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 33 毫秒
1.
2.
3.
During polyadenylation of mRNA precursors in metazoan cells, poly(A) polymerase is stimulated by the nuclear poly(A) binding protein PABPN1. We report that stimulation depends on binding of PABPN1 to the substrate RNA directly adjacent to poly(A) polymerase and results in an approximately 80-fold increase in the apparent affinity of poly(A) polymerase for RNA without significant effect on catalytic efficiency. PABPN1 associates directly with poly(A) polymerase either upon allosteric activation by oligo(A) or, in the absence of RNA, upon deletion of its N-terminal domain. The N-terminal domain of PABPN1 may function to inhibit undesirable interactions of the protein; the inhibition is relieved upon RNA binding. Tethering of poly(A) polymerase is mediated largely by the C-terminal domain of PABPN1 and is necessary but not sufficient for stimulation of the enzyme; an additional interaction dependent on a coiled-coil structure located within the N-terminal domain of PABPN1 is required for a productive interaction.  相似文献   

4.
5.
6.
Oculopharyngeal muscular dystrophy (OPMD), a late-onset disorder characterized by progressive degeneration of specific muscles, results from the extension of a polyalanine tract in poly(A) binding protein nuclear 1 (PABPN1). While the roles of PABPN1 in nuclear polyadenylation and regulation of alternative poly(A) site choice are established, the molecular mechanisms behind OPMD remain undetermined. Here, we show, using Drosophila and mouse models, that OPMD pathogenesis depends on affected poly(A) tail lengths of specific mRNAs. We identify a set of mRNAs encoding mitochondrial proteins that are down-regulated starting at the earliest stages of OPMD progression. The down-regulation of these mRNAs correlates with their shortened poly(A) tails and partial rescue of their levels when deadenylation is genetically reduced improves muscle function. Genetic analysis of candidate genes encoding RNA binding proteins using the Drosophila OPMD model uncovers a potential role of a number of them. We focus on the deadenylation regulator Smaug and show that it is expressed in adult muscles and specifically binds to the down-regulated mRNAs. In addition, the first step of the cleavage and polyadenylation reaction, mRNA cleavage, is affected in muscles expressing alanine-expanded PABPN1. We propose that impaired cleavage during nuclear cleavage/polyadenylation is an early defect in OPMD. This defect followed by active deadenylation of specific mRNAs, involving Smaug and the CCR4-NOT deadenylation complex, leads to their destabilization and mitochondrial dysfunction. These results broaden our understanding of the role of mRNA regulation in pathologies and might help to understand the molecular mechanisms underlying neurodegenerative disorders that involve mitochondrial dysfunction.  相似文献   

7.
8.
The nuclear poly(A) binding protein, PABPN1, promotes mRNA polyadenylation in the cell nucleus by increasing the processivity of poly(A) polymerase and contributing to poly(A) tail length control. In its C-terminal domain, the protein carries 13 arginine residues that are all asymmetrically dimethylated. The function of this modification in PABPN1 has been unknown. Part of the methylated domain serves as nuclear localization signal, binding the import receptor transportin. Here we report that arginine methylation weakens the affinity of PABPN1 for transportin. Recombinant, unmethylated PABPN1 binds more strongly to transportin than its methylated counterpart from mammalian tissue, and in vitro methylation reduces the affinity. Transportin and RNA compete for binding to PABPN1. Methylation favors RNA binding. Transportin also inhibits in vitro methylation of the protein. Finally, a peptide corresponding to the nuclear localization signal of PABPN1 competes with transportin-dependent nuclear import of the protein in a permeabilized cell assay and does so less efficiently when it is methylated. We hypothesize that transportin binding might delay methylation of PABPN1 until after nuclear import. In the nucleus, arginine methylation may favor the transition of PABPN1 to the competing ligand RNA and serve to reduce the risk of the protein being reexported to the cytoplasm by transportin.  相似文献   

9.
Poly (A) binding proteins are intimately implicated in controlling a number of events in mRNA metabolism from nuclear polyadenylation to cytoplasmic translation and stability. The known poly(A) binding proteins can be divided into three distinct structural groups (prototypes PABP1, PABPN1/PABP2 and Nab2p) and two functional families, showing that similar functions can be accomplished by differing structural units. This has prompted us to perform a screen for novel poly(A) binding proteins using Xenopus laevis. A novel poly(A) binding protein of 32 kDa (p32) was identified. Sequence analysis showed that p32 has about 50% identity to the known nuclear poly(A) binding proteins (PABPN1) but is more closely related to a group of mammalian proteins of unknown function. The expression of Xenopus laevis ePABP2 is restricted to early embryos. Accordingly, we propose that p32 is the founder member of a novel class of poly(A) binding proteins named ePABP2.  相似文献   

10.
11.
Poly(A) tails of mRNAs are synthesized in the cell nucleus with a defined length, ∼250 nucleotides in mammalian cells. The same type of length control is seen in an in vitro polyadenylation system reconstituted from three proteins: poly(A) polymerase, cleavage and polyadenylation specificity factor (CPSF), and the nuclear poly(A)-binding protein (PABPN1). CPSF, binding the polyadenylation signal AAUAAA, and PABPN1, binding the growing poly(A) tail, cooperatively stimulate poly(A) polymerase such that a complete poly(A) tail is synthesized in one processive event, which terminates at a length of ∼250 nucleotides. We report that PABPN1 is required to restrict CPSF binding to the AAUAAA sequence and to permit the stimulation of poly(A) polymerase by AAUAAA-bound CPSF to be maintained throughout the elongation reaction. The stimulation by CPSF is disrupted when the poly(A) tail has reached a length of ∼250 nucleotides, and this terminates processive elongation. PABPN1 measures the length of the tail and is responsible for disrupting the CPSF-poly(A) polymerase interaction.The poly(A) tails present at the 3′ end of almost all eukaryotic mRNAs have two major functions. The first function is in the control of mRNA decay; degradation of the poly(A) tail by a 3′ exonuclease (deadenylation) is the first step in both of the two main pathways of mRNA decay, and the completion of deadenylation triggers the second step, either cap hydrolysis or further 3′–5′ degradation. Because the rate of deadenylation is governed by sequence elements in the mRNA, it is specific for each mRNA species and serves as a major determinant of mRNA half-life (13). Obviously, a control of mRNA stability by the rate of deadenylation requires a defined poly(A) length as a starting point. The second function of the poly(A) tail is in the initiation of translation; the cytoplasmic poly(A)-binding protein associated with the poly(A) tail promotes the initiation of translation by an interaction with the initiation factor eIF4G and probably through additional mechanisms (47). In this process, poly(A) tail length can also be important. For example, gene regulation during oocyte maturation and early embryonic development of animals depends on translational regulation of maternal mRNAs, and changes in poly(A) tail lengths of specific mRNAs, determined both by deadenylation and by regulated cytoplasmic poly(A) extension, play a major role in this translational regulation. Long poly(A) tails favor translation, whereas a shortening of the tail promotes translational inactivation of the message (8, 9). Similar mechanisms seem to operate in neurons (10, 11) and possibly in other somatic cells (12).Because the length of the poly(A) tail is important for its function, it is not surprising that poly(A) tails are generally synthesized with a defined length, which is species-specific, ∼70–90 nucleotides in Saccharomyces cerevisiae (13, 14) and ∼250 nucleotides in mammalian cells (15). Subtle differences between newly made poly(A) tails of different mRNAs have been described (13), and there is even a class of mRNAs that never receives more than an oligo(A) tail (16, 17). However, the heterogeneous length distribution seen in the steady-state mRNA population is the result of cytoplasmic shortening starting from a relatively well defined initial tail length; heterogeneity of tail length reflects age differences of the mRNA molecules. The oligo(A) tails present on inactive mRNAs in oocytes or embryos are also generated by shortening of full-length tails made in the cell nucleus (18).The poly(A) tail is added during 3′ end processing of mRNA precursors in the cell nucleus (1921). This reaction consists of two steps: an endonucleolytic cleavage followed by the addition of the poly(A) tail to the upstream cleavage product. Whereas a large protein machinery of some 20 or more polypeptides (22) is required for the cleavage reaction, subsequent polyadenylation has much simpler protein requirements. In the mammalian system, it can be reconstituted from three proteins: poly(A) polymerase, the enzyme catalyzing primer-dependent polymerization of AMP using ATP as a precursor (2325); the cleavage and polyadenylation specificity factor (CPSF),6 which binds the cleavage and polyadenylation signal AAUAAA (26, 27); and the nuclear poly(A)-binding protein (PABPN1), which binds the growing poly(A) tail (28, 29). Note that PABPN1 is distinct from the family of cytoplasmic poly(A)-binding proteins (30). Roles of poly(A) polymerase and CPSF in polyadenylation in vivo have been most clearly demonstrated by genetic analysis of the orthologues in S. cerevisiae (21, 31). PABPN1 has no functional orthologue in budding yeast (32); its function in polyadenylation has been confirmed in mammalian cells (33) and in Drosophila (34).Whereas PABPN1 and poly(A) polymerase are monomeric proteins, CPSF is a hetero-oligomer, which has not yet been reconstituted from recombinant proteins (22, 26, 3540). Poly(A) polymerase on its own is barely active because of a low affinity for its RNA substrate and thus acts distributively, i.e. it dissociates from the RNA after each polymerization step, and presumably often before it has incorporated any nucleotide; the enzyme also has no significant sequence specificity and will elongate any RNA with a free 3′ OH (24). Both CPSF and PABPN1 enhance the activity of the polymerase by recruiting the enzyme to its substrate through direct interactions (38, 41). Sequence specificity of poly(A) addition reflects the RNA binding specificities of the two stimulatory factors: CPSF recruits the polymerase to RNAs containing the AAUAAA sequence in the vicinity of their 3′ ends (24, 42, 43), and PABPN1 recruits the enzyme to substrate RNAs carrying a terminal oligo(A) tract (29). Each factor alone endows the polymerase with modest processivity, such that it can incorporate maybe two to five nucleotides before dissociating (44). RNAs containing both the AAUAAA sequence and an oligo(A) tail and thus resembling intermediates of the polyadenylation reaction support a cooperative or synergistic stimulation of poly(A) polymerase by both CPSF and PABPN1. Under these conditions, addition of the poly(A) tail occurs in a processive manner, i.e. without intermittent dissociation of the protein complex from its substrate RNA (29, 44).Interestingly, the reconstituted polyadenylation reaction also shows proper length control, generating poly(A) tails of the same length as seen in vivo; tails grow to a relatively well defined length of 250–300 nucleotides in a rapid, processive reaction (29, 44). Length control is due to termination of this processive elongation; extension beyond 250 A residues is largely distributive and therefore slow (45). These kinetics of in vitro poly(A) tail synthesis are fully consistent with the in vivo kinetics derived from pulse-labeling studies (46). In vitro, poly(A) tail elongation rates beyond 250 A residues are similar when either CPSF or PABPN1 or both are present. In other words, substrates with long poly(A) tails no longer support the cooperative stimulation of poly(A) polymerase by both CPSF and PABPN1 that is the basis of processive elongation (45). The termination of processive elongation must be mediated by a change in the RNA-protein complex that remains to be defined. When RNAs carrying poly(A) tails of different lengths are used as substrates for polyadenylation, the tails are always elongated processively to 250 nucleotides, independently of the initial length, whereas extension of a tail of 250 or more nucleotides in length is slow and distributive from the start of the reaction. Thus, poly(A) tail length control is based on some kind of AMP residue counting or length measurement, not on a kinetic mechanism (45).In this paper, we address the two problems outlined above: first, how does the polyadenylation complex change to terminate processive poly(A) tail elongation, and second, how is the length of the tail measured? We provide evidence that PABPN1 is the active component in the mechanism of length control. The protein promotes the interaction between CPSF and poly(A) polymerase when bound to a short poly(A) tail. PABPN1 no longer promotes or even actively disrupts this interaction when bound to a poly(A) tail of 250 nucleotides or longer and thereby terminates the cooperative, processive elongation reaction in a poly(A) tail length-dependent manner. Only poly(A) sequences are counted as part of the tail. Because this reflects the binding specificity of PABPN1 and because disruption of the CPSF-poly(A) polymerase interaction requires complete coverage of the poly(A) tail by this protein, PABPN1 is also the protein that measures the length of the tail.  相似文献   

12.
Regulation of poly(A) tail length during mRNA 3'-end formation requires a specific poly(A)-binding protein in addition to the cleavage/polyadenylation machinery. The mechanism that controls polyadenylation in mammals is well understood and involves the nuclear poly(A)-binding protein PABPN1. In contrast, poly(A) tail length regulation is poorly understood in yeast. Previous studies have suggested that the major cytoplasmic poly(A)-binding protein Pab1p acts as a length control factor in conjunction with the Pab1p-dependent poly(A) nuclease PAN, to regulate poly(A) tail length in an mRNA specific manner. In contrast, we recently showed that Nab2p regulates polyadenylation during de novo synthesis, and its nuclear location is more consistent with a role in 3'-end processing than that of cytoplasmic Pab1p. Here, we investigate whether PAN activity is required for de novo poly(A) tail synthesis. Components required for mRNA 3'-end formation were purified from wild-type and pan mutant cells. In both situations, 3'-end formation could be reconstituted whether Nab2p or Pab1p was used as the poly(A) tail length control factor. However, polyadenylation was more efficient and physiologically more relevant in the presence of Nab2p as opposed to Pab1p. Moreover, cell immunofluorescence studies confirmed that PAN subunits are localized in the cytoplasm which suggests that cytoplasmic Pab1p and PAN may act at a later stage in mRNA metabolism. Based on these findings, we propose that Nab2p is necessary and sufficient to regulate poly(A) tail length during de novo synthesis in yeast.  相似文献   

13.
14.
15.
Translational activation in oocytes and embryos is often regulated via increases in poly(A) length. Cleavage and polyadenylation specificity factor (CPSF), cytoplasmic polyadenylation element binding protein (CPEB), and poly(A) polymerase (PAP) have each been implicated in cytoplasmic polyadenylation in Xenopus laevis oocytes. Cytoplasmic polyadenylation activity first appears in vertebrate oocytes during meiotic maturation. Data presented here shows that complexes containing both CPSF and CPEB are present in extracts of X. laevis oocytes prepared before or after meiotic maturation. Assessment of a variety of RNA sequences as polyadenylation substrates indicates that the sequence specificity of polyadenylation in egg extracts is comparable to that observed with highly purified mammalian CPSF and recombinant PAP. The two in vitro systems exhibit a sequence specificity that is similar, but not identical, to that observed in vivo, as assessed by injection of the same RNAs into the oocyte. These findings imply that CPSFs intrinsic RNA sequence preferences are sufficient to account for the specificity of cytoplasmic polyadenylation of some mRNAs. We discuss the hypothesis that CPSF is required for all polyadenylation reactions, but that the polyadenylation of some mRNAs may require additional factors such as CPEB. To test the consequences of PAP binding to mRNAs in vivo, PAP was tethered to a reporter mRNA in resting oocytes using MS2 coat protein. Tethered PAP catalyzed polyadenylation and stimulated translation approximately 40-fold; stimulation was exclusively cis-acting, but was independent of a CPE and AAUAAA. Both polyadenylation and translational stimulation required PAPs catalytic core, but did not require the putative CPSF interaction domain of PAP. These results demonstrate that premature recruitment of PAP can cause precocious polyadenylation and translational stimulation in the resting oocyte, and can be interpreted to suggest that the role of other factors is to deliver PAP to the mRNA.  相似文献   

16.
17.
Human ribosomal protein S13 inhibits splicing of the own pre-mRNA   总被引:1,自引:0,他引:1  
Recombinant human ribosomal protein S13 (rpS 13) is shown to bind specifically a fragment of its own pre-mRNA that includes exons 1 and 2, intron 1, and part of intron 2, and to inhibit the splicing of that fragment in vitro. The weaker binding of other recombinant human ribosomal proteins, S10 and S16, to this pre-mRNA fragment indicated that the binding of rpS 13 was specific. Besides, poly(AU) and adenovirus pre-mRNA fragment affected poorly the binding of rpS 13 to S13 pre-mRNA, providing another evidence that the interaction was specific. RpS 13 specifically inhibited the pre-mRNA splicing whereas recombinant rpS10 and rpS16 did not affect excision of intron from S13 pre-mRNA fragment in contrast to rpS 13. Those positions in S13 pre-mRNA that were protected by rpS13 protein against cleavage by RNases T1, T2 and V1 were found to be located closely to the 5' and 3' splice sites in the pre-mRNA. Intron 1 in S13 pre-mRNA is more highly conserved within mammals than the other introns in S13 pre-mRNA, which supports the possibility of an important role for intron 1 in the regulation of expression of rpS13 gene in mammals.  相似文献   

18.
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   

19.
During the early period of infection, the 4.9 kb (kb = 10(3) bases) E2A premRNA of adenovirus serotype 2 is matured mainly into a 2.0 kb mRNA by excision of introns of 2233 and 626 nucleotides. In order to define all the possible steps of splicing occurring in vivo, we characterized splicing intermediates present after a limiting treatment of cells with cycloheximide. Three complementary methods of analysis were used: RNA transfer analysis, S1 nuclease mapping and complementary DNA-RNase assay. Our principal conclusions concerning the poly(A)+ species are as follows. The RNA intermediate family detected is more complex than expected, since two major RNA intermediates of 4.6 kb and 4.3 kb, two minor intermediates of 2.9 kb and 2.6 kb, and a 2.3 kb RNA, which represents a minor alternative mRNA form, are revealed. Despite its large size and the presence of multiple internal donor and acceptor signals, intron 1 is exclusively excised as a whole. Intron 2 is either primarily excised as a whole, removing the standard 626-nucleotide sequence, or a smaller sequence of 337 nucleotides is removed, generating the 2.3 kb alternative mRNA. Kinetics of the ligation reaction demonstrate that the minimal time for excision of intron 2 is no more than two minutes, indicating a high level of co-ordination of the multiple individual reactions occurring during excision of an intron. Besides the major pathway for E2A premRNA splicing, namely the excision first of intron 2, followed by the excision of intron 1 after a lag time of five minutes, a minor pathway (used with a frequency of 10%) can be detected where the order of intron excision was inverted. With the alternative variant of excision of intron 2, at least three different pathways are therefore used to mature the E2A premRNA. RNA intermediates resulting from the cleavage at the 5' end of introns and branching can be detected by S1 mapping experiments, but their low accumulative level (1% relatively to the initial premRNA) precluded their direction by RNA transfer experiments and their complete characterization.  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