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 15 毫秒
1.
Chloroplast genomes in land plants harbor approximately 20 group II introns. Genetic approaches have identified proteins involved in the splicing of many of these introns, but the proteins identified to date cannot account for the large size of intron ribonucleoprotein complexes and are not sufficient to reconstitute splicing in vitro. Here, we describe an additional protein that promotes chloroplast group II intron splicing in vivo. This protein, RNC1, was identified by mass spectrometry analysis of maize (Zea mays) proteins that coimmunoprecipitate with two previously identified chloroplast splicing factors, CAF1 and CAF2. RNC1 is a plant-specific protein that contains two ribonuclease III (RNase III) domains, the domain that harbors the active site of RNase III and Dicer enzymes. However, several amino acids that are essential for catalysis by RNase III and Dicer are missing from the RNase III domains in RNC1. RNC1 is found in complexes with a subset of chloroplast group II introns that includes but is not limited to CAF1- and CAF2-dependent introns. The splicing of many of the introns with which it associates is disrupted in maize rnc1 insertion mutants, indicating that RNC1 facilitates splicing in vivo. Recombinant RNC1 binds both single-stranded and double-stranded RNA with no discernible sequence specificity and lacks endonuclease activity. These results suggest that RNC1 is recruited to specific introns via protein-protein interactions and that its role in splicing involves RNA binding but not RNA cleavage activity.  相似文献   

2.
3.
Mohr S  Stryker JM  Lambowitz AM 《Cell》2002,109(6):769-779
The Neurospora crassa CYT-18 protein, the mitochondrial tyrosyl-tRNA synthetase, functions in splicing group I introns by inducing formation of the catalytically active RNA structure. Here, we identified a DEAD-box protein (CYT-19) that functions in concert with CYT-18 to promote group I intron splicing in vivo and vitro. CYT-19 does not bind specifically to group I intron RNAs and instead functions as an ATP-dependent RNA chaperone to destabilize nonnative RNA structures that constitute kinetic traps in the CYT-18-assisted RNA-folding pathway. Our results demonstrate that a DExH/D-box protein has a specific, physiologically relevant chaperone function in the folding of a natural RNA substrate.  相似文献   

4.
Chloroplast genomes in angiosperms encode approximately 20 group II introns, approximately half of which are classified as subgroup IIB. The splicing of all but one of the subgroup IIB introns requires a heterodimer containing the peptidyl-tRNA hydrolase homolog CRS2 and one of two closely related proteins, CAF1 or CAF2, that harbor a recently recognized RNA binding domain called the CRM domain. Two CRS2/CAF-dependent introns require, in addition, a CRM domain protein called CFM2 that is only distantly related to CAF1 and CAF2. Here, we show that CFM3, a close relative of CFM2, associates in vivo with those CRS2/CAF-dependent introns that are not CFM2 ligands. Mutant phenotypes in rice and Arabidopsis support a role for CFM3 in the splicing of most of the introns with which it associates. These results show that either CAF1 or CAF2 and either CFM2 or CFM3 simultaneously bind most chloroplast subgroup IIB introns in vivo, and that the CAF and CFM subunits play nonredundant roles in splicing. These results suggest that the expansion of the CRM protein family in plants resulted in two subfamilies that play different roles in group II intron splicing, with further diversification within a subfamily to accommodate multiple intron ligands.  相似文献   

5.
The mitochondrial tyrosyl-tRNA synthetases (mt TyrRSs) of Pezizomycotina fungi are bifunctional proteins that aminoacylate mitochondrial tRNA(Tyr) and are structure-stabilizing splicing cofactors for group I introns. Studies with the Neurospora crassa synthetase (CYT-18 protein) showed that splicing activity is dependent upon Pezizomycotina-specific structural adaptations that form a distinct group I intron-binding site in the N-terminal catalytic domain. Although CYT-18's C-terminal domain also binds group I introns, it has been intractable to X-ray crystallography in the full-length protein. Here, we determined an NMR structure of the isolated C-terminal domain of the Aspergillus nidulans mt TyrRS, which is closely related to but smaller than CYT-18's. The structure shows an S4 fold like that of bacterial TyrRSs, but with novel features, including three Pezizomycontia-specific insertions. (15)N-(1)H two-dimensional NMR showed that C-terminal domains of the full-length A. nidulans and Geobacillus stearothermophilus synthetases do not tumble independently in solution, suggesting restricted orientations. Modeling onto a CYT-18/group I intron cocrystal structure indicates that the C-terminal domains of both subunits of the homodimeric protein bind different ends of the intron RNA, with one C-terminal domain having to undergo a large shift on its flexible linker to bind tRNA(Tyr) or the intron RNA on either side of the catalytic domain. The modeling suggests that the C-terminal domain acts together with the N-terminal domain to clamp parts of the intron's catalytic core, that at least one C-terminal domain insertion functions in group I intron binding, and that some C-terminal domain regions bind both tRNA(Tyr) and group I intron RNAs.  相似文献   

6.
Vogel J  Börner T 《The EMBO journal》2002,21(14):3794-3803
Lariat formation has been studied intensively only with a few self-splicing group II introns, and little is known about how the numerous diverse introns in plant organelles are excised. Several of these introns have branch-points that are not a single bulge but are adjoined by A:A, A:C, A:G and G:G pairs. Using a highly sensitive in vivo approach, we demonstrate that all but one of the barley chloroplast introns splice via the common pathway that produces a branched product. RNA editing does not improve domain 5 and 6 structures of these introns. The conserved branch-point in tobacco rpl16 is chosen even if an adjacent unpaired adenosine is available, suggesting that spatial arrangements in domain 6 determine correct branch-point selection. Lariats were not detected for the chloroplast trnV intron, which lacks an unpaired adenosine in domain 6. Instead, this intron is released as linear molecules that undergo further polyadenylation. trnV, which is conserved throughout plant evolution, constitutes the first example of naturally occurring hydrolytic group II intron splicing in vivo.  相似文献   

7.
Group II introns are catalytic RNAs that have been proposed to be the evolutionary precursors to the spliceosome. Most group II introns require accessory factors to splice efficiently in vivo, but few such factors have been identified. We have cloned the maize nuclear gene crs2, which is required for the splicing of nine group II introns in chloroplasts. CRS2 is related to peptidyl-tRNA hydrolase enzymes. However, CRS2 expression failed to rescue an Escherichia coli pth(ts) mutant and CRS2 lacks several conserved amino acids that are important for the activity of the E.coli enzyme, indicating that it may lack peptidyl-tRNA hydrolase activity. CRS2 is localized to the chloroplast stroma, where it is found in a large salt-stable complex that contains RNA. CRS2 co-sediments with group II intron RNA during centrifugation of stroma through sucrose gradients, suggesting that CRS2 facilitates splicing via direct interaction with intron RNA. Sequence comparisons indicate how evolutionary tinkering may have allowed an enzyme that interacts with peptidyl-tRNAs to acquire a function in group II intron splicing.  相似文献   

8.
A novel mechanism for protein-assisted group I intron splicing   总被引:3,自引:0,他引:3       下载免费PDF全文
Previously it was shown that the Aspergillus nidulans (A.n.) mitochondrial COB intron maturase, I-AniI, facilitates splicing of the COB intron in vitro. In this study, we apply kinetic analysis of binding and splicing along with RNA deletion analysis to gain insight into the mechanism of I-AniI facilitated splicing. Our results are consistent with I-AniI and A.n. COB pre-RNA forming a specific but labile encounter complex that is resolved into the native, splicing-competent complex. Significantly, kinetic analysis of splicing shows that the resolution step is rate limiting for splicing. RNA deletion studies show that I-AniI requires most of the A.n. COB intron for binding suggesting that the integrity of the I-AniI-binding site depends on overall RNA tertiary structure. These results, taken together with the observation that A.n. COB intron lacks significant stable tertiary structure in the absence of protein, support a model in which I-AniI preassociates with an unfolded COB intron via a "labile" interaction that facilitates correct folding of the intron catalytic core, perhaps by resolving misfolded RNAs or narrowing the number of conformations sampled by the intron during its search for native structure. The active intron conformation is then "locked in" by specific binding of I-Anil to its intron interaction site.  相似文献   

9.
10.
11.
Group II introns are autocatalytic RNAs which self-splice in vitro. However, in vivo additional protein factors might be involved in the splicing process. We used an affinity chromatography method called 'StreptoTag' to identify group II intron binding proteins from Saccharomyces cerevisiae. This method uses a hybrid RNA consisting of a streptomycin-binding affinity tag and the RNA of interest, which is bound to a streptomycin column and incubated with yeast protein extract. After several washing steps the bound RNPs are eluted by addition of streptomycin. The eluted RNPs are separated and the proteins identified by mass-spectrometric analysis. Using crude extract from yeast in combination with a substructure of the bl1 group II intron (domains IV-VI) we were able to identify four glycolytic enzymes; glucose-6-phosphate isomerase (GPI), 3-phosphoglycerate kinase (PGK), glyceraldehyde-3-phosphate dehydrogenase (GAPDH) and triosephosphate isomerase (TPI). From these proteins GAPDH increases in vitro splicing of the bl1 group II intron by up to three times. However, in vivo GAPDH is not a group II intron-splicing factor, since it is not localised in yeast mitochondria. Therefore, the observed activity reflects an unexpected property of GAPDH. Band shift experiments and UV cross linking demonstrated the interaction of GAPDH with the group II intron RNA. This novel activity expands the reaction repertoire of GAPDH to a new RNA species.  相似文献   

12.
Proteins of the DExH/D family are ATPases that can unwind duplex RNA in vitro. Individual members of this family coordinate many steps in ribonucleoprotein enzyme assembly and catalysis in vivo, but it is largely unknown how the action of these co-factors is specified and precisely timed. As a first step to address this question biochemically, we describe the development of a new protein-dependent group I intron splicing system that requires such an ATPase for coordinating successive steps in splicing. While genetic analysis in yeast has shown that at least five nuclear-encoded proteins are required for splicing of the mitochondrial aI5β group I intron, we show that efficient in vitro splicing of aI5β occurs with only two of these co-factors and, furthermore, they fulfill distinct functions in vitro. The Mrs1p protein stabilizes RNA structure and promotes the first step in splicing. In contrast, a DExH/D protein, Mss116p, acts after the first step and, utilizing ATP hydrolysis, specifically enhances the efficiency of exon ligation. An analysis of Mss116p variants with mutations that impair its RNA-stimulated ATP hydrolysis activity or reduce its ability to unwind duplexes show that the efficiency of ATP hydrolysis is a major determinant in promoting exon ligation. These observations suggest that Mss116p acts in aI5β splicing by catalyzing changes in the structure of the RNA/protein splicing intermediate that promote the second step. More broadly, these observations are consistent with a model in which the “functional-timing” of DExH/D-box protein action can be specified by a specific conformation of its substrate due to the “upstream” activity of other co-factors.  相似文献   

13.
A maize gene designated thylakoid assembly 8 (tha8) emerged from a screen for nuclear mutations that cause defects in the biogenesis of chloroplast thylakoid membranes. The tha8 gene encodes an unusual member of the pentatricopeptide repeat (PPR) family, a family of helical repeat proteins that participate in various aspects of organellar RNA metabolism. THA8 localizes to chloroplasts, where it associates specifically with the ycf3-2 and trnA group II introns. The splicing of ycf3-2 is eliminated in tha8 mutants, and trnA splicing is strongly compromised. Reverse-genetic analysis of the tha8 ortholog in Arabidopsis thaliana showed that these molecular functions are conserved, although null alleles are embryo lethal in Arabidopsis and seedling lethal in maize. Whereas most PPR proteins have more than 10 PPR motifs, THA8 belongs to a subfamily of plant PPR proteins with only four PPR motifs and little else. THA8 is the first member of this subfamily with a defined molecular function, and illustrates that even small PPR proteins have the potential to mediate specific intermolecular interactions in vivo.  相似文献   

14.
Crystal structure of a group I intron splicing intermediate   总被引:6,自引:3,他引:6  
A recently reported crystal structure of an intact bacterial group I self-splicing intron in complex with both its exons provided the first molecular view into the mechanism of RNA splicing. This intron structure, which was trapped in the state prior to the exon ligation reaction, also reveals the architecture of a complex RNA fold. The majority of the intron is contained within three internally stacked, but sequence discontinuous, helical domains. Here the tertiary hydrogen bonding and stacking interactions between the domains, and the single-stranded joiner segments that bridge between them, are fully described. Features of the structure include: (1) A pseudoknot belt that circumscribes the molecule at its longitudinal midpoint; (2) two tetraloop-tetraloop receptor motifs at the peripheral edges of the structure; (3) an extensive minor groove triplex between the paired and joiner segments, P6-J6/6a and P3-J3/4, which provides the major interaction interface between the intron's two primary domains (P4-P6 and P3-P9.0); (4) a six-nucleotide J8/7 single stranded element that adopts a mu-shaped structure and twists through the active site, making critical contacts to all three helical domains; and (5) an extensive base stacking architecture that realizes 90% of all possible stacking interactions. The intron structure was validated by hydroxyl radical footprinting, where strong correlation was observed between experimental and predicted solvent accessibility. Models of the pre-first and pre-second steps of intron splicing are proposed with full-sized tRNA exons. They suggest that the tRNA undergoes substantial angular motion relative to the intron between the two steps of splicing.  相似文献   

15.
The I-AniI maturase facilitates self-splicing of a mitochondrial group I intron in Aspergillus nidulans. Binding occurs in at least two steps: first, a specific but labile encounter complex rapidly forms and then this intermediate is slowly resolved into a native, catalytically active RNA/protein complex. Here we probe the structure of the RNA throughout the assembly pathway. Although inherently unstable, the intron core, when bound by I-AniI, undergoes rapid folding to a near-native state in the encounter complex. The next transition includes the slow destabilization and docking into the core of the peripheral stacked helix that contains the 5' splice site. Mutational analyses confirm that both transitions are important for native complex formation. We propose that protein-driven destabilization and docking of the peripheral stacked helix lead to subtle changes in the I-AniI binding site that facilitate native complex formation. These results support an allosteric-feedback mechanism of RNA-protein recognition in which proteins engaged in an intermediate complex can influence RNA structure far from their binding sites. The linkage of these changes to stable binding ensures that the protein and RNA do not get sequestered in nonfunctional complexes.  相似文献   

16.
The chloroplast RNA splicing and ribosome maturation (CRM) domain is a RNA-binding domain found in a plant-specific protein family whose characterized members play essential roles in splicing group I and group II introns in mitochondria and chloroplasts. Together, these proteins are required for splicing of the majority of the approximately 20 chloroplast introns in land plants. Here, we provide evidence from Setaria viridis and maize that an uncharacterized member of this family, CRM Family Member1 (CFM1), promotes the splicing of most of the introns that had not previously been shown to require a CRM domain protein. A Setaria mutant expressing mutated CFM1 was strongly disrupted in the splicing of three chloroplast tRNAs: trnI, trnV and trnA. Analyses by RNA gel blot and polysome association suggest that the tRNA deficiencies lead to compromised chloroplast protein synthesis and the observed whole-plant chlorotic phenotypes. Co-immunoprecipitation data demonstrate that the maize CFM1 ortholog is bound to introns whose splicing is disrupted in the cfm1 mutant. With these results, CRM domain proteins have been shown to promote the splicing of all but two of the introns found in angiosperm chloroplast genomes.  相似文献   

17.
18.
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   

19.
In mitochondria of flowering plants the nad5 open reading frame is assembled from five exons via two conventional cis-splicing and two trans-splicing events. Trans-splicing between exons c and d in wheat, petunia and Arabidopsis involves a bipartite group II intron structure, while in Oenothera a large portion of intron domains I–IV is missing from the major genomic locus. This intron region has been lost downstream of exon c and is now found in a distant genomic region. Intragenomic recombination across an 11 nucleotide sequence has separated these intron parts, which now have to be reassembled from three independent RNA precursors. This organisation coexists with highly substoichiometric copy numbers of the bipartite intron arrangement, consistent with an evolutionary origin of the tripartite intron by genomic disruption. Received: 28 August 1996 / Accepted: 11 December 1996  相似文献   

20.
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