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 626 毫秒
1.
We isolated RNAs by selection–amplification, selecting for affinity to Phe–Sepharose and elution with free l-phenylalanine. Constant sequences did not contain Phe condons or anticodons, to avoid any possible confounding influence on initially randomized sequences. We examined the eight most frequent Phe-binding RNAs for inclusion of coding triplets. Binding sites were defined by nucleotide conservation, protection, and interference data. Together these RNAs comprise 70% of the 105 sequenced RNAs. The K D for the strongest sites is ≈50 μM free amino acid, with strong stereoselectivity. One site strongly distinguishes free Phe from Trp and Tyr, a specificity not observed previously. In these eight Phe-binding RNAs, Phe codons are not significantly associated with Phe binding sites. However, among 21 characterized RNAs binding Phe, Tyr, Arg, and Ile, containing 1342 total nucleotides, codons are 2.7-fold more frequent within binding sites than in surrounding sequences in the same molecules. If triplets were not specifically related to binding sites, the probability of this distribution would be 4.8 × 10−11. Therefore, triplet concentration within amino acid binding sites taken together is highly likely. In binding sites for Arg, Tyr, and Ile cognate codons are overrepresented. Thus Arg, Tyr, and Ile may be amino acids whose codons were assigned during an era of direct RNA–amino acid affinity. In contrast, Phe codons arguably were assigned by another criterion, perhaps during later code evolution.  相似文献   

2.
Highly expressed plastid genes display codon adaptation, which is defined as a bias toward a set of codons which are complementary to abundant tRNAs. This type of adaptation is similar to what is observed in highly expressed Escherichia coli genes and is probably the result of selection to increase translation efficiency. In the current work, the codon adaptation of plastid genes is studied with regard to three specific features that have been observed in E. coli and which may influence translation efficiency. These features are (1) a relatively low codon adaptation at the 5′ end of highly expressed genes, (2) an influence of neighboring codons on codon usage at a particular site (codon context), and (3) a correlation between the level of codon adaptation of a gene and its amino acid content. All three features are found in plastid genes. First, highly expressed plastid genes have a noticeable decrease in codon adaptation over the first 10–20 codons. Second, for the twofold degenerate NNY codon groups, highly expressed genes have an overall bias toward the NNC codon, but this is not observed when the 3′ neighboring base is a G. At these sites highly expressed genes are biased toward NNT instead of NNC. Third, plastid genes that have higher codon adaptations also tend to have an increased usage of amino acids with a high G + C content at the first two codon positions and GNN codons in particular. The correlation between codon adaptation and amino acid content exists separately for both cytosolic and membrane proteins and is not related to any obvious functional property. It is suggested that at certain sites selection discriminates between nonsynonymous codons based on translational, not functional, differences, with the result that the amino acid sequence of highly expressed proteins is partially influenced by selection for increased translation efficiency. Received: 21 July 1999 / Accepted: 5 November 1999  相似文献   

3.
Summary Bumetanide-sensitive Na-K-Cl cotransporters and thiazide-sensitive Na-Cl cotransporters comprise a family of integral membrane transport proteins, the Na-K-Cl cotransporter (NKCC) family. Each of the members of this family is over 1,000 amino acids in length. We have multiply aligned the ten currently sequenced members of this family from human, rabbit, rodent, shark, flounder, moth, worm and yeast sources. Phylogenetic analyses suggest the presence of at least six isoforms of these full length proteins in eukaryotes. Average hydropathy and average similarity plots have been derived revealing that each of these proteins possesses a central, well conserved, hydrophobic domain of almost invariant length, possibly consisting of twelve transmembrane α-helical spanners, an N-terminal, poorly conserved, hydrophilic domain of variable length, and a C-terminal, moderately conserved, hydrophilic domain of moderately constant length. A functionally uncharacterized homologue of this family occurs in the cyanobacterium Synechococcus sp. Limited sequence similarity of these proteins with members of a family of basic amino acid transporters suggests that the NKCC family may be distantly related to the previously characterized, ubiquitous, amino acid-polyamine-choline (APC) family of facilitators. These observations suggest that the NKCC family is an old family that has its roots in the prokaryotic kingdom. Received: 27 July 1995/Revised: 8 November 1995  相似文献   

4.
Multidrug resistance to anti-cancer drugs is a major medical problem. Resistance is manifested largely by the product of the human MDR1 gene, P-glycoprotein, an ABC transporter that is an integral membrane protein of 1280 amino acids arranged into two homologous halves, each comprising 6 putative transmembrane α-helices and an ATP binding domain. Despite the plethora of data from site-directed, scanning and domain replacement mutagenesis, epitope mapping and photoaffinity labeling, a clear structural model for P-glycoprotein remains largely elusive. In this report, we propose a new model for P-glycoprotein that is supported by the vast body of previous data. The model comprises 2 membrane-embedded 16-strand β-barrels, attached by short loops to two 6-helix bundles beneath each barrel. Each ATP binding domain contributes 2 β-strands and 1 α-helix to the structure. This model, together with an analysis of the amino acid sequence alignment of P-glycoprotein isoforms, is used to delineate drug binding and translocation sites. We show that the locations of these sites are consistent with mutational, kinetic and labeling data. Received: 18 February 1998/Revised: 2 September 1998  相似文献   

5.
In this study, we analyzed the correlation between codon usage bias and Shine–Dalgarno (SD) sequence conservation, using complete genome sequences of nine prokaryotes. For codon usage bias, we adopted the codon adaptation index (CAI), which is based on the codon usage preference of genes encoding ribosomal proteins, elongation factors, heat shock proteins, outer membrane proteins, and RNA polymerase subunit proteins. To compute SD sequence conservation, we used SD motif sequences predicted by Tompa and systematically aligned them with 5′UTR sequences. We found that there exists a clear correlation between the CAI values and SD sequence conservation in the genomes of Escherichia coli, Bacillus subtilis, Haemophilus influenzae, Archaeoglobus fulgidus, Methanobacterium thermoautotrophicum, and Methanococcus jannaschii, and no relationship is found in M. genitalium, M. pneumoniae, and Synechocystis. That is, genes with higher CAI values tend to have more conserved SD sequences than do genes with lower CAI values in these organisms. Some organisms, such as M. thermoautotrophicum, do not clearly show the correlation. The biological significance of these results is discussed in the context of the translation initiation process and translation efficiency. Received: 22 June 2000 / Accepted: 18 October 2000  相似文献   

6.
Amino acid substitution tables are essential for the proper alignment of protein sequences, and alignment scores based on them can be transformed into distance measures by various means. In the simplest case, the negative log of the score is used. This Poisson relationship assumes that all sites are equally likely to change, however. A more accurate relationship would correct for different rates of change at each residue position. Recently, Grishin (J. Mol. Evol. 41:675–679, 1995) published a set of simple equations that correct for various circumstances, including different rates of change at different sites. We have used these equations in conjunction with similarity scores that take into account constraints on amino acid interchange. Simulation studies show a linear relationship between these calculated distances and the numbers of allowed mutations based on the observed variation of rate at all sites in various proteins. Received: 25 January 1996 / Accepted: 1 October 1996  相似文献   

7.
Endosymbiotic bacteria live in animal cells and are transmitted vertically at the time of the host's reproduction. In view of their small and asexual populations with infrequent chances of recombination, these endocellular bacteria are expected to accumulate mildly deleterious mutations. Previous studies showed that the DNA sequences of these bacteria evolved faster than those of free-living bacteria. In this study, we compared all the ORFs of Buchnera, an endocellular bacterial symbiont of aphids, with those of 34 other prokaryotic organisms and estimated the effect of the accelerated evolution of Buchnera on the functions of its proteins. It was revealed that Buchnera proteins contain many mutations at the sites where sequences are conserved in their orthologues in many other organisms. In addition, amino acid replacements at the conserved sites are mostly changes to physicochemically different amino acids. These results suggest that functions and conformations of Buchnera proteins have been seriously impaired or strongly modified. Indeed, extensive loss of functional motifs was observed in some Buchnera proteins. In many Buchnera proteins mutations were not detected evenly throughout each molecule but tended to accumulate in some functional units, possibly leading to loss of specific functions. As Buchnera has an unusual and limited gene repertory, it is conceivable that the manner of interactions among its proteins has been changed, and thus, functional constraints over their amino acid residues have also been changed during evolution. This may account for the loss of some functional units only in the Buchnera proteins. We obtained evidence that amino acid replacements in Buchnera were not always deleterious, but neutral or, in some cases, even positively selected. Received: 14 December 2000 / Accepted: 12 March 2001  相似文献   

8.
Biased codon usage is common in eukaryotic and prokaryotic genes. Evidence from Escherichia, Saccharomyces, and Drosophila indicates that it favors translational efficiency and accuracy. However, to date no functional advantages have been identified in the codon–anticodon interactions involving the most frequently used (preferred) codons. Here we present evidence that forces not related to the individual codon–anticodon interaction may be involved in determining which synonymous codons are preferred or avoided. We show that the ``off-frame' trinucleotide motif preferences inferrable from Drosophila coding regions are often in the same direction as Drosophila's ``in-frame' codon preferences, i.e., its codon usage. The off-frame preferences were inferred from the nonrandomness of the location of confamilial synonymous codons along coding regions—a pattern often described as a context dependence of nucleotide choice at synonymous positions or as codon-pair bias. We relied on randomizations of the location of confamilial codons that do not alter, and cannot be influenced by, the encoded amino acid sequences, codon usage, or base composition of the genes examined. The statistically significant congruency of in-frame and off-frame trinucleotide preferences suggests that the same kind of reading-frame-independent force(s) may also influence synonymous codon choice. These forces may have produced biases in codon usage that then led to the evolution of the translational advantages of these motifs as preferred codons. Under this scenario, tRNA pool size differences between preferred and nonpreferred codons initially were evolved to track the default overrepresentation of codons with preferred motifs. The motif preference hypothesis can explain the structuring of codon preferences and the similarities in the codon usages of distantly related organisms. Received: 10 November 1998 / Accepted: 23 February 1999  相似文献   

9.
Aquatic larvae of the midge, Chironomus tentans, synthesize a 185-kDa silk protein (sp185) with the cysteine-containing motif Cys-X-Cys-X-Cys (where X is any residue) every 20–28 residues. We report here the cloning and full-length sequence of cDNAs encoding homologous silk proteins from Chironomus pallidivittatus (sp185) and Chironomus thummi (sp220). Deduced amino acid sequences reveal proteins of nearly identical mass composed of 72 blocks of 20–28 residues, 61% of which can be described by the motif X5–8-Cys-X5-(Trp/Phe/Tyr)-X4-Cys-X-Cys-X-Cys. Spatial arrangement of these residues is preserved more than surrounding sequences. cDNA clones enabled us to map the genes on polytene chromosomes and identify for the first time the homolog of the Camptochironomus Balbiani ring 3 locus in Chironomus thummi. The apparent molecular weight difference between these proteins (185 vs 220 kDa) is not attributable to primary structure and may be due to differential N-linked glycosylation. DNA distances and codon substitutions indicate that the C. tentans and C. pallidivittatus genes are more related to each other than either is to C. thummi; however, substitution rates for the 5′- and 3′-halves of these genes are different. Blockwise sequence comparisons suggest intragenic variation in that some regions evolved slower or faster than the mean and may have been subjected to different selective pressures. Received: 30 August 1996 / Accepted: 6 November 1996  相似文献   

10.
11.
We previously found that proteinaceous protease inhibitors homologous to Streptomyces subtilisin inhibitor (SSI) are widely produced by various Streptomyces species, and we designated them ``SSI-like proteins' (Taguchi S, Kikuchi H, Suzuki M, Kojima S, Terabe M, Miura K, Nakase T, Momose H [1993] Appl Environ Microbiol 59:4338–4341). In this study, SSI-like proteins from five strains of the genus Streptoverticillium were purified and sequenced, and molecular phylogenetic trees were constructed on the basis of the determined amino acid sequences together with those determined previously for Streptomyces species. The phylogenetic trees showed that SSI-like proteins from Streptoverticillium species are phylogenetically included in Streptomyces SSI-like proteins but form a monophyletic group as a distinct lineage within the Streptomyces proteins. This provides an alternative phylogenetic framework to the previous one based on partial small ribosomal RNA sequences, and it may indicate that the phylogenetic affiliation of the genus Streptoverticillium should be revised. The phylogenetic trees also suggested that SSI-like proteins possessing arginine or methionine at the P1 site, the major reactive center site toward target proteases, arose multiple times on independent lineages from ancestral proteins possessing lysine at the P1 site. Most of the codon changes at the P1 site inferred to have occurred during the evolution of SSI-like proteins are consistent with those inferred from the extremely high G + C content of Streptomyces genomes. The inferred minimum number of amino acid replacements at the P1 site was nearly equal to the average number for all the variable sites. It thus appears that positive Darwinian selection, which has been postulated to account for accelerated rates of amino acid replacement at the major reaction center site of mammalian protease inhibitors, may not have dictated the evolution of the bacterial SSI-like proteins. Received: 23 August 1996 / Accepted: 20 November 1996  相似文献   

12.
In this study, the region corresponding to the Thr–Gly region of the period (per) gene in the Drosophila nasuta subgroup of species was sequenced. The results showed that this region was highly conserved in the D. nasuta subgroup. There were only nine variable sites found in this 300-bp-long region, all located in two small regions highly variable among Drosophila species. No length variation was observed either within this subgroup or in the Yunnan (YN) population of D. albomicans. The deduced amino acid sequences were identical for all 14 taxa in the D. nasuta subgroup, and a stretch of alternating Thr–Gly pairs was not observed in this subgroup. A phylogenetic tree was constructed. The clustering of some species was in general agreement with previous works, but it also raised some question on the phylogenetic relationship between the nasuta species. The data did not implicate the Thr–Gly region playing a role in behavioral isolation in this subgroup of Drosophila. Received: 8 February 1999 / Accepted: 22 April 1999  相似文献   

13.
A molecular phylogenetic analysis of elongation factor Tu (EF-Tu) proteins from plastids was performed in an attempt to identify the origin of chlorarachniophyte plastids, which are considered to have evolved from the endosymbiont of a photosynthetic eukaryote. Partial sequences of the genes for plastid EF-Tu proteins (1,080–1,089 bp) were determined for three algae that contain chlorophyll b, namely, Gymnochlora stellata (Chlorarachniophyceae), Bryopsis maxima (Ulvophyceae), and Pyramimonas disomata (Prasinophyceae). The deduced amino acid sequences were used to construct phylogenetic trees of the plastid and bacterial EF-Tu proteins by the maximum likelihood, the maximum parsimony, and the neighbor joining methods. The trees obtained in the present analysis suggest that all plastids that contain chlorophyll b are monophyletic and that the chlorarachniophyte plastids are closely related to those of the Ulvophyceae. The phylogenetic trees also suggest that euglenophyte plastids are closely related to prasinophycean plastids. The results indicate that the chlorarachniophyte plastids evolved from a green algal endosymbiont that was closely related to the Ulvophyceae and that at least two secondary endosymbiotic events have occurred in the lineage of algae with plastids that contain chlorophyll b. Received: 10 March 1997 / Accepted: 28 July 1997  相似文献   

14.
The available amino acid sequences of the α-amylase family (glycosyl hydrolase family 13) were searched to identify their domain B, a distinct domain that protrudes from the regular catalytic (β/α)8-barrel between the strand β3 and the helix α3. The isolated domain B sequences were inspected visually and also analyzed by Hydrophobic Cluster Analysis (HCA) to find common features. Sequence analyses and inspection of the few available three-dimensional structures suggest that the secondary structure of domain B varies with the enzyme specificity. Domain B in these different forms, however, may still have evolved from a common ancestor. The largest number of different specificities was found in the group with structural similarity to domain B from Bacillus cereus oligo-1,6-glucosidase that contains an α-helix succeeded by a three-stranded antiparallel β-sheet. These enzymes are α-glucosidase, cyclomaltodextrinase, dextran glucosidase, trehalose-6-phosphate hydrolase, neopullulanase, and a few α-amylases. Domain B of this type was observed also in some mammalian proteins involved in the transport of amino acids. These proteins show remarkable similarity with (β/α)8-barrel elements throughout the entire sequence of enzymes from the oligo-1,6-glucosidase group. The transport proteins, in turn, resemble the animal 4F2 heavy-chain cell surface antigens, for which the sequences either lack domain B or contain only parts thereof. The similarities are compiled to indicate a possible route of domain evolution in the α-amylase family. Received: 4 December 1996 / Accepted: 13 March 1997  相似文献   

15.
The Artemia hemoglobin is a dimer comprising two nine-domain covalent polymers in quaternary association. Each polymer is encoded by a gene representing nine successive globin domains which have different sequences and are presumed to have been copied originally from a single-domain gene. Two different polymers exist as the result of a complete duplication of the nine-domain gene, allowing the formation of either homodimers or the heterodimer. The total population size of 18 domains comprising nine corresponding pairs, coupled with the probability that they reflect several hundred million years of evolution in the same lineage, provides a unique model in which the process of gene multiplication can be analyzed. The outcome has important implications for the reliability of local molecular clocks. The two polymers differ from each other at 11.7% of amino acid sites; however when corresponding individual domains are compared between polymers, amino acid substitution fluctuates by a factor of 2.7-fold from lowest to highest. This variation is not obvious at the DNA level: Domain pair identity values fluctuate by 1.3-fold. Identity values are, however, uncorrected for multiple substitutions, and both silent and nonsilent changes are pooled. Therefore, to determine the variability in relative substitution rates at the DNA level, we have used the method of Li (1993, J Mol Evol 36:96–99) to determine estimates of nonsynonymous (K A ) and synonymous (K S ) substitutions per site for the nine pairs of domains. As expected, the overall level of silent substitutions (K S of 56.9%) far exceeded nonsilent substitutions (K A of 6.7%); however, for corresponding domain pairs, K A fluctuates by 2.3-fold and K S by 1.7-fold. The large discrepancies reflected in the expressed protein have accrued within a single lineage and the implication is that divergence dates of different genera based on amino acid sequences, even with well-studied proteins of reasonable size, can be wrong by a factor well in excess of 2. Received: 4 June 1997 / Accepted: 17 December 1997  相似文献   

16.
The extent to which base composition and codon usage vary among RNA viruses, and the possible causes of this bias, is undetermined in most cases. A maximum-likelihood statistical method was used to test whether base composition and codon usage bias covary with arthropod association in the genus Flavivirus, a major source of disease in humans and animals. Flaviviruses are transmitted by mosquitoes, by ticks, or directly between vertebrate hosts. Those viruses associated with ticks were found to have a significantly lower G+C content than non-vector-borne flaviviruses and this difference was present throughout the genome at all amino acids and codon positions. In contrast, mosquito-borne viruses had an intermediate G+C content which was not significantly different from those of the other two groups. In addition, biases in dinucleotide and codon usage that were independent of base composition were detected in all flaviviruses, but these did not covary with arthropod association. However, the overall effect of these biases was slight, suggesting only weak selection at synonymous sites. A preliminary analysis of base composition, codon usage, and vector specificity in other RNA virus families also revealed a possible association between base composition and vector specificity, although with biases different from those seen in the Flavivirus genus. Received: 29 August 2000 / Accepted: 19 December 2000  相似文献   

17.
The plasma complement system comprises several activation pathways that share a common terminal route involving the assembly of the terminal complement complex (TCC), formed by C5b–C9. The order of emergence of the homologous components of TCC (C6, C7, C8α, C8β, and C9) has been determined by phylogenetic analyses of their amino acid sequences. Using all the sequence data available for C6–C9 proteins, as well as for perforins, the results suggested that these TCC components originated from a single ancestral gene and that C6 and C7 were the earliest to emerge. Our evidence supports the notion that the ancestral gene had a complex modular composition. A series of gene duplications in combination with a tendency to lose modules resulted in successive complement proteins with decreasing modular complexity. C9 and perforin apparently are the result of different selective conditions to acquire pore-forming function. Thus C9 and perforin are examples of evolutionary parallelism. Received: 16 August 1998 / Accepted: 12 March 1999  相似文献   

18.
The microenvironment near the apical membrane of MDCK cells was studied by quantitation of the fluorescence of wheat germ agglutin attached to fluorescein (WGA). WGA was shown to bind to sialic acid residues attached to galactose at the α-2,3 position in the glycocalyx on the apical membrane. Young MDCK cells (5–8 days after splitting) showed a patchy distribution of WGA at stable sites that returned to the same locations after removal of sialic acid residues by neuraminidase treatment. Other lectins also showed stable binding to patches on the apical membrane of young cells. The ratio of WGA fluorescence emission at two excitation wavelengths was used to measure near-membrane pH. The near-membrane pH was markedly acidic to the pH 7.4 bathing solution in both young and older cells (13–21 days after splitting). Patches on the apical membrane of young cells exhibited a range of near-membrane pH values with a mean ±sem of 6.86 ± 0.04 (n= 121) while the near-membrane pH of older cells was 6.61 ± 0.04 (n= 120) with a uniform WGA distribution. We conclude that the distribution of lectin binding sites in young cells reflects the underlying nonrandom location of membrane proteins in the apical membrane and that nonuniformities in the pH of patches may indicate regional differences in membrane acid-base transport as well as in the location of charged sugars in the glycocalyx. Received: 15 December 1999/Revised: 16 March 2000  相似文献   

19.
The α-esterase cluster of D. melanogaster contains 11 esterase genes dispersed over 60 kb. Embedded in the cluster are two unrelated open reading frames that have sequence similarity with genes encoding ubiquitin-conjugating enzyme and tropomyosin. The esterase amino acid sequences show 37–66% identity with one another and all but one have all the motifs characteristic of functional members of the carboxyl/cholinesterase multigene family. The exception has several frameshift mutations and appears to be a pseudogene. Patterns of amino acid differences among cluster members in relation to generic models of carboxyl/cholinesterase protein structure are broadly similar to those among other carboxyl/cholinesterases sequenced to date. However the α-esterases differ from most other members of the family in: their lack of a signal peptide; the lack of conservation in cysteines involved in disulfide bridges; and in four indels, two of which occur in or adjacent to regions that align with proposed substrate-binding sites of other carboxyl/cholinesterases. Phylogenetic analyses clearly identify three simple gene duplication events within the cluster. The most recent event involved the pseudogene which is located in an intron of another esterase gene. However, relative rate tests suggest that the pseudogene remained functional after the duplication event and has become inactive relatively recently. The distribution of indels also suggests a deeper node in the gene phylogeny that separates six genes at the two ends of the cluster from a block of five in the middle. Received: 18 January 1996 / Accepted: 12 March 1996  相似文献   

20.
A Laminaria saccharina genomic library in the phage EMBL 4 was used to isolate and sequence a full-length gene encoding a fucoxanthin-chlorophyll a/c-binding protein. Contrary to diatom homologues, the coding sequence is interrupted by an intron of about 900 bp which is located in the middle of the transit peptide. The deduced amino acid sequence of the mature protein is very similar to those of related proteins from Macrocystis pyrifera (Laminariales) and, to a lesser extent, to those from diatoms and Chrysophyceae. Seven of the eight putative chlorophyll-binding amino acids determined in green plants are also present. Alignments of different sequences related to the light-harvesting proteins (LHC) demonstrate a structural similarity among the three transmembrane helices and suggest a unique ancestral helix preceded by two β-turns. The β-turns are conserved in front of the second helices of the chlorophyll a/c proteins more so than in chlorophyll a/b proteins. Phylogenetic trees generated from sequence data indicate that fucoxanthin-chlorophyll-binding proteins diverged prior to the separation of photosystem I and photosystem II LHC genes of green plants. Among the fucoxanthin-containing algae, LHC I or II families could not be distinguished at this time. Received: 14 February 1996 / Accepted: 4 April 1996  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