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.
Fu Y  Masuda A  Ito M  Shinmi J  Ohno K 《Nucleic acids research》2011,39(10):4396-4404
In pre-mRNA splicing, a conserved AG/G at the 3'-splice site is recognized by U2AF(35). A disease-causing mutation abrogating the G nucleotide at the first position of an exon (E(+1)) causes exon skipping in GH1, FECH and EYA1, but not in LPL or HEXA. Knockdown of U2AF(35) enhanced exon skipping in GH1 and FECH. RNA-EMSA revealed that wild-type FECH requires U2AF(35) but wild-type LPL does not. A series of artificial mutations in the polypyrimidine tracts of GH1, FECH, EYA1, LPL and HEXA disclosed that a stretch of at least 10-15 pyrimidines is required to ensure normal splicing in the presence of a mutation at E(+1). Analysis of nine other disease-causing mutations at E(+1) detected five splicing mutations. Our studies suggest that a mutation at the AG-dependent 3'-splice site that requires U2AF(35) for spliceosome assembly causes exon skipping, whereas one at the AG-independent 3'-splice site that does not require U2AF(35) gives rise to normal splicing. The AG-dependence of the 3'-splice site that we analyzed in disease-causing mutations at E(+1) potentially helps identify yet unrecognized splicing mutations at E(+1).  相似文献   

2.
3.
Expression of the L1 region of adenovirus is temporally regulated by alternative splicing to yield two major RNAs encoding the 52- to 55-kilodalton (52-55K) and IIIa polypeptides. The distal acceptor site (IIIa) is utilized only during the late phase of infection, whereas the proximal site (52-55K) is used at both early and late times. Several parameters that might affect this alternative splicing were tested by using expression vectors carrying the L1 region or mutated versions of it. In the absence of a virus-encoded or -induced factor(s), only the 52-55K acceptor was used. Decreasing the distance between the donor and the IIIa acceptor had no effect. Removal of the 52-55K acceptor induced IIIa splicing slightly, implying competition between the two acceptors. Fusion of the IIIa exon to the 52-55K intron greatly enhanced splicing of the IIIa junction, suggesting that the IIIa exon does not contain sequences that inhibit splicing. Thus, the lack of splicing to the IIIa acceptor in the absence of a virus-encoded or -induced factor(s) is probably due to the absence of a favorable sequence and/or the presence of a negative element 5' of the IIIa splice junction, or both. The presence of several adenovirus gene products, including VA RNAs, the E2A DNA-binding protein, and the products of E1A and E1B genes, did not facilitate use of the IIIa acceptor. In contrast, the simian virus 40 early proteins, probably large T antigen, induced IIIa splicing. This result, together with those of earlier studies, suggest that T antigen plays a role in modulation of alternative RNA splicing.  相似文献   

4.
Measurements of the in vitro splicing efficiency of deletion mutant RNA precursors containing the small intron of the rabbit beta-globin gene, which are truncated in the first or in the second exon, revealed that no more than approximately 20 nucleotides of either exon are necessary for efficient splicing. At least for the second exon, this minimal length requirement is globin sequence-independent. Reduction of the exon-2 length to 14 nucleotides resulted in very inefficient splicing, whereas further reduction to 5 nucleotides apparently abolished the second splicing step (3' cutting and ligation), whereas the first step (5' cutting and branching) still occurred. The splicing efficiency of a double-mutant substrate retaining approximately 20 nucleotides of each exon was reduced to 50%. A kinetic study indicated that in the reaction of this double-mutant substrate the second, but not the first, splicing step was delayed, in contrast to the reaction of the wild-type precursor. Duplication or triplication of the entire sequence of exon-1 did not affect the splicing efficiency, whereas elongation of this exon with approximately 100 nucleotides of 5'-flanking (nontranscribed) beta-globin sequence diminished the level of correct splicing with the simultaneous appearance of aberrant lariat forms. We conclude that for mono-intronic precursors in which there is only one choice of splice sites, most of the exon sequences are not mechanistically involved in the splicing process.  相似文献   

5.
The essential splicing factor SF2/ASF and the heterogeneous nuclear ribonucleoprotein A1 (hnRNP A1) modulate alternative splicing in vitro of pre-mRNAs that contain 5' splice sites of comparable strengths competing for a common 3' splice site. Using natural and model pre-mRNAs, we have examined whether the ratio of SF2/ASF to hnRNP A1 also regulates other modes of alternative splicing in vitro. We found that an excess of SF2/ASF effectively prevents inappropriate exon skipping and also influences the selection of mutually exclusive tissue-specific exons in natural beta-tropomyosin pre-mRNA. In contrast, an excess of hnRNP A1 does not cause inappropriate exon skipping in natural constitutively or alternatively spliced pre-mRNAs. Although hnRNP A1 can promote alternative exon skipping, this effect is not universal and is dependent, e.g., on the size of the internal alternative exon and on the strength of the polypyrimidine tract in the preceding intron. With appropriate alternative exons, an excess of SF2/ASF promotes exon inclusion, whereas an excess of hnRNP A1 causes exon skipping. We propose that in some cases the ratio of SF2/ASF to hnRNP A1 may play a role in regulating alternative splicing by exon inclusion or skipping through the antagonistic effects of these proteins on alternative splice site selection.  相似文献   

6.
The majority of mammalian pre-mRNAs contains multiple introns that are excised prior to export and translation. After intron excision, ligated exon intermediates participate in subsequent intron excisions. However, exon ligation generates an exon of increased size, a feature of pre-mRNA splicing that can interfere with downstream splicing events. These considerations raise the question of whether unique mechanisms exist that permit efficient removal of introns neighboring ligated exons. Kinetic analyses of multiple intron-containing pre-mRNAs revealed that splicing is more efficient following an initial intron removal event, suggesting that either the recruitment of the exon junction complex (EJC) to ligated exons increases the efficiency of multiple intron excisions or that the initial definition of splice sites is sufficient to permit efficient splicing of introns neighboring ligated exons. Knockdown experiments show that the deposition of the EJC does not affect subsequent splicing kinetics. Instead, spliceosomal components that are not involved in the initial splicing event remain associated with the pre-mRNA to ensure efficient removal of neighboring introns. Thus, ligated exons do not require redefinition, providing an additional kinetic advantage for exon defined splice sites.  相似文献   

7.
B Ruskin  M R Green 《Cell》1985,43(1):131-142
Biochemical components (splicing factors) interact with specific intron regions during pre-mRNA splicing in vitro. The pre-mRNA specifically associates with factors at both the branch point and the 5' splice site and these RNA-factor interactions are maintained in the intron-containing RNA processing products. The first detectable event, the ATP-dependent association of a factor (or factors) with the branch point, is mediated by at least one factor containing an essential nucleic acid component. Mutant RNA substrates that lack either the 5' splice site or the vast majority of exon sequences can still associate with the branch point binding factor(s). However, this branch point-factor interaction does not occur with a mutant RNA substrate that contains the branch point but that lacks the 3' splice site consensus sequence. These results suggest that selection of the 3' splice site accompanied by the association of a factor with the branch point may be the initial step in mammalian pre-mRNA splicing.  相似文献   

8.
The splicing of the c-src exon N1 is controlled by an intricate combination of positive and negative RNA elements. Most previous work on these sequences focused on intronic elements found upstream and downstream of exon N1. However, it was demonstrated that the 5' half of the N1 exon itself acts as a splicing enhancer in vivo. Here we examine the function of this regulatory element in vitro. We show that a mutation in this sequence decreases splicing of the N1 exon in vitro. Proteins binding to this element were identified as hnRNP A1, hnRNP H, hnRNP F, and SF2/ASF by site-specific cross-linking and immunoprecipitation. The binding of these proteins to the RNA was eliminated by a mutation in the exonic element. The activities of hnRNP A1 and SF2/ASF on N1 splicing were examined by adding purified protein to in vitro splicing reactions. SF2/ASF and another SR protein, SC35, are both able to stimulate splicing of c-src pre-mRNA. However, splicing activation by SF2/ASF is dependent on the N1 exon enhancer element whereas activation by SC35 is not. In contrast to SF2/ASF and in agreement with other systems, hnRNP A1 repressed c-src splicing in vitro. The negative activity of hnRNP A1 on splicing was compared with that of PTB, a protein previously demonstrated to repress splicing in this system. Both proteins repress exon N1 splicing, and both counteract the enhancing activity of the SR proteins. Removal of the PTB binding sites upstream of N1 prevents PTB-mediated repression but does not affect A1-mediated repression. Thus, hnRNP A1 and PTB use different mechanisms to repress c-src splicing. Our results link the activity of these well-known exonic splicing regulators, SF2/ASF and hnRNP A1, to the splicing of an exon primarily controlled by intronic factors.  相似文献   

9.
10.
Few point mutations have been described that specifically inhibit the second step of group II intron splicing. Furthermore, the effects of such mutations are typically not apparent unless the mutations are studied in the context of a substrate that harbors a very short 5' exon. Truncation of the 5' exon slows the second step of splicing. Once the second step has been slowed, the effects of point mutations can be seen. We report the unexpected observation that the deletion of a conserved GA dinucleotide dramatically inhibits the second step of splicing, even when the mutation is studied in the context of a full-length substrate. In contrast, we find that this mutation does not significantly affect the first step of splicing, unless the mutation is studied in combination with a second point mutation that is known to inhibit the first step. Even in that context, the effect of the GA deletion mutation on the first step is modest. These observations, together with the inferred location of the GA dinucleotide in the three-dimensional structure of the intron, suggest that this dinucleotide plays a particularly important role in the second step of splicing.  相似文献   

11.
12.
In vitro splicing was studied for a human ribosomal protein (rp) S26 pre-mRNA fragment containing the first exon, first intron, and a part of the second exon. Splicing yielded two products, the first was corresponded to a fragment of the mature rpS26 mRNA and another was retained the 19 3'-terminal nucleotides of the first intron between the first and second exons. Recombinant rpS26 inhibites generation of both splicing products in vitro. The inhibition was specific, because another recombinant human rp, S19, had no effect on the splicing of the pre-mRNA fragment. Toe-printing was used to map the spS26-binding sites of the per-mRNA within the regions of the conventional and alternative 3' splicing sites of the first intron. On the strength of the rusults, rpS26 was assumed to regulate the expression of its own gene at the level of pre-mRNA splicing via a feedback mechanism.  相似文献   

13.
14.
15.
16.
The alternative exon 5 of the striated muscle-specific cardiac troponin T (cTNT) gene is included in mRNA from embryonic skeletal and cardiac muscle and excluded in mRNA from the adult. The embryonic splicing pattern is reproduced in primary skeletal muscle cultures for both the endogenous gene and transiently transfected minigenes, whereas in nonmuscle cell lines, minigenes express a default exon skipping pattern. Using this experimental system, we previously showed that a purine-rich splicing enhancer in the alternative exon functions as a constitutive splicing element but not as a target for factors regulating cell-specific splicing. In this study, we identify four intron elements, one located upstream,and three located downstream of the alternative exon, which act in a positive manner to mediate the embryonic splicing pattern of exon inclusion. Synergistic interactions between at least three of the four elements are necessary and sufficient to regulate splicing of a heterologous alternative exon and heterologous splice sites. Mutations in these elements prevent activation of exon inclusion in muscle cells but do not affect the default level of exon inclusion in nonmuscle cells. Therefore, these elements function as muscle-specific splicing enhancers (MSEs) and are the first muscle-specific positive-acting splicing elements to be described. One MSE located downstream from the alternative exon is conserved in the rat and chicken cTNT genes. A related sequence is found in a third muscle-specific gene, that encoding skeletal troponin T, downstream from an alternative exon with a developmental pattern of alternative splicing similar to that of rat and chicken cTNT. Therefore, the MSEs identified in the cTNT gene may play a role in developmentally regulated alternative splicing in a number of different genes.  相似文献   

17.
H K?nig  H Ponta  P Herrlich 《The EMBO journal》1998,17(10):2904-2913
Alternative splicing of pre-mRNA is a fundamental mechanism of differential gene expression in that it can give rise to functionally distinct proteins from a single gene, according to the developmental or physiological state of cells in multicellular organisms. In the pre-mRNA of the cell surface molecule CD44, the inclusion of up to 10 variant exons (v1-v10) is regulated during development, upon activation of lymphocytes and dendritic cells, and during tumour progression. Using minigene constructs containing CD44 exon v5, we have discovered exonic RNA elements that couple signal transduction to alternative splicing. They form a composite splice regulator encompassing an exon recognition element and splice silencer elements. Both type of elements are necessary to govern cell type-specific inclusion of the exon as well as inducible inclusion in T cells after stimulation by concanavalin A, by Ras signalling or after activation of protein kinase C by phorbol ester. Inducible splicing does not depend on de novo protein synthesis. The coupling of signal transduction to alternative splicing by such elements probably represents the mechanism whereby splice patterns of genes are established during development and can be changed under physiological and pathological conditions.  相似文献   

18.
19.
20.
Alternative splicing of mRNA allows many gene products with different functions to be produced from a single coding sequence. Exon skipping is the most commonly known alternative splicing mechanism. A comprehensive database of alternative splicing by exon skipping is made available for the human genome data. 1,229 human genes are identified to exhibit alternative splicing by exon skipping. Availability: http://sege.ntu.edu.sg/wester/ashes/.  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