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 5′ cap and 3′ poly(A) tail of eukaryotic mRNAs cooperate to synergistically stimulate translation initiation in vivo. We recently described mammalian cytoplasmic extracts which, following ultracentrifugation to partially deplete them of ribosomes and associated initiation factors, reproduce cap–poly(A) synergy in vitro. Using these systems, we demonstrate that synergy requires interaction between the poly(A)-binding protein (PABP) and the eukaryotic initiation factor (eIF) 4F holoenzyme complex, which recognises the 5′ cap. Here we further characterise the requirements and constraints of cap–poly(A) synergy in reticulocyte lysates by evaluating the effects of different parameters on synergy. The extent of extract depletion and the amounts of different initiation factors in depleted extracts were examined, as well as the effects of varying the concentrations of KCl, MgCl2 and programming mRNA and of adding a cap analogue. The results presented demonstrate that maximal cap–poly(A) synergy requires: (i) limiting concentrations of ribosome-associated initiation factors; (ii) precise ratios of mRNA to translation machinery (low concentrations of ribosome-associated initiation factors and low, non-saturating mRNA concentrations); (iii) physiological concentrations of added KCl and MgCl2. Additionally, we show that the eIF4G–PABP interaction on mRNAs which are capped and polyadenylated significantly increases the affinity of eIF4E for the 5′ cap.  相似文献   

2.
Using sedimentation and cryo electron tomography techniques, the conformations of eukaryotic polyribosomes formed in a long-term cell-free translation system were analyzed over all the active system lifetime (20–30 translation rounds during 6–8 h in wheat germ extract at 25°C). Three distinct types of the conformations were observed: (i) circular polyribosomes, varying from ring-shaped forms to circles collapsed into double rows, (ii) linear polyribosomes, tending to acquire planar zigzag-like forms and (iii) densely packed 3D helices. At the start, during the first two rounds of translation mostly the circular (ring-shaped and double-row) polyribosomes and the linear (free-shaped and zigzag-like) polyribosomes were formed (‘juvenile phase’). The progressive loading of the polyribosomes with translating ribosomes induced the opening of the circular polyribosomes and the transformation of a major part of the linear polyribosomes into the dense 3D helices (‘transitional phase’). After 2 h from the beginning (about 8–10 rounds of translation) this compact form of polyribosomes became predominant, whereas the circular and linear polyribosome fractions together contained less than half of polysomal ribosomes (‘steady-state phase’). The latter proportions did not change for several hours. Functional tests showed a reduced translational activity in the fraction of the 3D helical polyribosomes.  相似文献   

3.
Enhancement of eukaryotic messenger RNA (mRNA) translation initiation by the 3′ poly(A) tail is mediated through interaction of poly(A)-binding protein with eukaryotic initiation factor (eIF) 4G, bridging the 5′ terminal cap structure. In contrast to cellular mRNA, translation of the uncapped, non-polyadenylated hepatitis C virus (HCV) genome occurs independently of eIF4G and a role for 3′-untranslated sequences in modifying HCV gene expression is controversial. Utilizing cell-based and in vitro translation assays, we show that the HCV 3′-untranslated region (UTR) or a 3′ poly(A) tract of sufficient length interchangeably stimulate translation dependent upon the HCV internal ribosomal entry site (IRES). However, in contrast to cap-dependent translation, the rate of initiation at the HCV IRES was unaffected by 3′-untranslated sequences. Analysis of post-initiation events revealed that the 3′ poly(A) tract and HCV 3′-UTR improve translation efficiency by enabling termination and possibly ribosome recycling for successive rounds of translation.  相似文献   

4.
Translation of most eukaryotic mRNAs involves the synergistic action between the 5′ cap structure and the 3′ poly(A) tail at the initiation step. The poly(A) tail has also been shown to stimulate translation of picornavirus internal ribosome entry sites (IRES)-directed translation. These effects have been attributed principally to interactions between eIF4G and poly(A)-binding protein (PABP) but also to the participation of PABP in other steps during translation initiation. As the rabbit reticulocyte lysate (RRL) does not recapitulate this cap/poly(A) synergy, several systems based on cellular cell-free extracts have been developed to study the effects of poly(A) tail in vitro but they generally exhibit low translational efficiency. Here, we describe that the non-nuclease-treated RRL (untreated RRL) is able to recapitulate the effects of poly(A) tail on translation in vitro. In this system, translation of a capped/polyadenylated RNA was specifically inhibited by either Paip2 or poly(rA), whereas translation directed by HCV IRES remained unaffected. Moreover, cleavage of eIF4G by FMDV L protease strongly stimulated translation directed by the EMCV IRES, thus recapitulating the competitive advantage that the proteolytic processing of eIF4G confers to IRES-driven RNAs.  相似文献   

5.
The time course of polysome formation was studied in a long-term wheat germ cell-free translation system using sedimentation and electron microscopy techniques. The polysomes were formed on uncapped luciferase mRNA with translation-enhancing 5′ and 3′ UTRs. The formation of fully loaded polysomes was found to be a long process that required many rounds of translation and proceeded via several phases. First, short linear polysomes containing no more than six ribosomes were formed. Next, folding of these polysomes into short double-row clusters occurred. Subsequent gradual elongation of the clusters gave rise to heavy-loaded double-row strings containing up to 30–40 ribosomes. The formation of the double-row polysomes was considered to be equivalent to circularization of polysomes, with antiparallel halves of the circle being laterally stuck together by ribosome interactions. A slow exchange with free ribosomes and free mRNA observed in the double-row type polysomes, as well as the resistance of translation in them to AMP-PNP, provided evidence that most polysomal ribosomes reinitiate translation within the circularized polysomes without scanning of 5′ UTR, while de novo initiation including 5′ UTR scanning proceeds at a much slower rate. Removal or replacements of 5′ and 3′ UTRs affected the initial phase of translation, but did not prevent the formation of the double-row polysomes during translation.  相似文献   

6.
The 5′-leader sequence (called Ω) of tobacco mosaic virus (TMV) functions as a translational enhancer in plants. A poly(CAA) region within Ω is responsible for the translation enhancement and serves as a binding site for the heat shock protein, HSP101, which is required for the translational enhancement. Genetic analysis of the HSP101-mediated enhancement of translation from Ω-containing mRNA suggested that two eukaryotic initiation factors (eIFs), i.e. eIF4G and eIF3, were necessary. In this study, the functional interaction between Ω and other RNA elements known to participate in the recruitment of eIF4G, i.e. the 5′-cap and the poly(A) tail, was examined. Ω exhibited functional overlap with the 5′-cap and the poly(A) tail but not with the native TMV 3′-UTR which contains an independent translational enhancer. Consistent with the role of HSP101 in mediating the translational function of Ω, the enhancement afforded by Ω increased following a heat shock, which elevates expression of HSP101. The use of a fractionated translation lysate revealed that of the two eIF4F proteins present in plants, eIF4F was specifically required for the activity of Ω. The data suggest that Ω is functionally similar to a 5′-cap and a poly(A) tail in that it serves to recruit eIF4F in order to enhance translation from an mRNA.  相似文献   

7.
Poly(A)-binding protein (PABP) is a translation initiation factor that interacts with the poly(A) tail of mRNAs. PABP bound to poly(A) stimulates translation by interacting with the eukaryotic initiation factor 4G (eIF4G), which brings the 3′ end of an mRNA close to its 5′ m7G cap structure through consecutive interactions of the 3′-poly(A)–PABP-eIF4G-eIF4E-5′ m7G cap. PABP is a highly abundant translation factor present in considerably larger quantities than mRNA and eIF4G in cells. However, it has not been elucidated how eIF4G, present in limited cellular concentrations, is not sequestered by mRNA-free PABP, present at high cellular concentrations, but associates with PABP complexed with the poly(A) tail of an mRNA. Here, we report that RNA-free PABPs dimerize with a head-to-head type configuration of PABP, which interferes in the interaction between PABP and eIF4G. We identified the domains of PABP responsible for PABP–PABP interaction. Poly(A) RNA was shown to convert the PABP–PABP complex into a poly(A)–PABP complex, with a head-to-tail-type configuration of PABP that facilitates the interaction between PABP and eIF4G. Lastly, we showed that the transition from the PABP dimer to the poly(A)–PABP complex is necessary for the translational activation function.  相似文献   

8.
Rabbit polyribosomal globin messenger RNA (mRNA) and messenger ribonucleoprotein (mRNP) were labelled at the 3′ poly(A) tail to high specific activity with T4 RNA ligase and [5′-32P]pCp without consequent loss of functional activity. Labelled message was translated in both micrococcal nuclease treated and untreated rabbit reticulocyte lysates, as shown by the formation of labelled polyribosomes. The utilisation of labelled messenger was abolished by T2 toxin or sodium fluoride which are known to inhibit protein synthesis.  相似文献   

9.
10.
The 5′-cap structure and poly(A) tail of eukaryotic mRNAs function synergistically to promote translation initiation through a physical interaction between the proteins that bind to these regulatory elements. In this study, we have examined the effect of leader length and the presence of secondary structure on the translational competence and the function of the cap and poly(A) tail for mRNAs microinjected into Xenopus oocytes. Increasing the length of the 5′-leader from 17 to 144 nt resulted in a 2- to 4-fold increase in expression from an mRNA containing an unstructured leader but increased expression up to 20-fold for an mRNA containing 5′-proximal structure. Consequently, the presence of secondary structure was less inhibitory for those mRNAs with a longer 5′-leader. Co-injection of poly(A)-binding protein (PABP) mRNA increased the function of the cap and poly(A) tail in promoting translation from poly(A)+ but not poly(A) mRNAs, particularly for mRNAs containing secondary structure. In the absence of an internal ribosome entry site, expression from the distal cistron of a dicistronic mRNA increased as a function of the length of the intercistronic region and the concentration of PABP. The inhibitory effect of intercistronic located secondary structure on translation was position-dependent. Indeed, the effect of secondary structure was abolished if positioned 134 nt upstream of the distal cistron. These data suggest that the length of a leader, the presence of secondary structure and the concentration of PABP determine the extent to which the cap and poly(A) tail regulate translation.  相似文献   

11.
12.
Alphaviruses are mosquito-transmitted RNA viruses that cause important diseases in both humans and livestock. Sindbis virus (SIN), the type species of the alphavirus genus, carries a 11.7-kb positive-sense RNA genome which is capped at its 5′ end and polyadenylated at its 3′ end. The 3′ nontranslated region (3′NTR) of the SIN genome carries many AU-rich motifs, including a 19-nucleotide (nt) conserved element (3′CSE) and a poly(A) tail. This 3′CSE and the adjoining poly(A) tail are believed to regulate the synthesis of negative-sense RNA and genome replication in vivo. We have recently demonstrated that the SIN genome lacking the poly(A) tail was infectious and that de novo polyadenylation could occur in vivo (K. R. Hill, M. Hajjou, J. Hu, and R. Raju, J. Virol. 71:2693–2704, 1997). Here, we demonstrate that the 3′-terminal 29-nt region of the SIN genome carries a signal for possible cytoplasmic polyadenylation. To further investigate the polyadenylation signals within the 3′NTR, we generated a battery of mutant genomes with mutations in the 3′NTR and tested their ability to generate infectious virus and undergo 3′ polyadenylation in vivo. Engineered SIN genomes with terminal deletions within the 19-nt 3′CSE were infectious and regained their poly(A) tail. Also, a SIN genome carrying the poly(A) tail but lacking a part or the entire 19-nt 3′CSE was also infectious. Sequence analysis of viruses generated from these engineered SIN genomes demonstrated the addition of a variety of AU-rich sequence motifs just adjacent to the poly(A) tail. The addition of AU-rich motifs to the mutant SIN genomes appears to require the presence of a significant portion of the 3′NTR. These results indicate the ability of alphavirus RNAs to undergo 3′ repair and the existence of a pathway for the addition of AU-rich sequences and a poly(A) tail to their 3′ end in the infected host cell. Most importantly, these results indicate the ability of alphavirus replication machinery to use a multitude of AU-rich RNA sequences abutted by a poly(A) motif as promoters for negative-sense RNA synthesis and genome replication in vivo. The possible roles of cytoplasmic polyadenylation machinery, terminal transferase-like enzymes, and the viral polymerase in the terminal repair processes are discussed.  相似文献   

13.
The poly(A) tail is a crucial determinant in the control of both mRNA translation and decay. Poly(A) tail length dictates the triggering of the degradation of the message body in the major 5′ to 3′ and 3′ to 5′ mRNA decay pathways of eukaryotes. In the 5′ to 3′ pathway oligoadenylated but not polyadenylated mRNAs are selectively decapped in vivo, allowing their subsequent degradation by 5′ to 3′ exonucleolysis. The conserved Lsm1p-7p-Pat1p complex is required for normal rates of decapping in vivo, and the purified complex exhibits strong binding preference for oligoadenylated RNAs over polyadenylated or unadenylated RNAs in vitro. In the present study, we show that two lsm1 mutants produce mutant complexes that fail to exhibit such higher affinity for oligoadenylated RNA in vitro. Interestingly, these mutant complexes are normal with regard to their integrity and retain the characteristic RNA binding properties of the wild-type complex, namely, binding near the 3′-end of the RNA, having higher affinity for unadenylated RNAs that carry U-tracts near the 3′-end over those that do not and exhibiting similar affinities for unadenylated and polyadenylated RNAs. Yet, these lsm1 mutants exhibit a strong mRNA decay defect in vivo. These results underscore the importance of Lsm1p-7p-Pat1p complex–mRNA interaction for mRNA decay in vivo and imply that the oligo(A) tail mediated enhancement of such interaction is crucial in that process.  相似文献   

14.
Animal microRNAs (miRNAs) typically regulate gene expression by binding to partially complementary target sites in the 3′ untranslated region (UTR) of messenger RNA (mRNA) reducing its translation and stability. They also commonly induce shortening of the mRNA 3′ poly(A) tail, which contributes to their mRNA decay promoting function. The relationship between miRNA-mediated deadenylation and translational repression has been less clear. Using transfection of reporter constructs carrying three imperfectly matching let-7 target sites in the 3′ UTR into mammalian cells we observe rapid target mRNA deadenylation that precedes measureable translational repression by endogenous let-7 miRNA. Depleting cells of the argonaute co-factors RCK or TNRC6A can impair let-7-mediated repression despite ongoing mRNA deadenylation, indicating that deadenylation alone is not sufficient to effect full repression. Nevertheless, the magnitude of translational repression by let-7 is diminished when the target reporter lacks a poly(A) tail. Employing an antisense strategy to block deadenylation of target mRNA with poly(A) tail also partially impairs translational repression. On the one hand, these experiments confirm that tail removal by deadenylation is not strictly required for translational repression. On the other hand they show directly that deadenylation can augment miRNA-mediated translational repression in mammalian cells beyond stimulating mRNA decay. Taken together with published work, these results suggest a dual role of deadenylation in miRNA function: it contributes to translational repression as well as mRNA decay and is thus critically involved in establishing the quantitatively appropriate physiological response to miRNAs.  相似文献   

15.
Barley yellow dwarf virus mRNA, which lacks both cap and poly(A) tail, has a translation element (3′-BTE) in its 3′-UTR essential for efficient translation initiation at the 5′-proximal AUG. This mechanism requires eukaryotic initiation factor 4G (eIF4G), subunit of heterodimer eIF4F (plant eIF4F lacks eIF4A), and 3′-BTE-5′-UTR interaction. Using fluorescence anisotropy, SHAPE (selective 2′-hydroxyl acylation analyzed by primer extension) analysis, and toeprinting, we found that (i) 40S subunits bind to BTE (Kd = 350 ± 30 nm), (ii) the helicase complex eIF4F-eIF4A-eIF4B-ATP increases 40S subunit binding (Kd = 120 ± 10 nm) to the conserved stem-loop I of the 3′-BTE by exposing more unpaired bases, and (iii) long distance base pairing transfers this complex to the 5′-end of the mRNA, where translation initiates. Although 3′-5′ interactions have been recognized as important in mRNA translation, barley yellow dwarf virus employs a novel mechanism utilizing the 3′-UTR as the primary site of ribosome recruitment.  相似文献   

16.
Ribosome profiling identifies ribosome positions on translated mRNAs. A prominent feature of published datasets is the near complete absence of ribosomes in 3′ untranslated regions (3′UTR) although substantial ribosome density can be observed on non-coding RNAs. Here we perform ribosome profiling in cultured Drosophila and human cells and show that different features of translation are revealed depending on the nuclease and the digestion conditions used. Most importantly, we observe high abundance of ribosome protected fragments in 3′UTRs of thousands of genes without manipulation of translation termination. Affinity purification of ribosomes indicates that the 3′UTR reads originate from ribosome protected fragments. Association of ribosomes with the 3′UTR may be due to ribosome migration through the stop codon or 3′UTR mRNA binding to ribosomes on the coding sequence. This association depends primarily on the relative length of the 3′UTR and may be related to translational regulation or ribosome recycling, for which the efficiency is known to inversely correlate with 3′UTR length. Together our results indicate that ribosome profiling is highly dependent on digestion conditions and that ribosomes commonly associate with the 3′UTR, which may have a role in translational regulation.  相似文献   

17.
The interferon-inducible protein with tetratricopeptide (IFIT) family proteins inhibit replication of some viruses by recognizing several types of RNAs, including 5′-triphosphate RNA and 5′ capped 2′-O unmethylated mRNA. However, it remains unclear how IFITs inhibit replication of some viruses through recognition of RNA. Here, we analyzed the mechanisms by which Ifit1 exerts antiviral responses. Replication of a Japanese encephalitis virus (JEV) 2′-O methyltransferase (MTase) mutant was markedly enhanced in mouse embryonic fibroblasts and macrophages lacking Ifit1. Ifit1 bound 5′-triphosphate RNA but more preferentially associated with 5′ capped 2′-O unmethylated mRNA. Ifit1 inhibited the translation of mRNA and thereby restricted the replication of JEV mutated in 2′-O MTase. Thus, Ifit1 inhibits replication of MTase-defective JEV by inhibiting mRNA translation through direct binding to mRNA 5′ structures.  相似文献   

18.
The eukaryotic mRNA 3′ poly(A) tail and its associated poly(A)-binding protein (Pab1p) are important regulators of gene expression. One role for this complex in the yeast Saccharomyces cerevisiae is in translation initiation through an interaction with a 115-amino-acid region of the translation initiation factor eIF4G. The eIF4G-interacting domain of Pab1p was mapped to its second RNA recognition motif (RRM2) in an in vitro binding assay. Moreover, RRM2 of Pab1p was required for poly(A) tail-dependent translation in yeast extracts. An analysis of a site-directed Pab1p mutation which bound to eIF4G but did not stimulate translation of uncapped, polyadenylated mRNA suggested additional Pab1p-dependent events during translation initiation. These results support the model that the association of RRM2 of yeast Pab1p with eIF4G is a prerequisite for the poly(A) tail to stimulate the translation of mRNA in vitro.  相似文献   

19.
MicroRNA (miRNA)-induced silencing complexes (miRISCs) repress translation and promote degradation of miRNA targets. Target degradation occurs through the 5′-to-3′ messenger RNA (mRNA) decay pathway, wherein, after shortening of the mRNA poly(A) tail, the removal of the 5′ cap structure by decapping triggers irreversible decay of the mRNA body. Here, we demonstrate that miRISC enhances the association of the decapping activators DCP1, Me31B and HPat with deadenylated miRNA targets that accumulate when decapping is blocked. DCP1 and Me31B recruitment by miRISC occurs before the completion of deadenylation. Remarkably, miRISC recruits DCP1, Me31B and HPat to engineered miRNA targets transcribed by RNA polymerase III, which lack a cap structure, a protein-coding region and a poly(A) tail. Furthermore, miRISC can trigger decapping and the subsequent degradation of mRNA targets independently of ongoing deadenylation. Thus, miRISC increases the local concentration of the decapping machinery on miRNA targets to facilitate decapping and irreversibly shut down their translation.  相似文献   

20.
The translation of specific maternal mRNAs is regulated during early development. For some mRNAs, an increase in translational activity is correlated with cytoplasmic extension of their poly(A) tails; for others, translational inactivation is correlated with removal of their poly(A) tails. Recent results in several systems suggest that events at the 3′ end of the mRNA can affect the state of the 5′ cap structure, m7G(5′)ppp(5′)G. We focus here on the potential role of cap modifications on translation during early development and on the question of whether any such modifications are dependent on cytoplasmic poly(A) addition or removal. To do so, we injected synthetic RNAs into Xenopus oocytes and examined their cap structures and translational activities during meiotic maturation. We draw four main conclusions. First, the activity of a cytoplasmic guanine-7-methyltransferase increases during oocyte maturation and stimulates translation of an injected mRNA bearing a nonmethylated GpppG cap. The importance of the cap for translation in oocytes is corroborated by the sensitivity of protein synthesis to cap analogs and by the inefficient translation of mRNAs bearing nonphysiologically capped 5′ termini. Second, deadenylation during oocyte maturation does not cause decapping, in contrast to deadenylation-triggered decapping in Saccharomyces cerevisiae. Third, the poly(A) tail and the N-7 methyl group of the cap stimulate translation synergistically during oocyte maturation. Fourth, cap ribose methylation of certain mRNAs is very inefficient and is not required for their translational recruitment by poly(A). These results demonstrate that polyadenylation can cause translational recruitment independent of ribose methylation. We propose that polyadenylation enhances translation through at least two mechanisms that are distinguished by their dependence on ribose modification.  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