experiments. (2) In most of the strains lethality of BB-IV embryos is the rule, leading to empty seeds. This can be ameliorated by including another plastome in the zygotes and developing embryos on account of the biparental plastid transmission inOenothera. It can best be demonstrated in crosses with a seed parent having normal green plastids of plastome IV and mutated chlorophyll deficient plastids from a different plastome in the pollen parent, leading to variegated progeny as well as a remainder of empty seeds. (3) In about one-half of the strains the BB-IV plants exhibit a temporary bleaching of thevirescens type. The incompatibily between genome B and plastome IV does not support the earlier assumption that plastome IV is the ancestor of plastomes II, III, and V. Instead, a precursor plastome is postulated from which plastomes II, III, and IV are descended. While plastome I can be derived from II, only plastome V can be descended from plastome IV.Deceased August 28, 1998.  相似文献   

20.
Plastid DNA (ptDNA) probes were used in RFLP analysis to determine ptDNA inheritance in interspecific hybrids in Zantedeschia. Biparental and maternal ptDNA inheritance was found in albino hybrids between the evergreen species Z. aethiopica and several winter-dormant species. From two albino hybrids, different types of ptDNA were detected in shoots derived from different parts of an embryo. This result indicates that plastids were sorted out during embryo development. Only maternal ptDNA was detected in the hybrids of Z. aethiopica × Z. odorata (a summer-dormant species) but paternal, biparental, and maternal ptDNA were found in the hybrids of the reciprocal cross. Z. odorata × Z. aethiopica. By correlating these ptDNA inheritance patterns with the leaf colour (albino, pale-green, and green) of the hybrids, it is suggested that the Z. odorata plastome is incompatible with the Z. aethiopica genome. The Z. aethiopica plastome is partially compatible with the Z. odorata genome but the development of Z. aethiopica plastids appears to be blocked by the presence of the Z. odorata plastids.  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