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 78 毫秒
1.
The first stable complex formed during the assembly of spliceosomes onto pre-mRNA substrates in mammals includes U1 snRNP, which recognizes the 5′ splice site, and the splicing factors SF1 and U2AF, which bind the branch point sequence, polypyrimidine tract, and 3′ splice site. The 5′ and 3′ splice site complexes are thought to be joined together by protein–protein interactions mediated by factors that ensure the fidelity of the initial splice site recognition. In this study, we identified and characterized PRPF40B, a putative mammalian ortholog of the U1 snRNP-associated yeast splicing factor Prp40. PRPF40B is highly enriched in speckles with a behavior similar to splicing factors. We demonstrated that PRPF40B interacts directly with SF1 and associates with U2AF65. Accordingly, PRPF40B colocalizes with these splicing factors in the cell nucleus. Splicing assays with reporter minigenes revealed that PRPF40B modulates alternative splice site selection. In the case of Fas regulation of alternative splicing, weak 5′ and 3′ splice sites and exonic sequences are required for PRPF40B function. Placing our data in a functional context, we also show that PRPF40B depletion increased Fas/CD95 receptor number and cell apoptosis, which suggests the ability of PRPF40B to alter the alternative splicing of key apoptotic genes to regulate cell survival.  相似文献   

2.
Alternative splicing is an important mechanism for the regulation of bovine papillomavirus type 1 (BPV-1) gene expression during the virus life cycle. Previous studies in our laboratory have identified two purine-rich exonic splicing enhancers (ESEs), SE1 and SE2, located between two alternative 3′ splice sites at nucleotide (nt) 3225 and nt 3605. Further analysis of BPV-1 late-pre-mRNA splicing in vitro revealed a 48-nt pyrimidine-rich region immediately downstream of SE1 that inhibits utilization of the nt 3225 3′ splice site. This inhibitory element, which we named an exonic splicing suppressor (ESS), has a U-rich 5′ end, a C-rich central part, and an AG-rich 3′ end (Z. M. Zheng, P. He, and C. C. Baker, J. Virol. 70:4691–4699, 1996). The present study utilized in vitro splicing of both homologous and heterologous pre-mRNAs to further characterize the ESS. The BPV-1 ESS was inserted downstream of the 3′ splice site in the BPV-1 late pre-mRNA, Rous sarcoma virus src pre-mRNA, human immunodeficiency virus tat-rev pre-mRNA, and Drosophila dsx pre-mRNA, all containing a suboptimal 3′ splice site, and in the human β-globin pre-mRNA, which contains a constitutive 3′ splice site. These studies demonstrated that suppression of splicing by the BPV-1 ESS requires an upstream suboptimal 3′ splice site but not an upstream ESE. Furthermore, the ESS functions when located either upstream or downstream of BPV-1 SE1. Mutational analyses demonstrated that the function of the ESS is sequence dependent and that only the C-rich region of the ESS is essential for suppression of splicing in all the pre-mRNAs tested.  相似文献   

3.
A T→G mutation at nucleotide 705 of human β-globin intron 2 creates an aberrant 5′ splice site and activates a cryptic 3′ splice site upstream. In consequence, the pre-mRNA is spliced via aberrant splice sites, despite the presence of the still functional correct sites. Surprisingly, when IVS2-705 HeLa or K562 cells were cultured at temperatures below 30°C, aberrant splicing was inhibited and correct splicing was restored. Similar temperature effects were seen for another β-globin pre-mRNA, IVS2-745, and in a construct in which a β-globin intron was inserted into a coding sequence of EGFP. Temperature-induced alternative splicing was affected by the nature of the internal aberrant splice sites flanking the correct sites and by exonic sequences. The results indicate that in the context of thalassemic splicing mutations and possibly in other alternatively spliced pre-mRNAs, temperature is one of the parameters that affect splice site selection.  相似文献   

4.
Despite a growing number of splicing mutations found in hereditary diseases, utilization of aberrant splice sites and their effects on gene expression remain challenging to predict. We compiled sequences of 346 aberrant 5′splice sites (5′ss) that were activated by mutations in 166 human disease genes. Mutations within the 5′ss consensus accounted for 254 cryptic 5′ss and mutations elsewhere activated 92 de novo 5′ss. Point mutations leading to cryptic 5′ss activation were most common in the first intron nucleotide, followed by the fifth nucleotide. Substitutions at position +5 were exclusively G>A transitions, which was largely attributable to high mutability rates of C/G>T/A. However, the frequency of point mutations at position +5 was significantly higher than that observed in the Human Gene Mutation Database, suggesting that alterations of this position are particularly prone to aberrant splicing, possibly due to a requirement for sequential interactions with U1 and U6 snRNAs. Cryptic 5′ss were best predicted by computational algorithms that accommodate nucleotide dependencies and not by weight-matrix models. Discrimination of intronic 5′ss from their authentic counterparts was less effective than for exonic sites, as the former were intrinsically stronger than the latter. Computational prediction of exonic de novo 5′ss was poor, suggesting that their activation critically depends on exonic splicing enhancers or silencers. The authentic counterparts of aberrant 5′ss were significantly weaker than the average human 5′ss. The development of an online database of aberrant 5′ss will be useful for studying basic mechanisms of splice-site selection, identifying splicing mutations and optimizing splice-site prediction algorithms.  相似文献   

5.
Splice site selection is a key element of pre-mRNA splicing. Although it is known to involve specific recognition of short consensus sequences by the splicing machinery, the mechanisms by which 5′ splice sites are accurately identified remain controversial and incompletely resolved. The human F7 gene contains in its seventh intron (IVS7) a 37-bp VNTR minisatellite whose first element spans the exon7–IVS7 boundary. As a consequence, the IVS7 authentic donor splice site is followed by several cryptic splice sites identical in sequence, referred to as 5′ pseudo-sites, which normally remain silent. This region, therefore, provides a remarkable model to decipher the mechanism underlying 5′ splice site selection in mammals. We previously suggested a model for splice site selection that, in the presence of consecutive splice consensus sequences, would stimulate exclusively the selection of the most upstream 5′ splice site, rather than repressing the 3′ following pseudo-sites. In the present study, we provide experimental support to this hypothesis by using a mutational approach involving a panel of 50 mutant and wild-type F7 constructs expressed in various cell types. We demonstrate that the F7 IVS7 5′ pseudo-sites are functional, but do not compete with the authentic donor splice site. Moreover, we show that the selection of the 5′ splice site follows a scanning-type mechanism, precluding competition with other functional 5′ pseudo-sites available on immediate sequence context downstream of the activated one. In addition, 5′ pseudo-sites with an increased complementarity to U1snRNA up to 91% do not compete with the identified scanning mechanism. Altogether, these findings, which unveil a cell type–independent 5′−3′-oriented scanning process for accurate recognition of the authentic 5′ splice site, reconciliate apparently contradictory observations by establishing a hierarchy of competitiveness among the determinants involved in 5′ splice site selection.  相似文献   

6.
The frequency distribution of mutation-induced aberrant 3′ splice sites (3′ss) in exons and introns is more complex than for 5′ splice sites, largely owing to sequence constraints upstream of intron/exon boundaries. As a result, prediction of their localization remains a challenging task. Here, nucleotide sequences of previously reported 218 aberrant 3′ss activated by disease-causing mutations in 131 human genes were compared with their authentic counterparts using currently available splice site prediction tools. Each tested algorithm distinguished authentic 3′ss from cryptic sites more effectively than from de novo sites. The best discrimination between aberrant and authentic 3′ss was achieved by the maximum entropy model. Almost one half of aberrant 3′ss was activated by AG-creating mutations and ~95% of the newly created AGs were selected in vivo. The overall nucleotide structure upstream of aberrant 3′ss was characterized by higher purine content than for authentic sites, particularly in position −3, that may be compensated by more stringent requirements for positive and negative nucleotide signatures centred around position −11. A newly developed online database of aberrant 3′ss will facilitate identification of splicing mutations in a gene or phenotype of interest and future optimization of splice site prediction tools.  相似文献   

7.
In Drosophila melanogaster, the fruitless (fru) gene controls essentially all aspects of male courtship behavior. It does this through sex-specific alternative splicing of the fru pre-mRNA, leading to the production of male-specific fru mRNAs capable of expressing male-specific fru proteins. Sex-specific fru splicing involves the choice between alternative 5′ splice sites, one used exclusively in males and the other used only in females. Here we report that the Drosophila sex determination genes transformer (tra) and transformer-2 (tra-2) switch fru splicing from the male-specific pattern to the female-specific pattern through activation of the female-specific fru 5′ splice site. Activation of female-specific fru splicing requires cis-acting tra and tra-2 repeat elements that are part of an exonic splicing enhancer located immediately upstream of the female-specific fru 5′ splice site and are recognized by the TRA and TRA-2 proteins in vitro. This fru splicing enhancer is sufficient to promote the activation by tra and tra-2 of both a 5′ splice site and the female-specific doublesex (dsx) 3′ splice site, suggesting that the mechanisms of 5′ splice site activation and 3′ splice site activation may be similar.  相似文献   

8.
Vertebrate internal exons are usually between 50 and 400 nt long; exons outside this size range may require additional exonic and/or intronic sequences to be spliced into the mature mRNA. The mouse polymeric immunoglobulin receptor gene has a 654 nt exon that is efficiently spliced into the mRNA. We have examined this exon to identify features that contribute to its efficient splicing despite its large size; a large constitutive exon has not been studied previously. We found that a strong 5′ splice site is necessary for this exon to be spliced intact, but the splice sites alone were not sufficient to efficiently splice a large exon. At least two exonic sequences and one evolutionarily conserved intronic sequence also contribute to recognition of this exon. However, these elements have redundant activities as they could only be detected in conjunction with other mutations that reduced splicing efficiency. Several mutations activated cryptic 5′ splice sites that created smaller exons. Thus, the balance between use of these potential sites and the authentic 5′ splice site must be modulated by sequences that repress or enhance use of these sites, respectively. Also, sequences that enhance cryptic splice site use must be absent from this large exon.  相似文献   

9.
Efficient splicing of the 5′-most intron of pre-mRNA requires a 5′ m7G(5′)ppp(5′)N cap, which has been implicated in U1 snRNP binding to 5′ splice sites. We demonstrate that the cap alters the kinetic profile of U1 snRNP binding, but its major effect is on U6 snRNA binding. With two alternative wild-type splice sites in an adenovirus pre-mRNA, the cap selectively alters U1 snRNA binding at the site to which cap-independent U1 snRNP binding is stronger and that is used predominantly in splicing; with two consensus sites, the cap acts on both, even though one is substantially preferred for splicing. However, the most striking quantitative effect of the 5′ cap is neither on U1 snRNP binding nor on the assembly of large complexes but on the replacement of U1 snRNP by U6 snRNA at the 5′ splice site. Inhibition of splicing by a cap analogue is correlated with the loss of U6 interactions at the 5′ splice site and not with any loss of U1 snRNP binding.  相似文献   

10.
11.
12.
Functional studies on the ATM intronic splicing processing element   总被引:1,自引:1,他引:0  
In disease-associated genes, the understanding of the functional significance of deep intronic nucleotide variants may represent a difficult challenge. We have previously reported a new disease-causing mechanism that involves an intronic splicing processing element (ISPE) in ATM, composed of adjacent consensus 5′ and 3′ splice sites. A GTAA deletion within ISPE maintains potential adjacent splice sites, disrupts a non-canonical U1 snRNP interaction and activates an aberrant exon. In this paper, we demonstrate that binding of U1 snRNA through complementarity within a ~40 nt window downstream of the ISPE prevents aberrant splicing. By selective mutagenesis at the adjacent consensus ISPE splice sites, we show that this effect is not due to a resplicing process occurring at the ISPE. Functional comparison of the ATM mouse counterpart and evaluation of the pre-mRNA splicing intermediates derived from affected cell lines and hybrid minigene assays indicate that U1 snRNP binding at the ISPE interferes with the cryptic acceptor site. Activation of this site results in a stringent 5′–3′ order of intron sequence removal around the cryptic exon. Artificial U1 snRNA loading by complementarity to heterologous exonic sequences represents a potential therapeutic method to prevent the usage of an aberrant CFTR cryptic exon. Our results suggest that ISPE-like intronic elements binding U1 snRNPs may regulate correct intron processing.  相似文献   

13.
Coffin–Lowry syndrome (CLS) is caused by mutations in the RSK2 gene encoding a protein kinase of the Ras signalling pathway. We have studied two point mutations which cause aberrant splicing but do not concern the invariant GT or AG nucleotides of splice sites. The first, an A→G transition at position +3 of the 5′ splice site of exon 6, results in vivo and in vitro in exon skipping and premature translation termination. The natural 5′ splice site, although intrinsically weak, is not transactivated under normal conditions. Consequently, replacement of an A/U by a G/U base pairing with U1 snRNA reduces its strength below a critical threshold. The second mutation, an A→G transition 11 nt upstream of exon 5, creates a new AG near the natural 3′ splice site. In vitro this synthetic 3′ AG is used exclusively by the splicing machinery. In vivo this splicing event is also observed, but is underestimated because the resulting RSK2 mRNA contains premature stop codons which trigger the nonsense-mediated decay process. We show that a particular mechanism is involved in the aberrant splicing of exon 5, implying involvement of the natural 3′ AG during the first catalytic step and the new 3′ AG during the second step. Thus, our results explain how these mutations cause severe forms of CLS.  相似文献   

14.
The guanosine-adenosine-rich exonic splicing enhancer (GAR ESE) identified in exon 5 of the human immunodeficiency virus type-1 (HIV-1) pre-mRNA activates either an enhancer-dependent 5′ splice site (ss) or 3′ ss in 1-intron reporter constructs in the presence of the SR proteins SF2/ASF2 and SRp40. Characterizing the mode of action of the GAR ESE inside the internal HIV-1 exon 5 we found that this enhancer fulfils a dual splicing regulatory function (i) by synergistically mediating exon recognition through its individual SR protein-binding sites and (ii) by conferring 3′ ss selectivity within the 3′ ss cluster preceding exon 5. Both functions depend upon the GAR ESE, U1 snRNP binding at the downstream 5′ ss D4 and the E42 sequence located between these elements. Therefore, a network of cross-exon interactions appears to regulate splicing of the alternative exons 4a and 5. As the GAR ESE-mediated activation of the upstream 3′ ss cluster also is essential for the processing of intron-containing vpu/env-mRNAs during intermediate viral gene expression, the GAR enhancer substantially contributes to the regulation of viral replication.  相似文献   

15.
Knowledge of the functional cis-regulatory elements that regulate constitutive and alternative pre-mRNA splicing is fundamental for biology and medicine. Here we undertook a genome-wide comparative genomics approach using available mammalian genomes to identify conserved intronic splicing regulatory elements (ISREs). Our approach yielded 314 ISREs, and insertions of ~70 ISREs between competing splice sites demonstrated that 84% of ISREs altered 5′ and 94% altered 3′ splice site choice in human cells. Consistent with our experiments, comparisons of ISREs to known splicing regulatory elements revealed that 40%–45% of ISREs might have dual roles as exonic splicing silencers. Supporting a role for ISREs in alternative splicing, we found that 30%–50% of ISREs were enriched near alternatively spliced (AS) exons, and included almost all known binding sites of tissue-specific alternative splicing factors. Further, we observed that genes harboring ISRE-proximal exons have biases for tissue expression and molecular functions that are ISRE-specific. Finally, we discovered that for Nova1, neuronal PTB, hnRNP C, and FOX1, the most frequently occurring ISRE proximal to an alternative conserved exon in the splicing factor strongly resembled its own known RNA binding site, suggesting a novel application of ISRE density and the propensity for splicing factors to auto-regulate to associate RNA binding sites to splicing factors. Our results demonstrate that ISREs are crucial building blocks in understanding general and tissue-specific AS regulation and the biological pathways and functions regulated by these AS events.  相似文献   

16.
17.
18.
Activation of pre-messenger RNA (pre-mRNA) splicing requires 5′ splice site recognition by U1 small nuclear RNA (snRNA), which is replaced by U5 and U6 snRNA. Here we use crosslinking to investigate snRNA interactions with the 5′ exon adjacent to the 5′ splice site, prior to the first step of splicing. U1 snRNA was found to interact with four different 5′ exon positions using one specific sequence adjacent to U1 snRNA helix 1. This novel interaction of U1 we propose occurs before U1-5′ splice site base pairing. In contrast, U5 snRNA interactions with the 5′ exon of the pre-mRNA progressively shift towards the 5′ end of U5 loop 1 as the crosslinking group is placed further from the 5′ splice site, with only interactions closest to the 5′ splice site persisting to the 5′ exon intermediate and the second step of splicing. A novel yeast U2 snRNA interaction with the 5′ exon was also identified, which is ATP dependent and requires U2-branchpoint interaction. This study provides insight into the nature and timing of snRNA interactions required for 5′ splice site recognition prior to the first step of pre-mRNA splicing.  相似文献   

19.
HPV16 E6 and E7, two viral oncogenes, are expressed from a single bicistronic pre-mRNA. In this report, we provide the evidence that the bicistronic pre-mRNA intron 1 contains three 5′ splice sites (5′ ss) and three 3′ splice sites (3′ ss) normally used in HPV16+ cervical cancer and its derived cell lines. The choice of two novel alternative 5′ ss (nt 221 5′ ss and nt 191 5′ ss) produces two novel isoforms of E6E7 mRNAs (E6*V and E6*VI). The nt 226 5′ ss and nt 409 3′ ss is preferentially selected over the other splice sites crossing over the intron to excise a minimal length of the intron in RNA splicing. We identified AACAAAC as the preferred branch point sequence (BPS) and an adenosine at nt 385 (underlined) in the BPS as a branch site to dictate the selection of the nt 409 3′ ss for E6*I splicing and E7 expression. Introduction of point mutations into the mapped BPS led to reduced U2 binding to the BPS and thereby inhibition of the second step of E6E7 splicing at the nt 409 3′ ss. Importantly, the E6E7 bicistronic RNA with a mutant BPS and inefficient splicing makes little or no E7 and the resulted E6 with mutations of 91QYNK94 to 91PSFW94 displays attenuate activity on p53 degradation. Together, our data provide structural basis of the E6E7 intron 1 for better understanding of how viral E6 and E7 expression is regulated by alternative RNA splicing. This study elucidates for the first time a mapped branch point in HPV16 genome involved in viral oncogene expression.  相似文献   

20.
During pre-mRNA splicing, the spliceosome must configure the substrate, catalyze 5′ splice site cleavage, reposition the substrate, and catalyze exon ligation. The highly conserved U2/U6 helix I, which adjoins sequences that define the reactive sites, has been proposed to configure the substrate for 5′ splice site cleavage and promote catalysis. However, a role for this helix at either catalytic step has not been tested rigorously and previous observations question its role at the catalytic steps. Through a comprehensive molecular genetic study of U2/U6 helix I, we found that weakening U2/U6 helix I, but not mutually exclusive structures, compromised splicing of a substrate limited at the catalytic step of 5′ splice site cleavage, providing the first compelling evidence that this helix indeed configures the substrate during 5′ splice site cleavage. Further, mutations that we proved weaken only U2/U6 helix I suppressed a mutation in PRP16, a DEAH-box ATPase required after 5′ splice site cleavage, providing persuasive evidence that helix I is destabilized by Prp16p and suggesting that this structure is unwound between the catalytic steps. Lastly, weakening U2/U6 helix I also compromised splicing of a substrate limited at the catalytic step of exon ligation, providing evidence that U2/U6 helix I reforms and functions during exon ligation. Thus, our data provide evidence for a fundamental and apparently dynamic role for U2/U6 helix I during the catalytic stages of splicing.  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