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 46 毫秒
1.
MOTIVATION: RNAs play an important role in many biological processes and knowing their structure is important in understanding their function. Due to difficulties in the experimental determination of RNA secondary structure, the methods of theoretical prediction for known sequences are often used. Although many different algorithms for such predictions have been developed, this problem has not yet been solved. It is thus necessary to develop new methods for predicting RNA secondary structure. The most-used at present is Zuker's algorithm which can be used to determine the minimum free energy secondary structure. However many RNA secondary structures verified by experiments are not consistent with the minimum free energy secondary structures. In order to solve this problem, a method used to search a group of secondary structures whose free energy is close to the global minimum free energy was developed by Zuker in 1989. When considering a group of secondary structures, if there is no experimental data, we cannot tell which one is better than the others. This case also occurs in combinatorial and heuristic methods. These two kinds of methods have several weaknesses. Here we show how the central limit theorem can be used to solve these problems. RESULTS: An algorithm for predicting RNA secondary structure based on helical regions distribution is presented, which can be used to find the most probable secondary structure for a given RNA sequence. It consists of three steps. First, list all possible helical regions. Second, according to central limit theorem, estimate the occurrence probability of every helical region based on the Monte Carlo simulation. Third, add the helical region with the biggest probability to the current structure and eliminate the helical regions incompatible with the current structure. The above processes can be repeated until no more helical regions can be added. Take the current structure as the final RNA secondary structure. In order to demonstrate the confidence of the program, a test on three RNA sequences: tRNAPhe, Pre-tRNATyr, and Tetrahymena ribosomal RNA intervening sequence, is performed. AVAILABILITY: The program is written in Turbo Pascal 7.0. The source code is available upon request. CONTACT: Wujj@nic.bmi.ac.cn or Liwj@mail.bmi.ac.cn  相似文献   

2.
Retroviruses encapsidate two molecules of genomic RNA that are noncovalently linked close to their 5' ends in a region called the dimer linkage structure (DLS). The dimerization initiation site (DIS) of human immunodeficiency virus type 1 (HIV-1) constitutes the essential part of the DLS in vitro and is crucial for efficient HIV-1 replication in cell culture. We previously identified the DIS as a hairpin structure, located upstream of the major splice donor site, that contains in the loop a six-nucleotide self-complementary sequence preceded and followed by two and one purines, respectively. Two RNA monomers form a kissing loop complex via intermolecular interactions of the six nucleotide self-complementary sequence. Here, we introduced compensatory mutations in the self-complementary sequence and/or a mutation in the flanking purines. We determined the kinetics of dimerization, the thermal stabilities and the apparent equilibrium dissociation constants of wild-type and mutant dimers and used chemical probing to obtain structural information. Our results demonstrate the importance of the 5'-flanking purine and of the two central bases of the self-complementary sequence in the dimerization process. The experimental data are rationalized by triple interactions between these residues in the deep groove of the kissing helix and are incorporated into a three-dimensional model of the kissing loop dimer. In addition, chemical probing and molecular modeling favor the existence of a non-canonical interaction between the conserved adenine residues at the first and last positions in the DIS loop. Furthermore, we show that destabilization of the kissing loop complex at the DIS can be compensated by interactions involving sequences located downstream of the splice donor site of the HIV-1 genomic RNA.  相似文献   

3.
The thiostrepton-resistance gene encoding the 23S rRNA A1067 methyltransferase from Streptomyces azureus has been overexpressed in Escherichia coli using a T7-RNA-polymerase-dependent expression vector. The protein was efficiently expressed at levels up to 20% of total soluble protein and purified to near homogeneity. Kinetic parameters for S-adenosyl-L-methionine (Km = 0.1 mM) and an RNA fragment containing nucleotides 1029-1122 of the 23S ribosomal RNA from E. coli (Km = 0.001 mM) were determined. S-Adenosyl-L-homocysteine showed competitive product inhibition (Ki = 0.013 mM). Binding of either thiostrepton or protein L11 inhibited methylation. RNA sequence variants of the RNA fragment with mutations in nucleotides 1051-1108 were tested as substrates for the methylase. The experimental data indicate that methylation is dependent on the secondary structure of the hairpin including nucleotide A1067 and the exact sequence U(1066)-A(1067)-G(1068)-A(1069)-A(1070) of the single strand.  相似文献   

4.
Telomerase is a ribonucleoprotein enzyme with an essential RNA component. Embedded within the telomerase RNA is a template sequence for telomere synthesis. We have characterized the structure of the 5' regions of the human and mouse telomerase-RNA genes, and have found a striking difference in the location of the template sequence: Whereas the 5'-end of the human telomerase RNA lies 45 nt from the telomerase-RNA template sequence, the 5'-end of the mouse telomerase RNA lies just 2 nt from the telomerase-RNA template sequence. Analysis of genomic sequences flanking the 5'-end of the human and mouse telomerase RNA-coding sequences reveals similar promoter-element arrangements typical of mRNA-type promoters: a TATA box-like element and an upstream region containing a consensus CCAAT box. This putative promoter structure contrasts with that of the ciliate telomerase-RNA genes whose structure resembles RNA polymerase III U6 small nuclear RNA (snRNA) promoters. These and other comparisons suggest that, during evolution, both the RNA-polymerase specificity of telomerase RNA-gene promoters and, more recently, the position of the template sequence in the telomerase RNA changed.  相似文献   

5.
6.
The sequence 5'-rUUGGCG-3' is conserved within the loop regions of antisense RNAs or their targets involved in replication of various prokaryotic plasmids. In IncIalpha plasmid ColIb-P9, the partially base paired 21-nucleotide loop of a stem-loop called structure I within RepZ mRNA contains this hexanucleotide sequence, and comprises the target site for the antisense Inc RNA. In this report, we find that the base pairing interaction at the 5'-rGGC-3' sequence in the hexanucleotide motif is important for interaction between Inc RNA and structure I. In addition, the 21-base loop domain of structure I is folded tighter than predicted, with the hexanucleotide sequence at the top. The second U residue in the sequence is favored for Inc RNA binding in a base-specific manner. On the other hand, the upper domain of the Inc RNA stem-loop is loosely structured, and maintaining the loop sequence single-stranded is important for the intermolecular interaction. Based on these results, we propose that a structural feature in the loop I domain, conferred probably by the conserved 5'-rUUGGCG-3' sequence, favors binding to a complementary, single-stranded RNA. This model also explains how the RepZ mRNA pseudoknot, described in the accompanying paper (Asano, K., and Mizobuchi, K. (1998) J. Biol. Chem. 273, 11815-11825) is formed specifically with structure I. A possible conformation adopted by the 5'-rUUGGCG-3' loop sequence is discussed.  相似文献   

7.
Ribonucleic acid (RNA) is a polymer composed of four bases denoted A, C, G, and U. It generally is a single-stranded molecule where the bases form hydrogen bonds within the same molecule leading to structure formation. In comparing different homologous RNA molecules it is important to consider both the base sequence and the structure of the molecules. Traditional alignment algorithms can only account for the sequence of bases, but not for the base pairings. Considering the structure leads to significant computational problems because of the dependencies introduced by the base pairings. In this paper we address the problem of optimally aligning a given RNA sequence of unknown structure to one of known sequence and structure. We phrase the problem as an integer linear program and then solve it using methods from polyhedral combinatorics. In our computational experiments we could solve large problem instances--23S ribosomal RNA with more than 1400 bases--a size intractable for former algorithms.  相似文献   

8.
9.
10.
Cis-acting RNA signals are required for replication of positive-strand viruses such as the picornaviruses. Although these generally have been mapped to the 5' and/or 3' termini of the viral genome, RNAs derived from human rhinovirus type 14 are unable to replicate unless they contain an internal cis-acting replication element (cre) located within the genome segment encoding the capsid proteins. Here, we show that the essential cre sequence is 83-96 nt in length and located between nt 2318-2413 of the genome. Using dicistronic RNAs in which translation of the P1 and P2-P3 segments of the polyprotein were functionally dissociated, we further demonstrate that translation of the cre sequence is not required for RNA replication. Thus, although it is located within a protein-coding segment of the genome, the cre functions as an RNA entity. Computer folds suggested that cre sequences could form a stable structure in either positive- or minus-strand RNA. However, an analysis of mutant RNAs containing multiple covariant and non-covariant nucleotide substitutions within these putative structures demonstrated that only the predicted positive-strand structure is essential for efficient RNA replication. The absence of detectable minus-strand synthesis from RNAs that lack the cre suggests that the cre is required for initiation of minus-strand RNA synthesis. Since a lethal 3' noncoding region mutation could be partially rescued by a compensating mutation within the cre, the cre appears to participate in a long-range RNA-RNA interaction required for this process. These data provide novel insight into the mechanisms of replication of a positive-strand RNA virus, as they define the involvement of an internally located RNA structure in the recognition of viral RNA by the viral replicase complex. Since internally located RNA replication signals have been shown to exist in several other positive-strand RNA virus families, these observations are potentially relevant to a wide array of related viruses.  相似文献   

11.
12.
BACKGROUND: One of the most significant questions in understanding the origin of life concerns the order of appearance of DNA, RNA and protein during early biological evolution. If an 'RNA world' was a precursor to extant life, RNA must be able not only to catalyze RNA replication but also to direct peptide synthesis. Iterative RNA selection previously identified catalytic RNAs (ribozymes) that form amide bonds between RNA and an amino acid or between two amino acids. RESULTS: We characterized peptidyl-transferase reactions catalyzed by two different families of ribozymes that use substrates that mimic A site and P site tRNAs. The family II ribozyme secondary structure was modeled using chemical modification, enzymatic digestion and mutational analysis. Two regions resemble the peptidyl-transferase region of 23S ribosomal RNA in sequence and structural context; these regions are important for peptide-bond formation. A shortened form of this ribozyme was engineered to catalyze intermolecular ('trans') peptide-bond formation, with the two amino-acid substrates binding through an attached AMP or oligonucleotide moiety. CONCLUSIONS: An in vitro-selected ribozyme can catalyze the same type of peptide-bond formation as a ribosome; the ribozyme resembles the ribosome because a very specific RNA structure is required for substrate binding and catalysis, and both amino acids are attached to nucleotides. It is intriguing that, although there are many different possible peptidyl-transferase ribozymes, the sequence and secondary structure of one is strikingly similar to the 'helical wheel' portion of 23S rRNA implicated in ribosomal peptidyl-transferase activity.  相似文献   

13.
The 3' untranslated region (UTR) of the positive-sense RNA genome of the coronavirus mouse hepatitis virus (MHV) contains sequences that are necessary for the synthesis of negative-strand viral RNA as well as sequences that may be crucial for both genomic and subgenomic positive-strand RNA synthesis. We have found that the entire 3' UTR of MHV could be replaced by the 3' UTR of bovine coronavirus (BCV), which diverges overall by 31% in nucleotide sequence. This exchange between two viruses that are separated by a species barrier was carried out by targeted RNA recombination. Our results define regions of the two 3' UTRs that are functionally equivalent despite having substantial sequence substitutions, deletions, or insertions with respect to each other. More significantly, our attempts to generate an unallowed substitution of a particular portion of the BCV 3' UTR for the corresponding region of the MHV 3' UTR led to the discovery of a bulged stem-loop RNA secondary structure, adjacent to the stop codon of the nucleocapsid gene, that is essential for MHV viral RNA replication.  相似文献   

14.
15.
16.
17.
The complete nucleotide sequence of the 16S RNA from Proteus vulgaris has been determined. The molecule (1544 nucleotides) shows 93% homology with the sequence of E. coli 16S RNA. Six methylated nucleotides have been localized in positions homologous to those observed in the E. coli RNA molecule. Both E. coli and P. vulgaris 16S RNA chains can be folded up into a common secondary structure scheme. Comparative sequence analysis of the two molecules has provided a valuable contribution to 16S RNA secondary structure model building.  相似文献   

18.
The RNA elements that are required for replication of defective interfering (DI) RNA of the JHM strain of mouse hepatitis virus (MHV) consist of three discontinuous genomic regions: about 0.46 to 0.47 kb from both terminal sequences and an internal 58-nucleotide (nt)-long sequence (58-nt region) present at about 0.9 kb from the 5' end of the DI genome. The internal region is important for positive-strand DI RNA synthesis (Y. N. Kim and S. Makino, J. Virol. 69:4963-4971, 1995). We further characterized the 58-nt region in the present study and obtained the following results. (i) The positive-strand RNA structure in solution was comparable with that predicted by computer modeling. (ii) Positive-strand RNA secondary structure, but not negative-strand RNA structure, was important for the biological function of the region. (iii) The biological function had a sequence-specific requirement. We discuss possible mechanisms by which the internal cis-acting signal drives MHV positive-strand DI RNA synthesis.  相似文献   

19.
Computer-generated thermodynamic predictions and solution structure probing indicated two stem-loop structures, stem-loop 1 (SL1; nt 32-106) and stem-loop 2 (SL2; nt 143-183), within the 5' 230 nt of potato virus X (PVX) RNA. Because the existence of SL1 was further supported by covariation analysis of several PVX strains, the functional significance of this structure was investigated by site-directed mutational analysis in a tobacco protoplast system. In general, mutations that reduced genomic plus-strand RNA accumulation similarly affected coat protein accumulation, indicating that subgenomic plus-strand RNA was also affected. In contrast, minus-strand RNA levels remained relatively unchanged. Mutational analysis of the stem C (SC) region of SL1 indicated that pairing was more important than sequence, which was consistent with the covariation analysis. Alterations that increased length and stability of either SC or stem D (SD) were deleterious to plus-strand RNA accumulation. The formation of internal loop C between SC and SD, as well as specific nucleotides within this loop, were also required. Several modifications were made to the terminal GAAA tetraloop, a motif known for enhanced RNA stability. Both GANA and GAAG motifs resulted in wild-type levels of RNA accumulation. However, a UUCG tetraloop was detrimental, indicating that the sequence of this element was important beyond just providing stabilization of the structure. These data indicate that multiple features of SL1 are critical for accumulation of PVX plus-strand RNA.  相似文献   

20.
Understanding which phenotypes are accessible from which genotypes is fundamental for understanding the evolutionary process. This notion of accessibility can be used to define a relation of nearness among phenotypes, independently of their similarity. Because of neutrality, phenotypes denote equivalence classes of genotypes. The definition of neighborhood relations among phenotypes relies, therefore, on the statistics of neighborhood relations among equivalence classes of genotypes in genotype space. The folding of RNA sequence (genotypes) into secondary structures (phenotypes) is an ideal case to implement these concepts. We study the extent to which the folding of RNA sequence induces a "statistical topology" on the set of minimum free energy secondary structures. The resulting nearness relation suggests a notion of "continuous" structure transformation. We can, then rationalize major transitions in evolutionary trajectories at the level of RNA structures by identifying those transformations which are irreducibly discontinuous. This is shown by means of computer simulations. The statistical topology organizing the set of RNA shapes explains why neutral drift in sequence space plays a key role in evolutionary optimization.  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