首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
2.
In coliphage MS2 RNA a long-distance interaction (LDI) between an internal segment of the upstream coat gene and the start region of the replicase gene prevents initiation of replicase synthesis in the absence of coat gene translation. Elongating ribosomes break up the repressor LDI and thus activate the hidden initiation site. Expression studies on partial MS2 cDNA clones identified base pairing between 1427-1433 and 1738-1744, the so-called Min Jou (MJ) interaction, as the molecular basis for the long-range coupling mechanism. Here, we examine the biological significance of this interaction for the control of replicase gene translation. The LDI was disrupted by mutations in the 3'-side and the evolutionary adaptation was monitored upon phage passaging. Two categories of pseudorevertants emerged. The first type had restored the MJ interaction but not necessarily the native sequence. The pseudorevertants of the second type acquired a compensatory substitution some 80 nt downstream of the MJ interaction that stabilizes an adjacent LDI. In one examined case we confirmed that the second site mutations had restored coat-replicase translational coupling. Our results show the importance of translational control for fitness of the phage. They also reveal that the structure that buries the replicase start extends to structure elements bordering the MJ interaction.  相似文献   

3.
Crystal structure of an RNA bacteriophage coat protein-operator complex   总被引:2,自引:0,他引:2  
The RNA bacteriophage MS2 is a convenient model system for the study of protein-RNA interactions. The MS2 coat protein achieves control of two distinct processes--sequence-specific RNA encapsidation and repression of replicase translation--by binding to an RNA stem-loop structure of 19 nucleotides containing the initiation codon of the replicase gene. The binding of a coat protein dimer to this hairpin shuts off synthesis of the viral replicase, switching the viral replication cycle to virion assembly rather than continued replication. The operator fragment alone can trigger self-assembly of the phage capsid at low protein concentrations and a complex of about 90 RNA operator fragments per protein capsid has been described. We report here the crystal structure at 3.0 A resolution of a complex between recombinant MS2 capsids and the 19-nucleotide RNA fragment. It is the first example of a structure at this resolution for a sequence-specific protein-RNA complex apart from the transfer RNA synthetase complexes. The structure shows sequence-specific interactions between conserved residues on the protein and RNA bases essential for binding.  相似文献   

4.
The uncB gene codes for the a subunit of the Fo proton channel sector of the Escherichia coli F1 Fo ATPase. Control of expression of uncB appears to be exerted at some step after translational initiation. Sequence analysis by the perceptron matrices (G. D. Stormo, T. D. Schneider, L. Gold, and A. Ehrenfeucht, Nucleic Acids Res. 10:2997-3011, 1982) identified a potential ribosome binding site within the uncB reading frame preceding a five-codon reading frame which is shifted one base relative to the uncB reading frame. Elimination of this binding site by mutagenesis resulted in a four- to fivefold increase in expression of an uncB'-'lacZ fusion gene containing most of uncB. Primer extension inhibition (toeprint) analysis to measure ribosome binding demonstrated that ribosomes could form an initiation complex at this alternative start site. Two fusions of lacZ to the alternative reading frame demonstrated that this site is recognized by ribosomes in vivo. The results suggest that expression of uncB is reduced by translational frameshifting and/or a translational false start at this site within the uncB reading frame.  相似文献   

5.
The crystal structure of bacteriophage Q beta at 3.5 A resolution   总被引:1,自引:0,他引:1  
BACKGROUND: The capsid protein subunits of small RNA bacteriophages form a T = 3 particle upon assembly and RNA encapsidation. Dimers of the capsid protein repress translation of the replicase gene product by binding to the ribosome binding site and this interaction is believed to initiate RNA encapsidation. We have determined the crystal structure of phage Q beta with the aim of clarifying which factors are the most important for particle assembly and RNA interaction in the small phages. RESULTS: The crystal structure of bacteriophage Q beta determined at 3.5 A resolution shows that the capsid is stabilized by disulfide bonds on each side of the flexible loops that are situated around the fivefold and quasi-sixfold axes. As in other small RNA phages, the protein capsid is constructed from subunits which associate into dimers. A contiguous ten-stranded antiparallel beta sheet facing the RNA is formed in the dimer. The disulfide bonds lock the constituent dimers of the capsid covalently in the T = 3 lattice. CONCLUSIONS: The unusual stability of the Q beta particle is due to the tight dimer interactions and the disulfide bonds linking each dimer covalently to the rest of the capsid. A comparison with the structure of the related phage MS2 shows that although the fold of the Q beta coat protein is very similar, the details of the protein-protein interactions are completely different. The most conserved region of the protein is at the surface, which, in MS2, is involved in RNA binding.  相似文献   

6.
Crystal structures of two complexes between recombinant MS2 capsids and RNA operator fragments have been determined at 2.7 A resolution. The coat protein of the RNA bacteriophage MS2 is bifunctional; it forms the icosahedral virus shell to protect the viral nucleic acid and it acts as a translational repressor by binding with high specificity to a unique site on the RNA, a single stem-loop structure, containing the initiation codon of the gene for the viral replicase. In order to determine the structure of these protein-RNA complexes, we have used chemically synthesized variants of the stem-loop fragment and soaked them into crystals of recombinant capsids. The RNA stem-loop, as bound to the protein, forms a crescent-like structure and interacts with the surface of the beta-sheet of a coat protein dimer. It makes protein contacts with seven phosphate groups on the 5' side of the stem-loop, with a pyrimidine base at position -5, which stacks onto a tyrosine, and with two exposed adenine bases, one in the loop and one at a bulge in the stem. Replacement of the wild-type uridine with a cytosine at position -5 increases the affinity of the RNA to the dimer significantly. The complex with RNA stem-loop having cytosine at this position differs from that of the wild-type complex mainly by having one extra intramolecular RNA interaction and one extra water-mediated hydrogen bond.  相似文献   

7.
The mRNA of vascular endothelial growth factor (VEGF), the major angiogenic growth factor, contains an unusually long (1,038 nucleotides) and structured 5' untranslated region (UTR). According to the classical translation initiation model of ribosome scanning, such a 5' UTR is expected to be a strong translation inhibitor. In vitro and bicistronic strategies were used to show that the VEGF mRNA translation was cap independent and occurred by an internal ribosome entry process. For the first time, we demonstrate that two independent internal ribosome entry sites (IRESs) are present in this 5' UTR. IRES A is located within the 300 nucleotides upstream from the AUG start codon. RNA secondary structure prediction and site-directed mutagenesis allowed the identification of a 49-nucleotide structural domain (D4) essential to IRES A activity. UV cross-linking experiments revealed that IRES A activity was correlated with binding of a 100-kDa protein to the D4 domain. IRES B is located in the first half of the 5' UTR. An element between nucleotides 379 and 483 is required for its activity. Immunoprecipitation experiments demonstrated that a main IRES B-bound protein was the polypyrimidine tract binding protein (PTB), a well-known regulator of picornavirus IRESs. However, we showed that binding of the PTB on IRES B does not seem to be correlated with its activity. Evidence is provided of an original cumulative effect of two IRESs, probably controlled by different factors, to promote an efficient initiation of translation at the same AUG codon.  相似文献   

8.
In addition to the two usual eukaryotic elongation factors (EF-1 alpha and EF-2) fungal ribosomes need a third protein, elongation factor 3, for translation. EF-3 is essential for in vivo and in vitro protein synthesis. Functionally, EF-3 stimulates EF-1 alpha dependent binding of aminoacyl-tRNA to the ribosomal A site when E site is occupied by deacylated tRNA. EF-3 has intrinsic ATPase activity which is regulated by the functional state of the ribosome. EF-3 ATPase is activated by both 40S and 60S ribosomal subunits. However intact 80S ribosomes are needed for efficient activation of EF-3 ATPase. EF-3 appears to be an RNA binding protein with high affinity for polynucleotides containing guanosine rich sequences. To determine whether guanosine rich sequence of ribosomal RNA is involved in EF-3 binding, an antisense oligonucleotide dC6 was used to block EF-3 interaction with the ribosome. The oligonucleotide suppresses activation of EF-3 ATPase by 40S ribosomal subunit and not by the 60S or the 80S particles. Poly(U)-directed polyphenylalanine synthesis by yeast ribosomes is inhibited by dC6. To define the binding site of the oligonucleotide and presumably of EF-3 on 18S ribosomal RNA, hydrolysis of rRNA by RNase H was followed in the presence of dC6. These experiments reveal an RNase H cleavage site at 1094GGGGGG1099 sequence of 18S ribosomal RNA. This guanosine rich sequence of rRNA is suggested to be involved in EF-3 binding to yeast ribosome. Data presented in this communication suggest that the activity of EF-3 involved a direct interaction with the guanosine rich sequence of rRNA.  相似文献   

9.
10.
Antisense RNAs in prokaryotic systems often inhibit translation of mRNAs. In some cases, this involves sequestration of Shine-Dalgarno (SD) sequences and start codons. In other cases, antisense/target RNA duplexes do not overlap these signals, but form upstream. We have performed toeprinting analyses on repA mRNA of plasmid R1, both free and in duplex with the antisense RNA, CopA. An intermolecular RNA duplex 2 nt upstream of the tap SD prevents ribosome binding. An intrastrand stem-loop at this location yields the same inhibition. Thus, stable secondary structures immediately upstream of the tap SD sequence inhibit translation, as shown by toeprinting in vitro and repA-lacZ expression in vivo. Previous work showed that repA (initiator protein) expression requires tap (leader peptide) translation. Toeprinting data confirm that the tap ribosome binding site (RBS) is accessible, whereas the repA RBS, which is sequestered by a stable stem-loop, is weakly recognized by the ribosome. Truncated CopA RNA (CopI) is unable to pair completely with target RNA, but proceeds normally to a kissing intermediate. This mutant RNA species inhibits repA expression in vivo. By a kinetic toeprint inhibition protocol, we have shown that the structure of the kissing complex is sufficient to sterically prevent ribosome binding. These results are discussed in comparison with the effect of RNA structures elsewhere in the ribosome-binding region of an mRNA.  相似文献   

11.
AUG-unrelated translation initiation was found in an insect picorna-like virus, Plautia stali intestine virus (PSIV). The positive-strand RNA genome of the virus contains two nonoverlapping open reading frames (ORFs). The capsid protein gene is located in the 3'-proximal ORF and lacks an AUG initiation codon. We examined the translation mechanism and the initiation codon of the capsid protein gene by using various dicistronic and monocistronic RNAs in vitro. The capsid protein gene was translated cap independently in the presence of the upstream cistron, indicating that the gene is translated by internal ribosome entry. Deletion analysis showed that the internal ribosome entry site (IRES) consisted of approximately 250 bases and that its 3' boundary extended slightly into the capsid-coding region. The initiation codon for the IRES-mediated translation was identified as the CUU codon, which is located just upstream of the 5' terminus of the capsid-coding region by site-directed mutagenesis. In vitro translation assays of monocistronic RNAs lacking the 5' part of the IRES showed that this CUU codon was not recognized by scanning ribosomes. This suggests that the PSIV IRES can effectively direct translation initiation without stable codon-anticodon pairing between the initiation codon and the initiator methionyl-tRNA.  相似文献   

12.
The human proto-oncogene c-myc encodes two proteins, c-Myc1 and c-Myc2, from two initiation codons, CUG and AUG, respectively. It is also transcribed from four alternative promoters (P0, P1, P2, and P3), giving rise to different RNA 5'-leader sequences, the long sizes of which suggest that they must be inefficiently translated by the classical ribosome scanning mechanism. Here we have examined the influence of three c-myc mRNA 5'-leaders on the translation of chimeric myc-CAT mRNAs. We observed that in the reticulocyte rabbit lysate, these 5'-leaders lead to cap-independent translation initiation. To determine whether this kind of initiation resulted from the presence of an internal ribosome entry site (IRES), COS-7 cells were transfected with bicistronic vectors containing the different c-myc 5'-leaders in the intercistronic region. An IRES was identified, requiring elements located within the P2 leader, between nucleotides -363 and -94 upstream from the CUG start codon. This is the first demonstration of the existence of IRES-dependent translation for a proto-oncogene. This IRES could be a translation enhancer, allowing activation of c-myc expression under the control of trans-acting factors and in response to specific cell stimuli.  相似文献   

13.
14.
Expression from the secA gene, encoding a key component of the general secretory pathway of Escherichia coli, is influenced by the secretion status of the cell, autogenous translational repression, and translational coupling to the upstream gene, X. SecA binds to its mRNA in a region overlapping its ribosome binding site, thus competing with ribosomes that would initiate secA translation. Mapping of the geneX-secA mRNA secondary structure has demonstrated that the RNA can adopt two distinct conformations in solution. The first conformation arises from the base-pairing of the secA Shine-Dalgarno (SD) sequence with the geneX terminus. The second conformation, in which the secA SD sequence is no longer paired with the geneX terminus, contains a GC-rich stem upstream of the secA SD sequence. The presence of this GC-rich stem is supported by structure mapping of a mutant RNA containing a deletion in the geneX terminus. The former structure appears to be involved in translational coupling by directly linking the geneX and secA sequences, where geneX translation activates secA translational initiation through the unpairing and unmasking of the secA SD sequence. As indicated by SecA-RNA binding assays, the latter structure is probably involved in SecA binding and translational repression of the secA gene. The stabilizing effect of magnesium ions toward occlusion of the secA SD sequence supports the presence of RNA tertiary structure in this regulatory domain.  相似文献   

15.
16.
catA86 is the second gene in a constitutively transcribed, two-gene operon cloned from Bacillus pumilus . The region that intervenes between the upstream gene, termed the leader, and the catA86 coding sequence contains a pair of inverted repeat sequences which cause sequestration of the catA86 ribosome binding site in mRNA secondary structure. As a consequence, the catA86 coding sequence is untranslatable in the absence of inducer. Translation of the catA86 coding sequence is induced by chloramphenicol in Gram-positives and induction requires a function of the leader coding sequence. The leader-encoded peptide has been proposed to instruct its translating ribosome to pause at leader codon 6, enabling chloramphenicol to stall the ribosome at that site. Ribosome stalling causes destabilization of the RNA secondary structure, exposing the catA86 ribosome binding site, allowing activation of its translation. A comparable mechanism of induction by chloramphenicol has been proposed for the regulated cmlA gene from Gram-negative bacteria. The catA86 and cmlA leader-encoded peptides are in vitro inhibitors of peptidyl transferase, which is thought to be the basis for selection of the site of ribosome stalling. Both leader-encoded peptides have been shown to alter the secondary structure of Escherichia coli 23S rRNA in vitro. All peptide-induced changes in rRNA conformation are within domains IV and V, which contains the peptidyl transferase center. Here we demonstrate that the leader peptides alter the conformation of domains IV and V of large subunit rRNA from yeast and a representative of the Archaea. The rRNA target for binding the leader peptides is therefore conserved across kingdoms.  相似文献   

17.
The regulation of the Escherichia coli secA gene, whose translation is auto-repressed except when protein secretion becomes limiting, was investigated using a combination of genetic and biochemical approaches. Oligonucleotide-directed deletion and point mutagenesis was used to show that only the last quarter of the upstream gene, geneX, and the geneX-secA intergenic are essential for proper regulation. This region previously shown to contain a secretion-responsive element contains two predicted helices, helix I and II, the latter of which would occlude the secA Shine-Dalgarno sequence. Mutations that destabilized the lower portion of helix II increased secA basal expression, reduced auto-repression by SecA protein, but retained a normal pattern of derepression of secA expression during a protein export block. The introduction of compensatory mutations into helix II that were predicted to restore base-pairing restored secA regulation to wild-type levels or nearly so, suggesting that this helix does play a role in secA auto-regulation in vivo. In contrast, mutations in the lower portion of helix I decreased secA basal expression, reduced auto-repression by SecA protein, and abolished the responsiveness of secA expression to a protein export block. In this latter case introduction of compensatory mutations into helix I that were predicted to restore base-pairing did not restore proper secA regulation, indicating that specific nucleotides in this region are required for normal secA regulation. Primer-extension inhibition (toeprint) analysis with 30 S ribosoma subunits, tRNAMet, and a model segment of geneX-secA RNA carrying the relevant mutations was used to show that mutations that destabilized helix II increased ribosome binding at the secA translation initiation site, while mutations that perturbed helix I decreased ribosome binding at this site. Our results suggest strongly that there is a system of dual regulation of secA translation, whereby helix I serves as an activator element while helix II serves as a repressor element.  相似文献   

18.
19.
We have analyzed both conformational and functional changes caused by two large cis-acting deletions (delta 159 and delta 549) located within the read-through domain, a 850 nucleotide hairpin, in coliphage Q beta genomic RNA. Studies in vivo show that co-translational regulation of the viral coat and replicase genes has been uncoupled in viral genomes carrying deletion delta 159. Translational regulation is restored in deletion delta 549, a naturally evolved pseudorevertant. Structural analysis by computer modeling shows that structural features within the read-through domain of delta 159 RNA are less well determined than they are in the read-through domain of wild-type RNA, whereas predicted structure in the read-through domain of evolved pseudorevertant delta 549 is unusually well determined. Structural analysis by electron microscopy of the genomic RNAs shows that several long range helices at the base of the read-through domain, that suppress translational initiation of the viral replicase gene in the wild-type genome, have been destabilized in delta 159 RNA. In addition, the structure of local hairpins within the read-through region is more variable in delta 159 RNA than in wild-type RNA. Stable RNA secondary structure is restored in the read-through domain of delta 549 RNA. Our analyses suggest that structure throughout the read-through domain affects the regulation of viral replicase expression by altering the likelihood that long-range interactions at the base of the domain will form. We discuss possible kinetic and equilibrium models that can explain this effect, and argue that observed changes in structural plasticity within the read-through domain of the mutant genomes are key in understanding the process. During the course of these studies, we became aware of the importance of the information contained in the energy dot plot produced by the RNA secondary structure prediction program mfold. As a result, we have improved the graphical representation of this information through the use of color annotation in the predicted optimal folding. The method is presented here for the first time.  相似文献   

20.
Hepatitis C virus (HCV) is a positive-strand RNA virus whose genome is replicated by a direct RNA-to-RNA mechanism. Initiation of negative-strand RNA synthesis is believed to proceed from the 3' end of the genomic RNA. The high conservation of the 3' terminus suggests that this region directs the assembly of proteins required for the initiation of RNA replication. We sought to determine whether host proteins bind specifically to this RNA structure. We observed specific binding of cellular proteins to labeled 3'-terminal RNA by mobility shift analysis. UV crosslinking revealed that the predominant 3'-terminal RNA-binding protein migrates as a single, 60-kDa species that can be precipitated by monoclonal antibodies directed against heterogeneous nuclear ribonucleoprotein I, also called polypyrimidine tract-binding protein (hnRNP-I/PTB), a protein previously shown to bind to the 5' internal ribosome entry site (IRES) of the HCV genome. Purified hnRNP-I/PTB also bound selectively to the 3' end of the HCV genome. hnRNP-I/PTB binding requires the upstream two stem-loop structures (SL2 and SL3) but not the most 3'-terminal stem-loop (SL1). Minor alteration of either the stem or loop sequences in SL2 or SL3 severely compromised hnRNP-I/PTB binding, suggesting extremely tight RNA structural requirements for interaction with this protein. hnRNP-I/PTB does not bind to either end of the antigenomic RNA strand and binds to the 5' IRES element of the genome at least 10-fold less avidly than to the 3' terminus. The strong, selective, and preferential binding of hnRNP-I/PTB to the 3' end of the HCV genome suggests that it may be recruited to participate in viral replication, helping to direct initiation of negative-strand RNA synthesis, stabilize the viral genome, and/or regulate encapsidation of genomic RNA.  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