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.
The aminoglycoside antibiotic neomycin B inhibits translation in prokaryotes and interferes with RNA-protein interactions in HIV both in vivo and in vitro. Hitherto, inhibition of ribozyme catalysis has only been observed in vitro. We therefore monitored the activity of neomycin B and several other aminoglycoside antibiotics on splicing of the T4 phage thymidylate synthase (td) intron in vivo. All antibiotics tested inhibited splicing, even chloramphenicol, which does not inhibit splicing in vitro. Splicing of the td intron in vivo requires translation for proper folding of the pre-mRNA. In the absence of translation, two interactions between sequences in the upstream exon and the 5' and 3' splice sites trap the pre-mRNA in splicing-incompetent conformations. Their disruption by mutations rendered splicing less dependent on translation and also less sensitive to neomycin B. Intron splicing was affected by neither neomycin B nor gentamicin in Escherichia coli strains carrying antibiotic-resistance genes that modify the ribosomal RNA. Taken together, this demonstrates that in vivo splicing of td intron is not directly inhibited by aminoglycosides, but rather indirectly by their interference with translation. This was further confirmed by assaying splicing of the Tetrahymena group I intron, which is inserted in the E. coli 23 S rRNA and, thus, not translated. Furthermore, neomycin B, paromomycin, and streptomycin enhanced missplicing in antibiotic-sensitive strains. Missplicing is caused by an alternative structural element containing a cryptic 5' splice site, which serves as a substrate for the ribozyme. Our results demonstrate that aminoglycoside antibiotics display different effects on ribozymes in vivo and in vitro.  相似文献   

2.
An Escherichia coli gene, stpA, has been identified and cloned based on its ability to suppress the Td- phenotype of a resident, splicing-defective phage T4 td (thymidylate synthase) gene. The stpA gene, which was localized to 60.24 min on the E. coli chromosome, encodes a 15.3-kDa protein. Overproduction of StpA in vivo led to an increase in td pre-mRNA levels and modest enhancement of td mRNA:pre-mRNA ratios. Consistent with its in vivo effect, purified StpA promoted RNA splicing in vitro, and facilitated RNA annealing and strand exchange with model substrates. These results suggest that StpA promotes splicing of the intron by binding RNA nonspecifically, resolving misfolded precursor molecules and facilitating association of critical base pair elements. Furthermore, proteinase K treatment of StpA-assembled precursors prior to the initiation of the splicing reaction still resulted in splicing enhancement, indicating that StpA is not required for the catalytic step, unlike the Neurospora splicing effector CYT-18, whose presence was necessary for catalysis to proceed. Together these results suggest that StpA has chaperone activity in vitro, with the property of promoting assembly of the precursors into an active conformation, in contrast to splicing effectors that stabilize the catalytically active intron structure.  相似文献   

3.
Here, we report an assay to evaluate the intracellular RNA chaperone activity of a protein of interest in vivo in bacterial cells. The method is based on self-splicing of the group I intron, which is located in the thymidylate synthase (td) gene of phage T4. A previously described td mutant (tdSH1) has significantly impaired splicing due to formation of splicing-incompetent alternative structures. In this procedure, overexpression of RNA chaperones in the presence of the td mutant SH1 is used to evaluate whether the putative RNA chaperone is able to rescue the incorrectly folded group I intron. The ability of the RNA chaperone to assist during folding is measured indirectly by assessing the difference between the splicing efficiencies of the td mutant in the absence and in the presence of the RNA chaperone. This procedure can be completed in 5-6 d, not including the time needed to clone the putative RNA chaperone.  相似文献   

4.
5.
Proteins with RNA chaperone activity are able to promote folding of RNA molecules by loosening their structure. This RNA unfolding activity is beneficial when resolving misfolded RNA conformations, but could be detrimental to RNAs with low thermodynamic stability. In order to test this idea, we constructed various RNAs with different structural stabilities derived from the thymidylate synthase (td) group I intron and measured the effect of StpA, an Escherichia coli protein with RNA chaperone activity, on their splicing activity in vivo and in vitro. While StpA promotes splicing of the wild-type td intron and of mutants with wild-type-like stability, splicing of mutants with a lower structural stability is reduced in the presence of StpA. In contrast, splicing of an intron mutant, which is not destabilized but which displays a reduced population of correctly folded RNAs, is promoted by StpA. The sensitivity of an RNA towards StpA correlates with its structural stability. By lowering the temperature to 25°C, a temperature at which the structure of these mutants becomes more stable, StpA is again able to stimulate splicing. These observations clearly suggest that the structural stability of an RNA determines whether the RNA chaperone activity of StpA is beneficial to folding.  相似文献   

6.
Revertants have been obtained from six mutants of the box9 cluster, which are supposed to be defective in RNA splicing as a result of alterations in a splice signal sequence. This sequence is in the 5' part of intron 4 of the cob gene, 330 to 340 bp downstream from the 5' splice site. Sequencing reveals that reversion to splicing competence is achieved by restoration of the wild-type box9 sequence; by creation of novel box9 sequences; and by introduction of a second site or suppressor mutation (sup-) compensating for the effect of the primary box9- mutation. The sup- mutation alters a sequence in intron 4,293 bp upstream from the box9- primary mutation. The box9 sequence and this upstream sequence can base pair to form an intramolecular hybrid in intron RNA in which box9- and sup- are compensatory base pair exchanges (G----A and C----U, respectively). Thus intramolecular hybrid structures of intron RNA are essential for RNA splicing.  相似文献   

7.
The intron-encoded proteins bI4 RNA maturase and aI4 DNA endonuclease can be faithfully expressed in yeast cytoplasm from engineered forms of their mitochondrial coding sequences. In this work we studied the relationships between these two activities associated with two homologous intron-encoded proteins: the bI4 RNA maturase encoded in the fourth intron of the cytochrome b gene and the aI4 DNA endonuclease (I-SceII) encoded in the fourth intron of the gene coding for the subunit I of cytochrome oxidase. Taking advantage of both the high recombinogenic properties of yeast and the similarities between the two genes, we constructed in vivo a family of hybrid genes carrying parts of both RNA maturase and DNA endonuclease coding sequences. The presence of a sequence coding for a mitochondrial targeting peptide upstream from these hybrid genes allowed us to study the properties of their translation products within the mitochondria in vivo. We thus could analyze the ability of the recombinant proteins to complement RNA maturase deficiencies in different strains. Many combinations of the two parental intronic sequences were found in the recombinants. Their structural and functional analysis revealed the following features. (i) The N-terminal half of the bI4 RNA maturase could be replaced in total by its equivalent from the aI4 DNA endonuclease without affecting the RNA maturase activity. In contrast, replacing the C-terminal half of the bI4 RNA maturase with its equivalent from the aI4 DNA endonuclease led to a very weak RNA maturase activity, indicating that this region is more differentiated and linked to the maturase activity. (ii) None of the hybrid proteins carrying an RNA maturase activity kept the DNA endonuclease activity, suggesting that the latter requires the integrity of the aI4 protein. These observations are interesting because the aI4 DNA endonuclease is known to promote the propagation, at the DNA level, of the aI4 intron, whereas the bI4 RNA maturase, which is required for the splicing of its coding intron, also controls the splicing process of the aI4 intron. We propose a scenario for the evolution of these intronic proteins that relies on a switch from DNA endonuclease to RNA maturase activity.  相似文献   

8.
Group I introns constitute excellent systems for analyzing the relationship between RNA tertiary folding and catalysis. Within a hierarchical framework interpretation of RNA folding, secondary structure motifs subtend RNA three-dimensional (3D) architecture. Thus, mutations in two-dimensional motifs are expected to have effects different from those disrupting 3D contacts. Using UV spectroscopy, we have studied the influence of nucleotide substitutions, in both secondary and tertiary structure elements, on the thermal stability of the tertiary folding of the bacteriophage T4 td group I intron. Further, we present a quantitative analysis of the relationship between the splicing efficiency in vivo and the stability of the intron structure as monitored by UV melting curves. We conclude that the stability of the tertiary structure of a group I intron as measured by UV melting is generally a good indication of its ability to splice in vivo.  相似文献   

9.
The fifth and terminal intron of yeast cytochrome b pre-mRNA (a group I intron) requires a protein encoded by the nuclear gene CBP2 for splicing. Because catalysis is intrinsic to the RNA, the protein is believed to promote formation of secondary and tertiary structure of the RNA, resulting in a catalytically competent intron. In vitro, this mitochondrial intron can be made to self-splice or undergo protein-facilitated splicing by varying the Mg(2+) and monovalent salt concentrations. This two-component system, therefore, provides a good model for understanding the role of proteins in RNA folding. A UV cross-linking experiment was initiated to identify RNA binding sites on Cbp2 and gain insights into Cbp2-intron interactions. A 12-amino acid region containing a presumptive contact site near the amino terminus was targeted for mutagenesis, and mutant proteins were characterized for RNA binding and stimulation of splicing. Mutations in this region resulted in partial or complete loss of function, demonstrating the importance of this determinant for stimulation of RNA splicing. Several of the mutations that severely reduced splicing did not significantly shift the overall binding isotherm of Cbp2 for the precursor RNA, suggesting that contacts critical for activity are not necessarily reflected in the dissociation constant. This analysis has identified a unique RNA binding motif of alternating basic and aromatic residues that is essential for protein facilitated splicing.  相似文献   

10.
RNA containing the aI3 group I intron of the yeast mitochondrial gene encoding cytochrome oxidase subunit I shows self-splicing in vitro. The excised intron, comprising 1514 nucleotides, is partially split into an upstream portion, containing the intronic reading frame, and a downstream portion, containing the typical group I conserved sequence elements. Full-length intron RNA and intron parts occur in linear and circular form. In the transesterification reactions leading to circle formation, only the guanosine nucleotide added during splicing is removed. Reincubation of isolated, complete circular intron RNA under self-splicing conditions leads to formation of free subintronic RNA circles. Under similar conditions, purified linear intron RNA gives rise to a number of circular and linear products, one of which consists of interlocked subintronic RNA circles. These observations suggest that the intron RNA possesses a dynamic structure in which subtle alterations in folding result in the formation of RNA products with different topology.  相似文献   

11.
RNA chaperone activity of the Sm-like Hfq protein   总被引:7,自引:0,他引:7  
  相似文献   

12.
Adenosine to inosine editing of mRNA from the human 5-HT2C receptor gene (HTR2C) occurs at five exonic positions (A–E) in a stable stem–loop that includes the normal 5′ splice site of intron 5 and is flanked by two alternative splice sites. Using in vitro editing, we identified a novel editing site (F) located in the intronic part of the stem–loop and demonstrated editing at this site in human brain. We have shown that in cell culture, base substitutions to mimic editing at different combinations of the six sites profoundly affect relative splicing at the normal and the upstream alternative splice site, but splicing at the downstream alternative splice site was consistently rare. Editing combinations in different splice variants from human brain were determined and are consistent with the effects of editing on splicing observed in cell culture. As RNA editing usually occurs close to exon/intron boundaries, this is likely to be a general phenomenon and suggests an important novel role for RNA editing.  相似文献   

13.
The RNA-catalyzed splicing of group I and group II introns is facilitated by proteins that stabilize the active RNA structure or act as RNA chaperones to disrupt stable inactive structures that are kinetic traps in RNA folding. In Neurospora crassa and Saccharomyces cerevisiae, the latter function is fulfilled by specific DEAD-box proteins, denoted CYT-19 and Mss116p, respectively. Previous studies showed that purified CYT-19 stimulates the in vitro splicing of structurally diverse group I and group II introns, and uses the energy of ATP binding or hydrolysis to resolve kinetic traps. Here, we purified Mss116p and show that it has RNA-dependent ATPase activity, unwinds RNA duplexes in a non-polar fashion, and promotes ATP-independent strand-annealing. Further, we show that Mss116p binds RNA non-specifically and promotes in vitro splicing of both group I and group II intron RNAs, as well as RNA cleavage by the aI5gamma-derived D135 ribozyme. However, Mss116p also has ATP hydrolysis-independent effects on some of these reactions, which are not shared by CYT-19 and may reflect differences in its RNA-binding properties. We also show that a non-mitochondrial DEAD-box protein, yeast Ded1p, can function almost as efficiently as CYT-19 and Mss116p in splicing the yeast aI5gamma group II intron and less efficiently in splicing the bI1 group II intron. Together, our results show that Mss116p, like CYT-19, can act broadly as an RNA chaperone to stimulate the splicing of diverse group I and group II introns, and that Ded1p also has an RNA chaperone activity that can be assayed by its effect on splicing mitochondrial introns. Nevertheless, these DEAD-box protein RNA chaperones are not completely interchangeable and appear to function in somewhat different ways, using biochemical activities that have likely been tuned by coevolution to function optimally on specific RNA substrates.  相似文献   

14.
J M Burke 《Gene》1988,73(2):273-294
In vivo and in vitro genetic techniques have been widely used to investigate the structure-function relationships and requirements for splicing of group-I introns. Analyses of group-I introns from extremely diverse genetic systems, including fungal mitochondria, protozoan nuclei, and bacteriophages, have yielded results which are complementary and highly consistent. In vivo genetic studies of fungal mitochondrial systems have served to identify cis-acting sequences within mitochondrial introns, and trans-acting protein products of mitochondrial and nuclear genes which are important for splicing, and to show that some mitochondrial introns are mobile genetic elements. In vitro genetic studies of the self-splicing intron within the Tetrahymena thermophila nuclear large ribosomal RNA precursor (Tetrahymena LSU intron) have been used to examine essential and nonessential RNA sequences and structures in RNA-catalyzed splicing. In vivo and in vitro genetic analysis of the intron within the bacteriophage T4 td gene has permitted the detailed examination of mutant phenotypes by analyzing splicing in vivo and self-splicing in vitro. The genetic studies combined with phylogenetic analysis of intron structure based on comparative nucleotide sequence data [Cech 73 (1988) 259-271] and with biochemical data obtained from in vitro splicing experiments have resulted in significant advances in understanding the biology and chemistry of group-I introns.  相似文献   

15.
Regulation of Rous sarcoma virus RNA splicing and stability.   总被引:40,自引:10,他引:30       下载免费PDF全文
  相似文献   

16.
In order to investigate in vivo splicing of group II introns in chloroplasts, we previously have integrated the mitochondrial intron rI1 from the green alga Scenedesmus obliquus into the Chlamydomonas chloroplast tscA gene. This construct allows a functional analysis of conserved intron sequences in vivo, since intron rI1 is correctly spliced in chloroplasts. Using site-directed mutagenesis, deletions of the conserved intron domains V and VI were performed. In another set of experiments, each possible substitution of the strictly conserved first intron nucleotide G1 was generated, as well as each possible single and double mutation of the tertiary base pairing gamma-gamma ' involved in the formation of the intron's tertiary RNA structure. In most cases, the intron mutations showed the same effect on in vivo intron splicing efficiency as they did on the in vitro self-splicing reaction, since catalytic activity is provided by the intron RNA itself. In vivo, all mutations have additional effects on the chimeric tscA -rI1 RNA, most probably due to the role played by trans -acting factors in intron processing. Substitutions of the gamma-gamma ' base pair lead to an accumulation of excised intron RNA, since intron stability is increased. In sharp contrast to autocatalytic splicing, all point mutations result in a complete loss of exon RNA, although the spliced intron accumulates to high levels. Intron degradation and exon ligation only occur in double mutants with restored base pairing between the gamma and gamma' sites. Therefore, we conclude that intron degradation, as well as the ligation of exon-exon molecules, depends on the tertiary intron structure. Furthermore, our data suggest that intron excision proceeds in vivo independent of ligation of exon-exon molecules.  相似文献   

17.
We have investigated the RNA structure of the region surrounding the muscle-specific exon 6B of the chicken beta-tropomyosin gene. We have used a variety of chemical and enzymatic probes: dimethylsulfate, N-cyclohexyl-N'-(2-(N-methylmorpholino)-ethyl)-carbodiimide-p-tolu enesulfonate) , RNase T1 and RNase V1. Lead acetate was also used to obtain some information on the tertiary structure of this region. Probing the wild-type sequence suggests a model involving one-stem and three-stem-loop structures in and around this exon. Two of these, hairpin I and stem III, have previously been implicated in repression of splicing of the intron following exon 6B in a HeLa nuclear extract. Stem I includes sequences at the beginning of exon 6B and stem III results from interaction of the intron upstream from exon 6B with sequences in the middle of the intron downstream from this exon (the intron whose splicing is repressed). Neither stem I nor stem III directly involves the consensus sequences (5' splice site, branch-point, 3' splice site) of the repressed intron. Probing RNAs that are derepressed for splicing of this intron show that there are structural changes around the 5' splice site and branch-point sequence that correlate with the derepression. This is true, despite the fact that the derepressed RNAs are altered in a region far from these consensus sequences. The most striking structural correlation with splicing capacity of the intron downstream from exon 6B is seen by probing with lead acetate. Lead ions cut RNA at specific residues; these sites are very sensitive to RNA tertiary structure. Repressed and derepressed RNAs show entirely different cleavage patterns after incubation with lead acetate. Remarkably, hybridizing a derepressed RNA with an RNA comprising the ascending arm of stem III not only re-establishes repression, but also converts the pattern of susceptibility to attack by lead ions over the whole molecule. We suggest that RNA conformation plays a role in keeping exon 6B from being spliced into non-muscle cell mRNA.  相似文献   

18.
19.
20.
CRM (chloroplast RNA splicing and ribosome maturation) is a recently recognized RNA-binding domain of ancient origin that has been retained in eukaryotic genomes only within the plant lineage. Whereas in bacteria CRM domains exist as single domain proteins involved in ribosome maturation, in plants they are found in a family of proteins that contain between one and four repeats. Several members of this family with multiple CRM domains have been shown to be required for the splicing of specific plastidic group II introns. Detailed biochemical analysis of one of these factors in maize, CRS1, demonstrated its high affinity and specific binding to the single group II intron whose splicing it facilitates, the plastid-encoded atpF intron RNA. Through its association with two intronic regions, CRS1 guides the folding of atpF intron RNA into its predicted "catalytically active" form. To understand how multiple CRM domains cooperate to achieve high affinity sequence-specific binding to RNA, we analyzed the RNA binding affinity and specificity associated with each individual CRM domain in CRS1; whereas CRM3 bound tightly to the RNA, CRM1 associated specifically with a unique region found within atpF intron domain I. CRM2, which demonstrated only low binding affinity, also seems to form specific interactions with regions localized to domains I, III, and IV. We further show that CRM domains share structural similarities and RNA binding characteristics with the well known RNA recognition motif domain.  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