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 46 毫秒
1.
Infection of cells by herpes simplex virus type 1 (HSV-1) triggers host cell shutoff whereby mRNAs are degraded and cellular protein synthesis is diminished. However, virus protein translation continues because the translational apparatus in HSV-infected cells is maintained in an active state. Surprisingly, poly(A)-binding protein 1 (PABP1), a predominantly cytoplasmic protein that is required for efficient translation initiation, is partially relocated to the nucleus during HSV-1 infection. This relocalization occurred in a time-dependent manner with respect to virus infection. Since HSV-1 infection causes cell stress, we examined other cell stress inducers and found that oxidative stress similarly relocated PABP1. An examination of stress-induced kinases revealed similarities in HSV-1 infection and oxidative stress activation of JNK and p38 mitogen-activated protein (MAP) kinases. Importantly, PABP relocalization in infection was found to be independent of the viral protein ICP27. The depletion of PABP1 by small interfering RNA (siRNA) knockdown had no significant effect on viral replication or the expression of selected virus late proteins, suggesting that reduced levels of cytoplasmic PABP1 are tolerated during infection.The lytic replication cycle of herpes simplex virus type 1 (HSV-1) can be divided into three phases, immediate-early (IE), early (E), and late (L), that occur in a coordinated sequential gene expression program. IE proteins can regulate E and L gene expression, which produces proteins involved in DNA replication, capsid production, and virion assembly. HSV infection results in host cell shutoff to facilitate the efficient production of viral proteins. First, mRNA is degraded by the virion-associated vhs protein, and then ICP27, a multifunctional regulator of gene expression, inhibits pre-mRNA splicing. As most viral mRNAs are intronless, this abrogates the production of stable cellular mRNAs that can be exported to the cytoplasm and compete for translation with viral mRNAs (44).HSV mRNAs are capped and polyadenylated and so are translated via a normal cap-dependent mechanism. Translation initiation, during which translationally active ribosomes are assembled, is a tightly regulated process (21). Eukaryotic initiation factor 4F (eIF4F) (composed of eIF4E, eIF4G, and eIF4A) that binds the cap at the 5′ end of the mRNA promotes the recruitment of the 40S ribosomal subunit and associated factors, including eIF2-GTP initiator tRNA. The recognition of the start codon then promotes large ribosomal subunit joining. Poly(A)-binding protein 1 (PABP1), which binds and multimerizes on mRNA poly(A) tails, enhances translation initiation through interactions with the eIF4G component of the eIF4F cap-binding complex (20, 29, 32, 51) to circularize the mRNA in a “closed-loop” conformation (24). Key protein-RNA and protein-protein interactions in the translation initiation complex are strengthened by this PABP1-mediated circularization (12).HSV-1 maintains active viral translation in the face of host translational shutoff. Infection activates protein kinase R (PKR), which phosphorylates eIF2α, resulting in translation inhibition. However, HSV-1 ICP34.5 redirects protein phosphatase 1α to reverse eIF2α phosphorylation, abrogating the block to translation (17, 38). In addition, the HSV-1 US11 protein inhibits PKR and may also block PKR-mediated eIF2α phosphorylation (40, 42). HSV-1 infection also enhances eIF4F assembly in quiescent cells by the phosphorylation and proteasome-mediated degradation of the eIF4E-binding protein (4E-BP), which, when hypophosphorylated, can negatively regulate eIF4F complex formation (54). However, ICP6 may also contribute to eIF4F assembly by binding to eIF4G (55). Finally, ICP6 is required for Mnk-1 phosphorylation of eIF4E, but the mechanisms behind this remain unclear (54). ICP27 has also been implicated in translation regulation during HSV infection (6, 8, 10, 30) and may also activate p38 mitogen-activated protein (MAP) kinase that can phosphorylate eIF4E (16, 59).PABP1 appears to be a common cellular target of RNA and DNA viruses. PABP1 can undergo proteolysis, intracellular relocalization, or modification of its interaction with other translation factors in response to infection. For example, poliovirus induces host cell shutoff by cleaving PABP1, thus disrupting certain PABP1-containing complexes (28, 29). The rotavirus NSP3 protein can displace PABP1 from translation initiation complexes (41). However, NSP3 also interacts with a cellular protein, RoXaN, which is required to relocate PABP1 to the nucleus (13). Similarly, the Kaposi''s sarcoma herpesvirus (KSHV) SOX protein plays a role in relocating PABP1, its cofactor in cellular mRNA decay, to the nucleus (33). Although steady-state levels of PABP1 are highest in the cytoplasm of normal cells, where it has cytoplasmic functions, it is a nucleocytoplasmic shuttling protein (1). However, it is unclear how PABP1 enters or exits the nucleus, as it contains neither a canonical nuclear export nor an import signal.Here we describe the loss of PABP1 from cap-binding complexes and the partial relocation of PABP1 to the nucleus in HSV-1-infected cells in a time-dependent manner. Relocation is specific for PABP1, as other translation factors remained in the cytoplasm. Cells undergo stress during HSV-1 infection, and analysis of a variety of cell stresses revealed that PABP relocalization was also observed upon oxidative stress. Paxillin, a potential PABP1 nuclear chaperone, was phosphorylated, and the paxillin-PABP1 interaction was reduced during virus infection. However, the interaction was weak and cell type dependent, indicating that other effectors of PABP1 relocation in the infected cell must exist. Recently, the HSV-1 ICP27 protein was suggested to alter the PABP1 cellular location (6). However, infections with ICP27-null mutant viruses clearly demonstrated that ICP27 is not required for PABP1 nuclear relocation in the context of infection. Although HSV-1 mRNAs are translated by a normal cap-dependent mechanism known to be enhanced by PABP1, small interfering RNA (siRNA) knockdown of PABP1 indicated that at late times of infection, the translation of certain virus late proteins tolerates very low levels of PABP1.  相似文献   

2.
GW182 family proteins are essential in animal cells for microRNA (miRNA)-mediated gene silencing, yet the molecular mechanism that allows GW182 to promote translational repression and mRNA decay remains largely unknown. Previous studies showed that while the GW182 N-terminal domain interacts with Argonaute proteins, translational repression and degradation of miRNA targets are promoted by a bipartite silencing domain comprising the GW182 middle and C-terminal regions. Here we show that the GW182 C-terminal region is required for GW182 to release silenced mRNPs; moreover, GW182 dissociates from miRNA targets at a step of silencing downstream of deadenylation, indicating that GW182 is required to initiate but not to maintain silencing. In addition, we show that the GW182 bipartite silencing domain competes with eukaryotic initiation factor 4G for binding to PABPC1. The GW182-PABPC1 interaction is also required for miRNA target degradation; accordingly, we observed that PABPC1 associates with components of the CCR4-NOT deadenylase complex. Finally, we show that PABPC1 overexpression suppresses the silencing of miRNA targets. We propose a model in which the GW182 silencing domain promotes translational repression, at least in part, by interfering with mRNA circularization and also recruits the deadenylase complex through the interaction with PABPC1.In multicellular eukaryotes, the regulation of gene expression by microRNAs (miRNAs) is critical for biological processes as diverse as cell differentiation and proliferation, apoptosis, metabolism, and development (4). To exert a regulatory function, miRNAs associate with Argonaute proteins to form RNA-induced silencing complexes, which repress translation and trigger the degradation of target mRNAs (4, 10, 16). The extent to which translational repression and degradation contribute to silencing depends on the specific target-miRNA combination; some targets are regulated predominantly at the translational level, whereas others can be regulated mainly at the mRNA level (3). A large-scale proteomic analysis performed in parallel with measurements of mRNA levels showed that for the vast majority of miRNA targets, silencing correlates with changes at both the protein and mRNA levels (1, 27).In animal cells, the degradation of miRNA targets is initiated by deadenylation and decapping, which are followed by the exonucleolytic decay of the mRNA body (2, 3, 9, 11, 12, 17, 19, 24, 30, 31). miRNA-dependent mRNA degradation requires a variety of proteins: an Argonaute and a GW182 protein, the CCR4-NOT deadenylase complex, the decapping enzyme DCP2, and several decapping activators including DCP1, Ge-1, HPat, EDC3, and Me31B (also known as RCK/p54) (3, 6, 9, 12, 19). Several studies previously demonstrated that miRNAs trigger deadenylation and decapping even when the mRNA target is not translated (9, 12, 19, 24, 30, 31), indicating that mRNA decay is not merely a consequence of a primary effect of miRNAs on translation but rather is an independent mechanism by which miRNAs silence gene expression.Although how miRNAs trigger mRNA degradation is well established, the mechanisms driving the inhibition of translation are unclear. Multiple mechanisms have been proposed: the displacement of eukaryotic initiation factor 4E (eIF4E) from the mRNA cap structure, interference with the function of the eIF4F complex, a block of 60S ribosomal subunit joining, or an inhibition of translation elongation (4, 10, 16). Regardless of the precise mechanism, the translational repression of miRNA targets also requires GW182 family proteins (11, 13).GW182 proteins are essential components of the miRNA pathway in animal cells, as their depletion suppresses miRNA-mediated gene silencing (reviewed in references 8 and 13). Recent studies have revealed that the silencing activity of these proteins resides predominantly in a bipartite silencing domain containing the middle and C-terminal regions (14, 22, 33). The precise molecular function of the GW182 silencing domain is not fully understood, yet it is known that the domain is not required for GW182 proteins to interact with Argonaute proteins or to localize to P bodies (3, 14, 22). Furthermore, when the silencing domains of GW182 proteins are artificially tethered to mRNAs, their expression is silenced; therefore, tethering bypasses the requirement for Argonaute proteins and miRNAs (5, 22, 33). These observations suggest that the silencing domains of GW182 proteins exhibit intrinsic silencing activity and therefore likely play a role at the effector step of silencing (13, 14, 22, 33).Here we investigate what role the Drosophila melanogaster GW182 silencing domain plays in the miRNA pathway. Overall, our results reveal that the very C-terminal region of this domain is required for the release of GW182 from silenced mRNPs. Indeed, we unexpectedly found that we could detect D. melanogaster GW182 bound to miRNA targets only in cells depleted of components of the deadenylase complex. These results suggest that GW182 dissociates from Argonaute-1 (AGO1) and miRNA targets at a step of silencing downstream of deadenylation. In contrast, GW182 mutants lacking the C-terminal region remain stably bound to miRNA targets, even in wild-type cells, indicating that this region plays a role in the dissociation of GW182 from effector complexes. We further show that the bipartite silencing domain of GW182 interacts with PABPC1 and interferes with the binding of PABPC1 to eIF4G. The interaction of GW182 with PABPC1 is also required for the degradation of miRNA targets, most likely because the interaction facilitates the recruitment of the CCR4-NOT deadenylase complex. Accordingly, overexpressing PABPC1 suppresses the silencing of miRNA targets. Our findings uncover an unexpected role for PABPC1 in the miRNA pathway.  相似文献   

3.
4.
5.
6.
7.
8.
Translation initiation plays an important role in cell growth, proliferation, and survival. The translation initiation factor eIF4B (eukaryotic initiation factor 4B) stimulates the RNA helicase activity of eIF4A in unwinding secondary structures in the 5′ untranslated region (5′UTR) of the mRNA in vitro. Here, we studied the effects of eIF4B depletion in cells using RNA interference (RNAi). In agreement with the role of eIF4B in translation initiation, its depletion resulted in inhibition of this step. Selective reduction of translation was observed for mRNAs harboring strong to moderate secondary structures in their 5′UTRs. These mRNAs encode proteins, which function in cell proliferation (Cdc25C, c-myc, and ODC [ornithine decarboxylase]) and survival (Bcl-2 and XIAP [X-linked inhibitor of apoptosis]). Furthermore, eIF4B silencing led to decreased proliferation rates, promoted caspase-dependent apoptosis, and further sensitized cells to camptothecin-induced cell death. These results demonstrate that eIF4B is required for cell proliferation and survival by regulating the translation of proliferative and prosurvival mRNAs.Targeting the translation initiation pathway is emerging as a potential therapy for inhibiting cancer cell growth (35, 38). Ribosome recruitment to the 5′ ends of eukaryotic mRNAs proceeds via translation initiation mechanisms that are dependent either on the 5′ cap structure (m7GpppN, where N is any nucleotide) or an internal ribosome entry site (IRES). The majority of translation initiation events in eukaryotes are mediated through cap-dependent translation whereby the 40S ribosomal subunit is recruited to the vicinity of the mRNA 5′ cap structure by the eukaryotic initiation factor 4F (eIF4F) complex. eIF4F is comprised of eIF4E (the cap-binding subunit), eIF4A (an RNA helicase), and eIF4G (a large scaffolding protein for eIF4E, eIF4A, and other initiation factors). Once assembled at the 5′ cap, the 40S ribosomal subunit in association with several initiation factors scans the 5′ untranslated region (5′UTR) of the mRNA until it encounters a start codon in a favorable context, followed by polypeptide synthesis (37).Early in vitro studies have shown that the initiation factor eIF4B acts to potentiate ribosome recruitment to the mRNA (3, 45). eIF4B stimulates translation of both capped and uncapped mRNAs in vitro (1, 36). This function is exerted through stimulation of the helicase activity of eIF4A (43), possibly through direct interactions with eIF4A (44) or with mRNA, the ribosome-associated eIF3, and 18S rRNA (28, 29, 44). Thus, eIF4B is thought to form auxiliary bridges between the mRNA and the 40S ribosomal subunit. Toeprinting studies using mammalian eIF4B underscored its importance in the assembly of the 48S initiation complex, especially on mRNAs harboring secondary structures in the 5′UTRs (11).In vivo studies of eIF4B are limited. Ectopic expression of eIF4B in cultured Drosophila melanogaster cells and in developing eye imaginal discs stimulated cell proliferation (16). Enhanced cell proliferation is most likely mediated by increased translation of a subset of mRNAs, since knockdown of Drosophila eIF4B by RNA interference (RNAi) caused a modest reduction in global translation but compromised the survival of insect cells grown under low serum conditions (16). Studies of eIF4B in mammalian cells yielded contradictory results. Transient overexpression of eIF4B stimulated translation initiation in a phosphorylation-dependent manner in some cells (18, 49) while inhibiting translation in others (30, 31, 41). These differences might be attributed to disparate levels of eIF4B overexpression.To address the physiological role of eIF4B in mRNA translation in the cell, RNAi knockdown of eIF4B was used here. We demonstrate that eIF4B is required for optimal translation. Importantly, the translation of mRNAs bearing structured 5′UTRs, such as the cell cycle regulators Cdc25C, c-myc, and ODC (ornithine decarboxylase), and the antiapoptotic factors Bcl-2 and XIAP (X-linked inhibitor of apoptosis), was reduced as a result of eIF4B silencing by RNAi. Furthermore, eIF4B silencing promoted caspase-dependent apoptosis. Thus, we show that mammalian eIF4B is required for cell proliferation and survival, whereby it acts by regulating the translation of a functionally related subset of mRNAs.  相似文献   

9.
ICP27 is a multifunctional protein that is required for herpes simplex virus 1 mRNA export. ICP27 interacts with the mRNA export receptor TAP/NXF1 and binds RNA through an RGG box motif. Unlike other RGG box proteins, ICP27 does not bind G-quartet structures but instead binds GC-rich sequences that are flexible in structure. To determine the contribution of arginines within the RGG box, we performed in vitro binding assays with N-terminal proteins encoding amino acids 1 to 160 of wild-type ICP27 or arginine-to-lysine substitution mutants. The R138,148,150K triple mutant bound weakly to sequences that were bound by the wild-type protein and single and double mutants. Furthermore, during infection with the R138,148,150K mutant, poly(A)+ RNA and newly transcribed RNA accumulated in the nucleus, indicating that viral RNA export was impaired. To determine if structural changes had occurred, nuclear magnetic resonance (NMR) analysis was performed on N-terminal proteins consisting of amino acids 1 to 160 from wild-type ICP27 and the R138,148,150K mutant. This region of ICP27 was found to be highly flexible, and there were no apparent differences in the spectra seen with wild-type ICP27 and the R138,148,150K mutant. Furthermore, NMR analysis with the wild-type protein bound to GC-rich sequences did not show any discernible folding. We conclude that arginines at positions 138, 148, and 150 within the RGG box of ICP27 are required for binding to GC-rich sequences and that the N-terminal portion of ICP27 is highly flexible in structure, which may account for its preference for binding flexible sequences.The herpes simplex virus 1 (HSV-1) protein ICP27 is a multifunctional regulatory protein that is required for productive viral infection. ICP27 interacts with a number of cellular proteins, and it binds RNA (35). One of the functions that ICP27 performs is to escort viral mRNAs from the nucleus to the cytoplasm for translation (2, 3, 5, 10, 13, 21, 34). ICP27 binds viral RNAs (5, 34) and interacts directly with the cellular mRNA export receptor TAP/NXF1 (2, 21), which is required for the export of HSV-1 mRNAs (20, 21). ICP27 also interacts with the export adaptor proteins Aly/REF (2, 3, 23) and UAP56 (L. A. Johnson, H. Swesey, and R. M. Sandri-Goldin, unpublished results), which form part of the TREX complex that binds to the 5′ end of mRNA through an interaction with CBP80 (26, 32, 41). Aly/REF does not appear to bind viral RNA directly (3), and it is not essential for HSV-1 RNA export based upon small interfering RNA (siRNA) knockdown studies (20), but it contributes to the efficiency of viral RNA export (3, 23). ICP27 also interacts with the SR splicing proteins SRp20 and 9G8 (11, 36), which have been shown to shuttle between the nucleus and the cytoplasm (1). SRp20 and 9G8 have also been shown to facilitate the export of some cellular RNAs (16, 17, 27) by binding RNA and interacting with TAP/NXF1 (14, 16, 18). The knockdown of SRp20 or 9G8 adversely affects HSV-1 replication and specifically results in a nuclear accumulation of newly transcribed RNA during infection (11). Thus, these SR proteins also contribute to the efficiency of viral RNA export. However, the overexpression of SRp20 was unable to rescue the defect in RNA export during infection with an ICP27 mutant that cannot bind RNA (11), suggesting that ICP27 is the major HSV-1 RNA export protein that links viral RNA to TAP/NXF1.ICP27 was shown previously to bind RNA through an RGG box motif located at amino acids 138 to 152 within the 512-amino-acid protein (28, 34). Using electrophoretic mobility shift assays (EMSAs), we showed that the N-terminal portion of ICP27 from amino acids 1 to 160 bound specifically to viral oligonucleotides that are GC rich and that are flexible and relatively unstructured (5). Here we report the importance of three arginine residues within the RGG box for ICP27 binding to GC-rich sequences in vitro and for viral RNA export during infection. We also performed nuclear magnetic resonance (NMR) structural analysis of the N-terminal portion of ICP27 for both the wild-type protein and an ICP27 mutant in which three arginines were replaced with lysines. The NMR data showed that the N-terminal portion of ICP27 is relatively unstructured but compact, and NMR analysis in the presence of oligonucleotide substrates to which the N-terminal portion of ICP27 binds did not show any discernible alterations in this highly flexible structure, nor did the arginine-to-lysine substitutions.  相似文献   

10.
11.
The dicistrovirus is a positive-strand single-stranded RNA virus that possesses two internal ribosome entry sites (IRES) that direct translation of distinct open reading frames encoding the viral structural and nonstructural proteins. Through an unusual mechanism, the intergenic region (IGR) IRES responsible for viral structural protein expression mimics a tRNA to directly recruit the ribosome and set the ribosome into translational elongation. In this study, we explored the mechanism of host translational shutoff in Drosophila S2 cells infected by the dicistrovirus, cricket paralysis virus (CrPV). CrPV infection of S2 cells results in host translational shutoff concomitant with an increase in viral protein synthesis. CrPV infection resulted in the dissociation of eukaryotic translation initiation factor 4G (eIF4G) and eIF4E early in infection and the induction of deIF2α phosphorylation at 3 h postinfection, which lags after the initial inhibition of host translation. Forced dephosphorylation of deIF2α by overexpression of dGADD34, which activates protein phosphatase I, did not prevent translational shutoff nor alter virus production, demonstrating that deIF2α phosphorylation is dispensable for host translational shutoff. However, premature induction of deIF2α phosphorylation by thapsigargin treatment early in infection reduced viral protein synthesis and replication. Finally, translation mediated by the 5′ untranslated region (5′UTR) and the IGR IRES were resistant to impairment of eIF4F or eIF2 in translation extracts. These results support a model by which the alteration of the deIF4F complex contribute to the shutoff of host translation during CrPV infection, thereby promoting viral protein synthesis via the CrPV 5′UTR and IGR IRES.For productive viral protein expression, viruses have to compete for and hijack the host translational machinery (45). Some viruses such as poliovirus, vesicular stomatitis virus (VSV), and influenza virus selectively antagonize the translation apparatus to shut off host translation, resulting in the release of ribosomes from host mRNAs and the inhibition of antiviral responses. On the other hand, the host cell can counteract through antiviral mechanisms to shutdown viral translation. For instance, viral RNA replication intermediates can trigger PKR, leading to an inhibition of overall translation. To bypass the block in translation, viruses have evolved unique mechanisms to preferentially recruit the ribosome for viral protein synthesis. Thus, the control of the translational machinery during infection is a major focal point in the battle between the host and the virus and often, elucidation of these viral translational shutoff strategies reveals key targets of translational regulation.The majority of cellular mRNAs initiate translation through the recruitment of the cap-binding complex, eukaryotic translation initiation factor 4F (eIF4F), to the 5′ cap of the mRNA (56). eIF4F consists of the cap-binding protein eIF4E, the RNA helicase, eIF4A, and the adaptor protein eIF4G. eIF4G acts as a bridge to join eIF4E and the 40S subunit via eIF3. With the ternary eIF2-Met-tRNAi-GTP complex bound, the 40S subunit scans in a 5′-to-3′ direction until an AUG start codon is encountered. Here, eIF5 mediates GTP hydrolysis on the ternary complex, releasing the eIFs and subsequently leading to 60S subunit joining to assemble an elongation-competent 80S ribosome. The ternary eIF2-Met-tRNAi-GTP complex is reactivated for another round of translation by exchange of GDP for GTP, which is mediated by the guanine nucleotide exchange factor, eIF2B. The 3′ poly(A) tail of the mRNA also stimulates translational initiation by binding to the poly(A) binding protein (PABP), which in turn interacts with eIF4G at the 5′end, resulting in a circularized mRNA. PABP has been proposed to enhance eIF4E affinity for the 5′cap and promote 60S joining, indicating that PABP functions at multiple steps of translational initiation (33).A common tactic viruses use to inhibit host translation is to selectively target eIFs. One of the best studied is the cleavage of eIF4G by viral proteases during picornavirus infection. In humans, two isoforms, eIF4GI and eIF4GII, are cleaved early in poliovirus infection by the viral protease 2A, where cleavage of eIFGII correlates more precisely with host translation shutoff (20). Cleavage of eIF4G produces an amino-terminal fragment that binds to eIF4E and a C-terminal fragment that binds to eIF4A and eIF3 (26, 39, 42). PABP is also cleaved by the viral protease 3C during poliovirus infection, thus contributing to shutoff of both host and viral translation and thereby enabling the switch from viral translation to replication (3, 31, 38). Another major target is the availability of the cap-binding protein eIF4E, which is regulated by binding to the repressor protein 4E-BP (21, 41). 4E-BP and eIF4G compete for an overlapping site on eIF4E (42). In its hypophosphorylated state, 4E-BP binds to and sequesters eIF4E, preventing eIF4G recruitment. Dephosphorylation and activation of 4E-BP has been observed during poliovirus, encephalomyocarditis (EMCV), and VSV infections (7, 18).During virus infection, host antiviral responses are triggered that also inhibit translation to counteract viral protein synthesis. An integral antiviral response is phosphorylation at Ser51 of eIF2α, which reduces the pool of the ternary complex by blocking the eIF2B-dependent exchange of GDP to GTP. In mammals, four known eIF2α kinases exist including the endoplasmic reticulum (ER)-stress-inducible PERK, GCN2, which senses the accumulation of deacylated tRNAs during amino acid starvation conditions; the heme-regulated kinase HRI; and the interferon-inducible double-stranded RNA-binding PKR (64). In mammalian cells, PKR is activated by binding to double-stranded viral RNA replication intermediates, leading to eIF2α phosphorylation and inhibition of overall host and viral translation. PERK and GCN2 have also been shown to be activated during virus infections by VSV and members of the alphavirus family (2, 6, 43, 65, 79). Often, viruses rely on the ER for synthesis and proper folding of viral proteins. The large burden on the ER activates PERK to phosphorylate eIF2α, thereby inhibiting global protein synthesis to reduce the load on the ER (23). Some viruses such as HCV and herpes simplex viruses have adapted to responses that induce eIF2α phosphorylation by producing viral proteins that counteract PKR or modulate the ER stress response (27, 76). Thus, virus infection can trigger several eIF2α kinases that lead to translational shutoff to counteract viral protein synthesis.To circumvent these translation blocks, viruses such as poliovirus and hepatitis C virus utilize internal ribosome entry sites (IRES), which are RNA elements that directly recruit ribosomes in a cap-independent manner and require only a subset of canonical eIFs (15, 25). It is generally thought that IRES-containing viral mRNAs can be translated under conditions when specific eIFs are compromised during infection. Except for a few cases, the specific mechanisms and factors that lead to IRES stimulation is poorly understood. For example, poliovirus and the related EMCV possess an IRES that allows viral translation despite cleavage of eIF4G during infection or inhibiting eIF4E by 4E-BP binding. This type of IRES can still bind to the central domain of eIF4G and mediate 40S subunit recruitment (11, 37, 57).One of the most unique and simplest IRES is found within the intergenic region (IGR) of the Dicistroviridae family (for extensive reviews, see references 28, 36, and 49). Members of this family include the cricket paralysis virus (CrPV), drosophila C virus (DCV), taura syndrome virus, the Plautia stali intestine virus (PSIV), the Rhopalosiphum padi virus (RhPV), and several bee viruses such as the black queen cell virus and the Israeli acute paralysis virus, which has been recently linked to colony collapse disorder (10). The dicistroviruses encode a positive-strand 8- to 10-kb single-stranded RNA genome, which contains two main open reading frames, ORF1 and ORF2, encoding the nonstructural and structural proteins, respectively, separated by an IGR (see Fig. Fig.1A).1A). The 5′ end of the CrPV RNA is linked to the viral protein VpG and the 3′ end contains a poly(A) tail (16). Radiolabeling of intracellular RNA in infected cells reveals no subgenomic RNA species smaller than the full-length genomic RNA, and this has been supported by Northern blot analysis (16, 81). Translation of ORF2 is directed by the IGR IRES, whereas ORF1 expression is mediated by an IRES within the 5′ untranslated region (5′UTR) (35, 67, 81, 82). Remarkably, the IGR IRES element can directly recruit the ribosome independently of eIFs or the initiator Met-tRNAi (29, 30, 54, 80). Furthermore, the IRES occupies the P-site of the ribosome to initiate translation from the ribosomal A-site encoding non-AUG codon (35, 81). Extensive biochemical and structural analyses from several groups have revealed that the IGR IRES mimics a tRNA that occupies the mRNA cleft of the ribosome and sets the ribosome into an elongation state (9, 29, 30, 34, 51, 55, 58, 68, 72, 83). Using reporter constructs, it has also been demonstrated that CrPV IGR IRES-mediated translation is active under a number of cellular conditions when the activity of the ternary complex eIF2-Met-tRNAi-GTP is compromised (17, 63, 78, 80). Because IGR IRES-mediated translation does not require initiation factors, the IRES can direct translation under a number of cellular conditions when the activity of multiple eIFs is compromised (12). Although the majority of studies have focused on the IGR IRES of CrPV, PSIV, and TSV, it is predicted that the IGRs within this viral family all function similarly based on the predicted conserved RNA structures (28, 36, 49). In contrast, only the 5′UTR IRES mechanism of RhPV has been studied in detail (77). Despite the wealth of studies on the mechanics of these IRES, the mechanisms that lead to translational shutoff during dicistrovirus infection and the interaction of dicistrovirus with the host machinery to allow virus production have been relatively unexplored.Open in a separate windowFIG. 1.Kinetics of host protein synthesis and viral protein expression in CrPV-infected Drosophila S2 cells. (A) Genomic arrangement of the CrPV RNA. The viral open reading frames, ORF1 and ORF2, that encode nonstructural (NS) and structural (S) proteins, respectively, are shown, which are separated by the intergenic internal ribosome entry site (IGR IRES). Translation of ORF1 and ORF2 is directed by the 5′UTR IRES and the IGR IRES, respectively. The first amino acid of ORF2 directed by the IGR IRES is encoded by a GCU alanine codon. (B) Autoradiography of protein lysates resolved on a SDS-12% PAGE gel. The protein lysates were collected from S2 cells that were untreated (U), mock infected (M), CrPV infected (5 FFU/cell), or thapsigargin treated (Tg; 0.4 μM) for the indicated times (h p.i.) and metabolically labeled with [35S]methionine for 30 min at each time point. The migration of proteins with known molecular masses is shown on the left. The expression of detectable nonstructural (NS) and structural (S) proteins is denoted. (C) Quantitation of host protein synthesis during CrPV infection. To calculate the host translation at each time point, the amount of radioactivity of the bands between 55 and 70 kDa in panel A was quantitated by using ImageQuant, and the percent translation was calculated at each time point of virus infection or thapsigargin treatment compared to the mock infection. Shown are averages (± the standard deviation) from at least three independent experiments. (D) Immunoblots of viral ORF1 and ORF2 during CrPV infection at various times postinfection (h p.i.). Antibodies were raised against peptides within ORF1 and ORF2. The expression of ORF1 and ORF2 was quantitated by a LI-COR Odyssey system, plotted against time of infection, and normalized to the amount of ORF1 or ORF2 expression at 6 h p.i. (100%). As a comparison, viral RNA synthesis as detected by Northern blot analysis (see Fig. Fig.2B)2B) is plotted on the same graph.Previous studies have shown that the CrPV and the related DCV can infect a wide range of insect hosts, including the Drosophila melanogaster S2 cell line (60, 69). In the present study, we have explored how CrPV infection leads to host translational shutoff in S2 cells. Two steps of translational initiation are targeted during CrPV infection. First, the interaction of deIF4G with deIF4E is disrupted early in infection and remains dissociated during the course of infection. Second, deIF2α is phosphorylated at a time that lags after the initial host translational shutoff during infection. Premature phosphorylation of deIF2α early in infection inhibited translation directed by the 5′UTR IRES, but IGR IRES-mediated translation remained relatively resistant. These results support the model that multiple mechanisms, including impairment of deIF4F complex formation and induction of deIF2α phosphorylation, contribute to the host translational shutoff during CrPV infection. The inhibition of host translation and the release of ribosomes from host mRNAs ensures that translation mediated by the 5′UTR and IGR IRES is optimal to produce sufficient viral nonstructural and structural proteins for proper CrPV maturation and assembly.  相似文献   

12.
13.
14.
Autophagy is an important component of host innate and adaptive immunity to viruses. It is critical for the degradation of intracellular pathogens and for promoting antigen presentation. Herpes simplex virus type 1 (HSV-1) infection induces an autophagy response, but this response is antagonized by the HSV-1 neurovirulence gene product, ICP34.5. This is due, in part, to its interaction with the essential autophagy protein Beclin 1 (Atg6) via the Beclin-binding domain (BBD) of ICP34.5. Using a recombinant virus lacking the BBD, we examined pathogenesis and immune responses using mouse models of infection. The BBD-deficient virus (Δ68H) replicated equivalently to its marker-rescued counterpart (Δ68HR) at early times but was cleared more rapidly than Δ68HR from all tissues at late times following corneal infection. In addition, the infection of the cornea with Δ68H induced less ocular disease than Δ68HR. These results suggested that Δ68H was attenuated due to its failure to control adaptive rather than innate immunity. In support of this idea, Δ68H stimulated a significantly stronger CD4+ T-cell-mediated delayed-type hypersensitivity response and resulted in significantly more production of gamma interferon and interleukin-2 from HSV-specific CD4+ T cells than Δ68HR. Taken together, these data suggest a role for the BBD of ICP34.5 in precluding autophagy-mediated class II antigen presentation, thereby enhancing the virulence and pathogenesis of HSV-1.Autophagy is a conserved cellular pathway that eliminates defective proteins and organelles, prevents abnormal protein aggregate accumulation, and removes intracellular pathogens (11, 22, 32, 56). This process begins with the formation and elongation of a double membrane that fuses to form an autophagosome. The cytoplasmic contents are nonspecifically sequestered inside the autophagosome and then are degraded once the autophagosome fuses with the lysosome. Autophagy is upregulated during starvation, growth factor withdrawal, hypoxia, and infection (10). Following metabolic stress, autophagy can generate metabolic precursors that can be recycled for the de novo synthesis of proteins. The autophagic pathway has important roles in development, immune defense, apoptosis, tumor suppression, and the prevention of neuronal degeneration (reviewed in references 21 and 31).Autophagy is not limited to the degradation of self proteins; it also can engulf and break down invading microorganisms in a process termed xenophagy (26). Xenophagy can limit the replication of pathogens (3, 4, 29, 34, 42), but some infectious agents can exploit autophagy to enhance their replication (2, 18, 38). There are pathogens that actively inhibit autophagy through interaction with the essential autophagy-promoting protein, Beclin 1 (24, 35). Beclin 1 is the mammalian homolog of yeast Atg6 and is required for the formation of the autophagosome membrane through its interaction with VPS34, a class III phosphatidylinositol 3-kinase (19, 27). Autophagy/xenophagy is also an important process for the adaptive immune response to infection through the delivery of antigens for major histocompatibility complex class I and II (MHC-I and -II) presentation (7, 9, 12, 36, 45). The inhibition of autophagy by pathogens therefore would serve to block CD4+ and CD8+ cell responses and allow pathogens to remain underrecognized by the adaptive immune response.The interferon (IFN)-inducible double-strand RNA-inducible protein kinase (PKR) pathway is required for virus- and starvation-induced autophagy (50). The PKR-mediated induction of autophagy requires the phosphorylation of the translation initiation factor eIF2α (50). Herpes simplex virus type 1 (HSV-1) inhibits autophagy through at least two domains and the activities of the late protein ICP34.5, its C-terminal mediation of dephosphorylation of eIF2α, and its N-terminal binding to Beclin 1 (5, 16, 35). HSV-1 strains lacking ICP34.5 show significant attenuation in vivo, increased sensitivity to IFN (33), an inability to counteract the PKR-induced phosphorylation of eIF2α, and the induction of the generalized shutoff of protein synthesis in infected cells. These activities originally were ascribed solely to ICP34.5''s ability to recruit PP1α and redirect its activity to dephosphorylate eIF2α to counteract the general shutoff of protein synthesis mediated by PKR (17). This function is mediated by the C-terminal domain of ICP34.5 that contains homology to the growth arrest and DNA damage 34 (GADD34) gene (6). Independently of the role of ICP34.5 in countering the PKR-induced antiviral state, viruses lacking ICP34.5 also exhibit altered patterns of autophagy. This manifests as increased long-lived protein degradation, the increased formation of autophagosomes, increased autophagic vacuole volume density, and the enhanced xenophagic degradation of virions (35, 50, 51). Such alterations in the autophagy pathway now can be ascribed to ICP34.5''s ability to bind Beclin 1 in addition to its mediation of eIF2α dephosphorylation (35). The deletion of the Beclin 1-binding domain (BBD) of ICP34.5 renders HSV-1 less able to regulate autophagosome formation, and viruses lacking this domain are neuroattenuated (35).To determine the impact of Beclin 1-ICP34.5 interactions on the pathogenesis of HSV-1 during infection, we examined the ability of BBD-deficient virus to replicate, cause disease, and stimulate an immune response in mice. We determined that mice infected with HSV-1 lacking the ICP34.5 BBD were able to clear virus more efficiently than mice infected with wild-type virus. We observed the significantly enhanced stimulation of CD4+ T cells by the virus lacking the BBD compared to that of mice infected with wild-type virus. These data suggest that Beclin 1 binding and the inhibition of autophagy by ICP34.5 are important for HSV-1 pathogenesis through its ability to suppress autophagy and to dampen the activation of CD4+ T cells.  相似文献   

15.
16.
17.
Eukaryotic mRNA translation begins with recruitment of the 40S ribosome complex to the mRNA 5′ end through the eIF4F initiation complex binding to the 5′ m7G-mRNA cap. Spliced leader (SL) RNA trans splicing adds a trimethylguanosine (TMG) cap and a sequence, the SL, to the 5′ end of mRNAs. Efficient translation of TMG-capped mRNAs in nematodes requires the SL sequence. Here we define a core set of nucleotides and a stem-loop within the 22-nucleotide nematode SL that stimulate translation of mRNAs with a TMG cap. The structure and core nucleotides are conserved in other nematode SLs and correspond to regions of SL1 required for early Caenorhabditis elegans development. These SL elements do not facilitate translation of m7G-capped RNAs in nematodes or TMG-capped mRNAs in mammalian or plant translation systems. Similar stem-loop structures in phylogenetically diverse SLs are predicted. We show that the nematode eukaryotic translation initiation factor 4E/G (eIF4E/G) complex enables efficient translation of the TMG-SL RNAs in diverse in vitro translation systems. TMG-capped mRNA translation is determined by eIF4E/G interaction with the cap and the SL RNA, although the SL does not increase the affinity of eIF4E/G for capped RNA. These results suggest that the mRNA 5′ untranslated region (UTR) can play a positive and novel role in translation initiation through interaction with the eIF4E/G complex in nematodes and raise the issue of whether eIF4E/G-RNA interactions play a role in the translation of other eukaryotic mRNAs.Cap-dependent translation initiation in eukaryotes is a complex process involving many factors and serves as the primary mechanism for eukaryotic translation (37, 44). The first step in the initiation process, recruitment of the m7G (7-methylguanosine)-capped mRNA to the ribosome, is widely considered the rate-limiting step. It begins with recognition of and binding to the m7G cap at the 5′ end of the mRNA by the eukaryotic translation initiation factor 4F (eIF4F) complex, which contains three proteins: eIF4E (a cap-binding protein), eIF4G (a scaffold protein with RNA binding sites), and eIF4A (an RNA helicase). eIF4G''s interaction with eIF3, itself a multisubunit complex that interacts with the 40S ribosome, facilitates the actual recruitment of capped RNA to the ribosome. With the help of several other initiation factors, the small ribosomal subunit scans the mRNA from 5′ to 3′ until a translation initiation codon (AUG) in appropriate context is identified and an 80S ribosomal complex is formed, after which the first peptide bond is formed, thus ending the initiation process (37, 44). The AUG context can play an important role in the efficiency of translation initiation (23, 44). The length, structure, and presence of AUGs or open reading frames in the mRNA 5′ untranslated region (UTR) can negatively affect cap-dependent translation and ribosomal scanning. In general, long and highly structured 5′ UTRs, as well as upstream AUGs leading to short open reading frames, can impede ribosome scanning and lead to reduced translation (23, 44). In addition, 5′ UTRs less than 10 nucleotides (nt) in length are thought to be too short to enable preinitiation complex assembly and scanning (24). Thus, several attributes of the mRNA 5′ UTR are known to negatively affect translation initiation, whereas only the AUG context and the absence of negative elements are known to have a positive effect on translation initiation (44).Two of the important mRNA features associated with cap-dependent translation, the cap and the 5′ UTR, are significantly altered by an RNA processing event known as spliced leader (SL) trans splicing (3, 8, 17, 26, 36, 47). This takes place in members of a diverse group of eukaryotic organisms, including some protozoa, sponges, cnidarians, chaetognaths, flatworms, nematodes, rotifers, crustaceans, and tunicates (17, 28, 39, 55, 56). In SL trans splicing, a separately transcribed small exon (16 to 51 nucleotides [nt]) with its own cap gets added to the 5′ end of pre-mRNAs. This produces mature mRNAs with a unique cap and a conserved sequence in the 5′ UTR. In metazoa, the m7G cap is replaced with a trimethylguanosine (TMG) cap (m2,2,7GpppN) (27, 30, 46, 49). In nematodes, ∼70% of all mRNAs are trans spliced and therefore have a TMG cap and an SL (2). In general, eukaryotic eIF4E proteins do not effectively recognize the TMG cap (35). This raises the issues of how the translation machinery in trans-splicing metazoa effectively recognizes TMG-capped trans-spliced mRNAs, what role the SL sequence plays in translation initiation, and how the conserved translation initiation machinery has adapted to effectively translate trans-spliced mRNAs.Previous work has shown that efficient translation of TMG-capped messages in nematodes requires the SL sequence (22 nt) immediately downstream of the cap (5, 25, 29). In the current studies, we sought to understand the manner in which the SL enhanced the translation of TMG-capped mRNAs. Using a cell-free nematode in vitro translation system, we carried out mutational analyses that define the specific sequences in the SL that are required and sufficient for efficient translation of TMG-capped mRNAs. These analyses led to the discovery of a small, discrete stem-loop immediately adjacent to the TMG cap in trans-spliced messages required for efficient translation. Notably, the sequences involved in the base pairing of the stem are highly conserved in alternative SL sequences found in nematodes. We further show that the nematode eIF4E/G complex plays a major role in facilitating the SL enhancement of TMG-capped mRNA that likely occurs after the initial cap-binding step. The results demonstrate the importance of specific enhancing elements in the 5′ UTR and adaptation in the eIF4F complex necessary for optimal cap-dependent translation.  相似文献   

18.
19.
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