Next Article in Journal
The p22 RNA Silencing Suppressor of the Crinivirus Tomato chlorosis virus is Dispensable for Local Viral Replication but Important for Counteracting an Antiviral RDR6-Mediated Response during Systemic Infection
Next Article in Special Issue
Regulation of Stress Responses and Translational Control by Coronavirus
Previous Article in Journal
Development of Neutralization Assay Using an eGFP Chikungunya Virus
Previous Article in Special Issue
Stress Beyond Translation: Poxviruses and More
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Who Regulates Whom? An Overview of RNA Granules and Viral Infections

by
Natalia Poblete-Durán
1,
Yara Prades-Pérez
1,
Jorge Vera-Otarola
2,
Ricardo Soto-Rifo
1 and
Fernando Valiente-Echeverría
1,*
1
Molecular and Cellular Virology Laboratory, Virology Program, Institute of Biomedical Sciences, Faculty of Medicine, Universidad de Chile, Independencia 1027, Santiago, 8389100, Chile
2
Laboratorio de Virología Molecular, Instituto Milenio de Inmunología e Inmunoterapia, Centro de Investigaciones Médicas, Departamento de Enfermedades Infecciosas e Inmunología Pediátrica, Escuela de Medicina, Pontificia Universidad Católica de Chile, Marcoleta 391, Santiago 8330024, Chile
*
Author to whom correspondence should be addressed.
Submission received: 28 April 2016 / Revised: 10 June 2016 / Accepted: 21 June 2016 / Published: 28 June 2016
(This article belongs to the Special Issue Viral Subversion of Stress Responses and Translational Control)

Abstract

:
After viral infection, host cells respond by mounting an anti-viral stress response in order to create a hostile atmosphere for viral replication, leading to the shut-off of mRNA translation (protein synthesis) and the assembly of RNA granules. Two of these RNA granules have been well characterized in yeast and mammalian cells, stress granules (SGs), which are translationally silent sites of RNA triage and processing bodies (PBs), which are involved in mRNA degradation. This review discusses the role of these RNA granules in the evasion of anti-viral stress responses through virus-induced remodeling of cellular ribonucleoproteins (RNPs).

1. Introduction

The control of the translation and turnover of mRNAs plays fundamental roles in the regulation of gene expression. Eukaryotic cells have developed different mechanisms to control these processes in order to respond to environmental stress, such as heat shock, UV irradiation, hypoxia, endoplasmic reticulum (ER) stress and viral infection. Among those mechanisms is the assembly of non-membrane-delimited bodies, called RNA granules, that contain RNA-binding proteins (RBPs) and translationally silenced mRNAs [1,2].
The classification of RNA granules is based on a number of factors, including their subcellular localization (nuclear, cytoplasmic, axonal, etc.), the presence of specific markers, the cell type where they are assembled (germ cells, neurons), response to stimuli (stress, viral infection), dynamic behavior and proposed functions (sites of mRNA storage/decay, stress response, etc.) [3]. The nucleus contains the nucleolus, nuclear speckles, nuclear stress bodies, the transcription factory, Cajal bodies, the Gemini of Cajal bodies, the histone locus body, paraspeckles and PML bodies, all of which have specific regulatory functions (reviewed in [4,5]). The cytoplasm harbors the polar and germinal granules, stress granules (SGs), processing bodies (P-bodies or PBs) and neuronal granules [6,7]. Many RNA viruses have evolved mechanisms to modulate the assembly of different RNA granules. In this review, we will summarize the regulation of SGs and PBs by different virus families.

2. RNA Granules: Stress Granules (SBs) and Processing Bodies (PBs)

SGs and PBs are associated with mRNA storage and decay, respectively [8]. Importantly, increasing evidence suggests that abnormal SG formation may promote several neurodegenerative disorders, including amyotrophic lateral sclerosis (ALS), frontotemporal lobar degeneration (FTLD) and spinal muscular atrophy (SMA) [9]. The formation of SGs in response to chemotherapeutic treatments has also been associated with cancer cell survival [3]. SGs are specifically induced upon cellular stress [10], triggering global translational silencing typically through the phosphorylation of the translation initiation factor eIF2α. Four eIF2α kinases sense environmental stress: HRI (heme-regulated eIF2α kinase) is activated in heme deprivation and oxidative stress [11]; PKR, (Protein Kinase RNA-dependent) is a double-stranded RNA (dsRNA)-dependent protein kinase activated by viral infection [12]; PERK/PEK (PKR-like ER kinase) is activated during hypoxia and in response to misfolded proteins in the ER [13]; and GCN2 (general control non-derepressible-2) is activated during amino acid depravation and UV irradiation [14]. The assembly of SGs can also occur independently of eIF2α phosphorylation, upon treatment with hippuristanol or pateamine A, drugs that inhibit the helicase activity of eIF4A [15,16]; oligomycin, FCCP or 2-deoxy-D-glucose, metabolic inhibitors that deplete cellular adenosine triphosphate [17]; hydrogen peroxide, a known inducer of reactive oxygen species [18]; as well as overexpression of the SG markers TIA1/TIAR [19] or G3BP-1 [20]. Recently, Kedersha and colleagues reported that SG assembly is finely regulated by the binding of G3BPs to Caprin1 or USP10 [21]. By contrast, PBs are constitutively present and respond to stimuli that affect mRNA translation and decay [7]. Three proteins are known to play a critical role in PB assembly, Edc3, Pat1 and Lsm4 [22,23], due to their Q/N-rich domains that have the potential to form aggregates [24]. PBs can release mRNA to allow their translation [25,26,27]. Additionally, emetine and cycloheximide, drugs that stabilize polysomes, dissolve both SGs and PBs, whereas puromycin, a drug that disrupts polysomes, promotes their assembly [27].

2.1. SG Components

SGs are composed of target mRNAs bound to translation initiation factors (eIF4G, eIF3, PABPC1, eIF2α-P, eIF5A) to allow the rapid restoration of translation once the stress is gone [19,28]. Many other proteins are involved in SG formation, such as mRNA binding proteins that provide translational control or mRNA stability (TIA-1, TIAR, HuR/ELAVL1, FXR1, Pum1) [6,29,30,31], proteins related to mRNA metabolism (G3BP1, G3BP2, p54/rck/DDX6, PMR1, SMN, Staufen1, DHX36, Caprin1, ZBP1, HDAC6, ADAR) [20,32,33,34,35,36] and signaling proteins (mTOR, RACK1, TRAF2) [37]. Moreover, interferon-stimulated gene (ISG) products, such as, PKR, ADAR1, RNA-sensing RIG-I-like receptors (RIG-I, MDA5, LGP2), RNase L and OAS, have been shown to colocalize with SGs following viral infections [38,39].

2.2. PB Components

PBs contain components of the mRNA decapping machinery (Dcp1/2, Lsm1-7, Edc3proteins) [40,41], scaffolding proteins (Ge-1/Hedls) [41], deadenylation factors (Ccr1, Caf1, Not1) [42], nonsense-mediated decay (NMD) proteins (SMG5-6-7, UPF1) [40,43] and translation control factors (CPEB, eIF4E-T) [32,44]. Most recently, Patel and colleagues showed that GW-bodies, which also are ribonucleoproteins (RNPs) involved in mRNA decay, are distinct from P-bodies. GW-bodies are unique in that they contain GW182, a large scaffolding protein containing an N-terminal domain composed of GW/WG motifs. The N-terminal GW/WG motif-bearing domain has been shown to bind to AGO2, while the C-terminus interacts with the CCR4-NOT deadenylation complex [45]. Moreover, GW-bodies have a different spatial-temporal regulation because they play a major functional role in miRNA-mediated gene silencing rather than degradation [45]. As a result of virus infection, ISGs can also be found in PBs [46]. Given that SGs and PBs are highly dynamic, many proteins have been described as being part of both RNA granules (APOBEC3G, Ago2, BRF1, DDX3, FAST, TTP, etc.) [29,47,48,49], suggesting that mRNAs can move between both mRNPs, thus regulating RNA homeostasis. A comprehensive list of SG, PB and SG/PB components is provided in Table 1.

3. Viral Infection and Stress Granules

Several viruses modulate the assembly of SGs to promote viral replication and suppress the cellular stress response. The discussion below was performed upon the interaction between virus families and SGs (Table 2).

3.1. Double-Stranded DNA (dsDNA) Viruses

Members of the family Herpesviridae and Poxviridae are enveloped double-stranded DNA (dsDNA) viruses. Herpes simplex virus type 1 (HSV-1) shuts off host protein synthesis by impairing the activation of eIF2 kinases through the virion host shutoff (vhs) protein [129], Us11 [130], ICP34.5 [131] and glycoprotein B (gB) [132]. During HSV-1 infection, the SG components TIA-1, TIAR and TTP are upregulated, but do not form SG, however, infection with vhs-defective HSV-1 triggers SG assembly [133,134], relying on PKR activity in the absence of vhs [135]. Recently, Finnen and colleagues have shown that infection with HSV type 2 (HSV-2) impairs arsenite-mediated SG assembly, while the SG induced by treatment with pateamine A are not affected [136]. The blockage in arsenite-induced SG assembly is dependent on vhs, as cells infected with a vhs defective HSV-2 mutant form SGs late during infection [137]. Human cytomegalovirus (HCMV) infection suppresses the assembly of SGs in cells treated with the ER stressor thapsigargin [138], while simultaneously inducing an unfolded protein response (UPR) and activating PERK, but limiting eIF2α phosphorylation to maintain viral RNA translation [139].
Vaccinia virus (VV), a member of Poxviridae family, replicates within the cytoplasm in large foci called DNA factories that co-opt SG proteins, such as G3BP1, Caprin1, eIF4E, PABP and eIF4G [140,141]. VV appears to utilize SG components for different purposes. The G3BP1/Caprin1 complex aids VV transcription [140]; viral translation initiation is dependent on eIF4E/eIF4G/PABP; and the viral protein I3 associates with eIF4G to recruit viral ssDNA [142], suggesting that SG components may link VV transcription and translation [141]. TIA-1 is not sequestered in DNA factories [143]; however, infection with a VV mutant lacking E3L, which activates PKR, induces aggregates that contain TIA-1, eIF3b, G3BP1 and USP10, called antiviral granules (AVGs), given that they restrict viral replication [94].

3.2. Double-Stranded RNA (dsRNA) Viruses

The family Reoviridae is composed of non-enveloped virions with a 9–12 dsRNA segment genome. The prototypical member of the family, rotavirus, causes the shut off of host protein synthesis [144], and despite inducing eIF2α phosphorylation, SG assembly is not induced, likely due to the translocation of PABP from the cytoplasm to the nucleus [144]. The persistent phosphorylation of eIF2α during rotavirus infection is PKR-dependent as a consequence of the high amount of viral dsRNA in the cytoplasm [145]. By contrast, mammalian orthoreovirus (MRV) induces SG assembly during the early stages of infection, at a step between viral uncoating and viral mRNA transcription, and requires phosphorylation of eIF2α, which is important to promote virus replication [146]. However, despite high levels of phosphorylated eIF2α, SGs are disrupted at later times during MRV infection [147]. Recently, Carroll and colleagues showed that the nonstructural protein μNS is recruited to SGs, but its expression alone was not able to modulate assembly, suggesting a relationship between viral factories and SGs [148].

3.3. Positive-Sense Single Strand RNA ((+)ssRNA) Viruses

All members of the Picornaviridae family are composed of non-enveloped particles. Evidence indicates that poliovirus (PV) proteinase 2A induces assembly of SGs early post-infection (between 2 and 4 h) [16,149], which are dispersed later in the infection through the cleavage of G3BP1 by the PV 3C proteinase (3Cpro) [150]. Interestingly, at a later time post-infection, aggregates containing viral RNA and TIA-1, but excluding the bona fide SG components eIF4G and PABP, are observed, suggesting that TIA-1 aggregation is unlinked from SG formation [151,152]. Encephalomyocarditis virus (EMCV) and Coxsackievirus B3 (CVB3) also disrupt SGs by cleavage of G3BP1 through a mechanism similar to that used by PV [50,153]. By contrast, Theiler murine encephalomyelitis virus (TMEV) and mengovirus, a strain of EMCV, inhibit SG assembly through the expression of the leader (L) protein, maintaining the G3BP-1 intact [154,155]. A mutant mengovirus, in which the Zn-finger domain of L is disrupted, induces G3BP1 aggregation in a PKR-dependent manner [155], suggesting that a G3BP1-Caprin1-PKR complex could induce innate immune activation during mengovirus infection [156].
In 2005, McInerney and colleagues showed that Semliki Forest virus (SFV), a member of the Togaviridae family composed of enveloped virions, is able to induce eIF2α phosphorylation promoting SG assembly at early stages of infection [157]. However, at late times post-infection, concomitant to an increase in vRNA levels, SFV is no longer able to promote SG assembly [157]. It was determined that the SFV-induced SG disassembly is caused by the C-terminal domain of the viral nonstructural protein 3 (nsP3), which forms a complex with G3BP1, sequestering it into the vRNA replication complex [158]. Chikungunya virus (CHIKV) nsP3 also sequesters G3BP1 and forms specific virus-induced cytoplasmic foci that lack the SG marker eIF3 and are resistant to cycloheximide treatment [159]. Later, Scholte and colleagues demonstrated that G3BP2, a close relative of G3BP1 containing a similar domain architecture, which is recruited into SGs [160], colocalized with CHIKV nsP2/3 [161]. G3BP2-containing complexes differs from the replication and transcription complexes (RTCs), suggesting a role of G3BPs in the early step of the CHIKV replication cycle [161]. In contrast, Rubella virus (RUBV) induces G3BP1 aggregates that colocalize with viral ssRNA and non-structural viral protein P150, suggesting a potential post-replicative role in encapsidation [162]. Finally, the infection with Sindbis virus (SINV), another alphavirus, induces the assembly of G3BP1 aggregates, which interact with nsP2, nsP3 [163,164] and nsP4 [165]. In addition, it has been shown that SINV-derived vectors induce the assembly of bona fide SGs, containing TIA-1, eIF4E and eIF4G in a PKR-dependent manner [166].
The Flaviviridae family are composed of enveloped virions and West Nile virus (WNV) was the first virus reported to block SG assembly. Li and colleagues showed that the 3` stem loop in the viral genome is able to capture TIA-1 and TIAR [167]. In addition, Emara and colleagues expanded these observations to dengue virus DENV-infected cells, where TIA-1/TIAR are found in RTCs [168]. However, the chimeric WNV W956IC, which produces high levels of early viral RNA, activates PKR and subsequently induces the SG assembly [169]. In another report, Xia and colleagues showed that DENV-infected A549 cells were able to induce G3BP1 aggregates independently of TIA-1 [170]. In addition, a proteomic analysis found that G3BP1, G3BP2, Caprin1 and USP10 interact with DENV RNA [171], and it was reported that G3BP1, G3BP2 and Caprin1 regulate the translation of ISG mRNAs, thus protecting DENV replication from IFN-mediated antiviral effects [172]. Consistent with these findings, TIA-1 and TIAR were also recruited to sites of tick-borne encephalitis virus (TBEV) replication. Indeed, eIF2α becomes phosphorylated, and SGs containing G3BP1, eIF3 and eIF4B are induced in TBEV-infected cells [173]. On the other hand, the Japanese encephalitis virus (JEV) core protein was shown to sequester G3BP1 and USP10 through an interaction with Caprin1, resulting in the suppression of SG formation [174]. There is also evidence showing that infection with bovine viral diarrhea virus (BVDV), a Flaviviridae pestivirus, impairs the arsenite-mediated SG assembly; despite the fact that BVDV N-terminal protease (Npro) is able to interact with several RNA granules components, such as YB-1, IGFBP2, DDX3, ILF2 and RHA (DXH9) [175]. Contrasting the aforementioned flavivirus, HCV modulates the SG assembly by controlling the phosphorylation of eIF2α. Ariumi and colleagues reported evidence indicating that HCV infection delocalizes components of SGs (G3BP1, ATX2, PABP1) to lipid droplets (LDs), while Garaigorta and colleagues reported that HCV induces the assembly of bona fide SGs in a PKR-dependent manner. This report also demonstrated that TIA-1, TIAR and G3BP1 play a pivotal role in several steps of the HCV life cycle [176]. In addition, Ruggieri and colleagues observed rapid cycles of SG assembly and disassembly during the HCV infection depending on the eIF2α phosphorylation state [177]. They reported that while SG assembly is regulated by dsRNA promoting phosphorylation of eIF2α mediated by PKR, SG disassembly is regulated by a rapid eIF2α dephosphorylation through both protein phosphatase 1 (PP1) and its regulatory subunit, growth arrest and DNA-damage-inducible 34 (GADD34) [177]. Recently, another component of SGs, DDX3, has been reported to bind the HCV 3′ UTR and IKKα, leading to its activation to induce LD biogenesis [178]. However, late in infection, DDX3 and G3BP1 localize with the HCV core protein around LDs in order to initiate viral particle assembly [179,180]. These observations could explain the oscillation of SG assembly/disassembly detected in HCV-infected cells [177] and how SG formation is necessary for HCV RNA replication, assembly and egress [176,181,182].
The infection with Cricket paralysis virus (CrPV), a member of Dicistroviridae family, prevents the aggregation of Rox8 and Rin, Drosophila SG marker homologs of TIA-1 and G3BP-1, respectively, even in the presence of various stressor, such as arsenite, pateamine A or heat shock [183]. Although the CrPV 3C proteinase is sequestered in SGs, cleavage of the Rox8 or Rin is not detected in CrPV-infected cells [183]
The family Coronaviridae is composed of membrane-enveloped virions and one of the most significant features of coronaviruses is the expression of downstream genes via transcription of multiple 3` nested subgenomic mRNAs [184]. It has been shown that both mouse hepatitis coronavirus (MHV) and transmissible gastroenteritis virus (TGEV) induce TIA-1/TIAR aggregates with a concomitant increase in eIF2α phosphorylation levels [185,186]. MHV promotes SG assembly at early times post-infection [185], while TGEV induces SG assembly at later times post-infection [186]. During TGEV infection, the polypyrimidine tract-binding protein (PTB) is redistributed to the cytoplasm, associated with both TGEV gRNA and subgenomic mRNA (sgmRNA) and confined to TIA-1/TIAR aggregates [186].
Finally, Caliciviridae family is composed of non-enveloped virions and recently, Humoud and colleagues showed that the feline calicivirus (FCV) infection triggers eIF2α phosphorylation without inducing SG assembly [187]. This report demonstrated that the FCV viral 3C-like proteinase NS6 inhibit SG assembly by cleaving G3BP1 [187]. On the other hand, the murine norovirus 1 (MNV1) infection triggers eIF4E phosphorylation to control the translational machinery of the host cell [188], but does not impair SG assembly [187]

3.4. Negative-Sesnse Single Strand RNA ((-)ssRNA) Viruses

The family of Orthomyxoviridae is composed of enveloped virions with a segmented, negative ssRNA genome [184]. Influenza A virus (IAV) suppresses the assembly of SG during viral infection by expressing the non-structural protein NS1, which inhibits PKR activity [189]. The ability of IAV to interfere with SG assembly is reverted with the expression of an NS1-mutatant virus unable to bind dsRNA [189] or by expressing an NS1-deficient IAV, which promote SG assembly containing retinoic acid inducible gene I (RIG-I)-like receptors (RLRs), such as RIG-I, MDA5 and LGP2, or ISGs, such as PKR, RnaseL and OAS [39]. The NS1-mediated inhibition of SG assembly is dependent on the interaction with the RNA-associated protein 55 (RAP55) [190]. Besides NS1, the nucleoprotein (NP) and polymerase acidic protein X (PA-X) also contribute to block the SG assembly independent of eIF2α phosphorylation [191]. In addition, given that DDX3 has been implicated in the sensing of viral RNA to modulate IFN production and SG assembly (reviewed in [180]), Thulasi Raman and colleagues showed that DDX3 interacts with NP and colocalizes in SGs upon infection with an IAV mutant lacking NS1, suggesting a novel antiviral function for DDX3 [192].
Junin virus (JUNV) is a virus belonging to Arenaviridae, a family of enveloped virions with a bi-segmented negative-sense single strand RNA ((-)ssRNA) genome, that prevents SG assembly by impairing the phosphorylation of eIF2α [193], while the expression of JUNV nucleoprotein (N) and/or the glycoprotein precursor (GPC) are enough to inhibit SG assembly [193]. In addition, arenavirus RTCs contain ribosomal proteins L10a and S6 and translation initiation factors eIF4G and eIF4A and G3BP1 [194], suggesting a role of the SG component in the viral replication cycle.
Vesicular stomatitis virus (VSV), a virus in the family Rhabdoviridae, induces the phosphorylation of eIF2α and promotes the assembly of SG-like particles that contain TIA-1, TIAR, PCBP2, viral replication proteins and RNA, but not the eukaryotic initiation factor 3 (eIF3), nor eIF4A [195], suggesting that the SG-like structures may play a role in the VSV infection cycle.
Members of Paramyxoviridae family are composed of enveloped virions. Respiratory syncytial virus (RSV) induces SG assembly mediated by PKR, and its formation enhanced RSV replication [196,197]. Given that RSV infection forms cytoplasmic inclusion bodies (IBs) to mediate viral replication, Lindquist and colleagues observed that HuR, another SG component, localized to IBs [196] as well as MDA5 and MAVS proteins, suggesting a role in the suppression of IFN production [198]. Moreover, it was recently reported that RSV sequesters phosphorylated p38 (p38 P) and O-linked N-acetylglucosamine (OGN) transferase (OGT) into viral IBs, inhibiting the MAPK-activated protein kinase 2 (MK2) pathway and suppressing the assembly of SGs, respectively [199]. Measles virus (MeV) has the capacity to affect the innate and adaptive immune response through the accessory viral proteins V and C [200]. Interestingly, while infection with a protein C-deficient MeV efficiently induces SG assembly mediated by PKR, wild-type MeV does not [38]. However, wild-type MeV induces the assembly of SGs in ADAR1-knockdown cells, suggesting a role between the viral C protein and ADAR1 in modulating the assembly of SGs [38]. Sendai virus (SeV) infection slightly induces SG assembly, while short transcripts generated from the 3′-ends of antigenome RNA, SeV trailer RNA, interact with TIAR to prevent the assembly of SG [201]. In addition, recent data showed that SeV viral C protein also play a role in impairing SG assembly and IFN production [202]. This study revealed that SeV C-deficient recombinant 4C(-) infection forms SG-like structures containing RIG-I and unusual viral RNA species, suggesting that these structures could be reminiscent of anti-viral stress granules (avSGs) (reviewed in [180])
Bunyaviridae are a large family composed of enveloped virions with a tri-segmented (-)ssRNA genome [184]. The peculiarity of Bunyaviruses is that they require a capped oligonucleotide, which is scavenged from host mRNAs, to initiate its own mRNA synthesis, a process called “cap-snatching” [184]. These mechanisms take place in PBs, where the viral N protein binds to cap structures of cellular mRNAs, and the endonuclease domain of the RNA-dependent RNA polymerase (RdRp) cleaves cap-downstream sequences in a range of 10–18 ribonucleotides [203,204]. This capped-RNA fragment is used as a primer for mRNA synthesis by the viral RdRp. Interestingly, one report with Sin Nombre virus (SNV), a member of the hantaviruses genera, showed that the cellular mRNAs target of cap-snatching are mRNAs with a premature stop codon (PTC), which have been addressed to PBs for degradation through the non-sense mediated decay pathway (NMD) [205]. Another report with the Rift Valley fever virus (RVFV) from the Phlebovirus genera showed attenuation of the Akt/mTOR signaling pathway, which in turn increases the activity of 4EBP1/2 proteins leading to the arrest of cap-dependent translation [206]. Although SG formation would be expected upon inhibition of translation, this report demonstrated that SGs are instead disassembled during RVFV infection. It is noteworthy that during attenuation of Akt/mTOR 5′ TOP-containing mRNAs (including those coding for translation initiation factors), these are selectively targeted to SGs. Notably, the authors observed that 5′-end sequences from TOP mRNAs are included in RVFV’s mRNAs captured by the cap-snatching mechanism in PBs. Those data suggest that RVFV temporally induces SGs to nucleate important mRNAs and in turn promote mRNAs cargo from SGs to PBs where, finally, RVFV uses these cellular mRNAs for cap-snatching. Indeed, decay of the core of translation machinery would not represent a problem for bunyaviruses, since it has been described that SNV N protein is capable of replacing the entire eIF4F complex (eIF4E, eIF4A and eIF4G) and binding the 40S ribosomal subunit, thus recruiting the translation machinery to its own mRNAs [207,208]. Moreover, it is widely accepted that Bunyaviruses do not have a poly(A) tail at the 3′-end of their mRNAs. Indeed, there is evidence showing that Bunyamwera virus (BUNV) and Andes virus (ANDV) have a 3′ UTR, which replaces the poly(A) tail function, and interestingly, the translations of those mRNAs are poly(A) binding protein (PABP)-independent [209,210]. These findings show that bunyaviruses use complex mechanisms of transcription/translation that are potentially interconnected with cytoplasmic RNA granules.
On the other hand, Bunyaviral proteins, like other viral proteins, have evolved mechanisms to inactivate PKR and inhibit the IFN response. Similarly to the non-structural protein NS1 from IAV [211] or NS4A from Dengue virus [212], studies with Orthobunyaviruses, Hantaviruses and Phleboviruses have shown that the non-structural protein from the S segment (NSs) acts as an inhibitor of the IFN response [213,214,215,216], as well as glycoprotein Gn and the capsid N protein from Hantaviruses [217,218,219]. For example, the capsid N protein from Andes Hantavirus (ANDV) is an inhibitor of PKR dimerization, impairing its activation; however, the translation shut-off is not observed in Hantavirus-infected cells. [220]. In contrast, RVFV infection promotes a shut-off of global translation, where NSs protein mediate PKR degradation by the proteasome [221,222]. Undoubtedly, the field of cytoplasmic RNA granules related to the transcription and translation of bunyaviruses already began to show interesting findings, which will help to answer some ancient questions linked to the molecular mechanisms of its replication cycle, as it has been thought that transcription and translation are coupled like in prokaryotic systems [223,224]. However, the SG formation has not been addressed to date.
Recently, Nelson and colleagues showed that the infection with Ebola virus(EBOV), member of Filoviridae family composed of enveloped virions, does not trigger eIF2α phosphorylation or SG formation. However, SG components are sequestered within viral inclusions where they colocalize with viral mRNA [225]. In addition, viral protein VP35 not only prevents SG formation by blocking PKR activation, but also disrupts SG formation independently of eIF2α [225].

3.5. Single Strand RNA Retroviruses (ssRNA-RTs)

Retroviruses are enveloped, positive ssRNA viruses, which synthesize complementary DNA (cDNA) by reverse transcription that is integrated into the host chromosomal DNA [184]. The oncoretrovirus, human T-cell leukemia virus (HTLV-1), causes a blockade of SG assembly mediated by the viral regulatory protein, Tax. Legros and colleagues observed that Tax interacts with histone deacetylase 6 (HDAC6), a critical component of SG, to block SG assembly [226]. Tax also interacts with USP10, inhibiting SG assembly and enhancing the production of reactive oxygen species [227]. On the other hand, human immunodeficiency virus type 1 (HIV-1) significantly impairs SG assembly in favor of the assembly of Staufen1-containing HIV-1-dependent ribonucleoproteins [228]. Indeed, we showed that HIV-1 Gag blocks SG assembly irrespective of eIF2α phosphorylation [229]. In addition, we reported that the interaction between the N-terminal domain (NTD) of the capsid domain of Gag (p24) and host eukaryotic elongation factor 2 (eEF2) are critical for the SG blockade, while that eEF2 depletion not only lifted the SG blockade, but also resulted in impaired virus production and infectivity [229]. Interestingly, we also reported that HIV-1 Gag mediates the disassembly of preexisting SGs via an interaction with G3BP1 [229], and more recently, it was shown that G3BP1 binds HIV-1 unspliced mRNA (gRNA) in the cytoplasm of macrophages to inhibit viral replication [230]. At the same time, we reported that the HIV-1 unspliced mRNA (gRNA) promotes the assembly of a pre-translation initiation intermediate with DDX3, another SG component, and a subset of translation initiation factors, such as eIF4GI and PABPC1, suggesting that these intermediates may serve to concentrate the gRNA and eIFs in order to enhance the efficiency of polysome association [231]. It is noteworthy that the mechanism associated with SGs-blockage by HIV-1 Gag was dependent on the kind of stressor. As such, sodium selenite (Se) causes 4EBP1-mediated mRNA translational arrest and the subsequent assembly of non-canonical type II SGs [232]. Recently, Cinti and colleagues showed that in HIV-1-expressing cells under Se treatment, Gag interacts with eIF4E, reducing hypophosphorylated 4EBP1 associated with the 5′ cap in order to promote the disassembly of SGs [233]. In contrast to HIV-1, the replication of human immunodeficiency virus type 2 (HIV-2) induces the spontaneous assembly of SG [231]. As such, the HIV-2 gRNA recruits TIAR to form a novel viral mRNP, where the switch between translation to packaging could occur [231].

4. Viral Infections and Processing Bodies

Several viruses are known to disrupt P-bodies. Here, we will summarize the most up-to-date information known about the relationships between viruses and PBs (Table 3).

4.1. dsDNA Viruses

Adenovirus regulates its gene expression over the time of infection, triggering an accumulation of viral late mRNA in the cytoplasm [184]. To prevent viral mRNA degradation, the early protein E4 11k binds Rck/p54/DDX6, relocalizing it with PB components such as Lsm-1, Ge-1, Ago2 and Xrn1, to aggresomes, sites where these proteins are inactivated [237]. In contrast, within the human papilloma virus (HPV), the oncoprotein E6 commands the re-colocalization of PKR in PBs, suggesting an antiviral effect for these granules [46]. Finally, during HCMV infection, PB components, such as Dcp1a, Edc4, Rck/p54/DDX6 and RAP55, increased in a translation-independent manner requiring cellular, but not viral RNA synthesis [238].

4.2. dsRNA Viruses

A recent report showed that rotavirus infection triggers a significant time-dependent decline of PBs components XRN1, DCP1, Pan3, but not GW182 [239]. Pan3, but not XRN1 and DCP1, undergoes accelerated turn over in response to rotavirus infection, while XRN1 and DCP1 were translocated to the nucleus [239], as well as PABP [144].

4.3. (+)ssRNA Viruses

In the case of flaviviruses, WNV and DENV infections reveal a reduction in PB formation [168]. Indeed, WNV sequesters several components of PBs, such as Lsm1, GW182, Xrn1, DDX3 and Rck/p54/DDX6, in viral replication factories to promote viral replication [240]. Furthermore, it has been shown that Rck/p54/DDX6 binds a stem-loop present in the DENV 3′ UTR, inhibiting PB formation [171]. Interestingly, flaviviruses (including yellow fever virus (YFV) and DENV) generate a subgenomic flavivirus RNA (sfRNA), as a product of gRNA degradation by Xrn1 at pseudoknot 3 [241]. The sfRNA and Xrn1 colocalized in PBs and were essential for viral cytopathogenicity [242]. In addition, Moon and colleagues showed that inhibition of XRN1 by sfRNA-interaction results in accumulation of uncapped cellular mRNA [243]. Likewise, HCV infection relocalized PB components to viral factories close to lipid droplets, such as RCK/p54/DDX6, Lsm1, Xrn1, PatL1, Ago2 and DDX3 [181,182]. The depletion of these components has a detrimental role over HCV replication [181,182,244], while RCK/p54/DDX6, Lsm1 and PatL1 play a central role in HCV translation and replication [245,246]. However, the disruption of PB mediated by depletion of Rap55 does not affect HCV replication [247]. Together, these data strongly suggest that HCV co-opts several PB components to ensure viral replication, but if PBs are formed, they do not have an inhibitory effect during viral infection.
On the other hand, infection with picornaviruses, such as PV and CVB3, entirely disrupts PB formation through the virally-induced cleavage of Dcp1a mediated by viral protease 3C, as well as the degradation of Xrn1 and Pan3 mediated by the proteasome [248]. It has been recently shown that expression of PV-protease 2A also blocks PB formation more efficiently than protease 3C [149]; however, the mechanism remains unclear. Furthermore, CrPV infection only disrupts GW182/Dcp1 aggregates, but not Ago1/Ago2, suggesting that these PBs components could have a differential role in viral infection [183]. Moreover, the U-rich region close to the 3′-end of SINV mRNAs interacts with HuR, generating a dramatic translocation of the host protein out of the nucleus, stabilizing viral transcripts during infection and subsequently preventing the assembly of PBs [249].

4.4. (-)ssRNA Viruses

The negative-strand RNA viruses are known to use cap-snatching mechanism, to initiate its own mRNA synthesis. However, for IAV, this process occurs in the nucleus, while that for Bunyavirus occurs in PBs. Mir and colleagues showed that hantavirus nucleocapsid protein (N) avoids the 5′ cap degradation of cellular mRNAs, protecting them from Dcp1a/Dcp2-mediated decapping, [204]. In addition, the interaction of N protein with 5′-Cap, instead of eIF4E/eIF4F, allows the Hantavirus transcripts to escape from PB and recruit ribosomes [207]. In the case of IAV, the interaction of RAP55 with NS1 impairs PB formation, but also prevents the capture of NP in the PBs [190].

4.5. ssRNA-RT Viruses

Lastly, the protein APOBEC3 (apolipoprotein B mRNA editing enzyme catalytic polypeptide-like 3) provides anti-HIV-1 activity while being a PB component that interacts with Ago2, Dcp1a, Dcp2 and DDX6 [47,115]. However, the viral infectivity factor Vif protein induces APOBEC3 degradation, preventing its incorporation into virions [47], affecting its subcellular localization, degradation rates and antiviral properties [250]. Several groups have reported that depletion of PB components or RISC components (including DDX6, LSM-1, GW182, XRN1, DGCR8, Dicer and Drosha) increases the viral production and that gRNA and Gag protein localize to PBs [251,252,253]. Nevertheless, gRNA localization in PBs has not been detected by others [228,254], and the depletion of Ago2 or DDX6 resulted in the inhibition of HIV-1 replication [255,256]. Likewise, Abrahamyan and colleagues reported a dramatic decrease in the abundance of PBs around gRNA-foci in HIV-1-expressing cells, even in cells treated with sodium arsenite [228], suggesting a relocalization of PB during the HIV-1 infection. Given that the events of HIV-1 capsid assembly have not been associated with PBs, Reed and colleagues demonstrated that assembly intermediates (AIs), containing HIV-1 Gag, GagPol and Vif [257], are formed by the recruitment of DDX6 and ATP-binding cassette protein E1, ABCE1 [256]. On the other hand, the overexpression of Mov10, a putative RNA helicase that associates with APOBEC in RISC complexes and that was found in PBs, can inhibit viral replication and reduce Gag expression [258,259]. In addition, it was reported that the recruitment of Mov10 and APOBEC3G into virions is independent of localization on their PBs [260]. However, given the discrepancy between these reports, more work is needed to elucidate the role of PBs and their components during the HIV-1 replicative cycle.

5. Conclusions

Although significant advances have been made to understand how viruses modulate the assembly/disassembly of RNA granules, there are still outstanding questions that need to be addressed. For example, could these mechanisms be targets of new antiviral drugs? What are the molecular mechanisms and/or signaling pathways that transport mRNAs from one RNA granule to another? Do post-translational modifications that serve as signals to modulate the formation of RNA granules exist? To answer these questions, several reports have helped us to comprehend the molecular biology of RNA granules [261,262,263,264,265,266,267,268,269,270], however, further work is necessary to determine the viral mechanisms that modulate the RNA granules. In addition, emerging evidence has related RNA granules with innate antiviral immunity as part of the integrated stress response. Finally, understanding how viruses counter anti-viral stress responses lays the groundwork for new strategies to bolster host cell immune defenses against invading pathogens.

Acknowledgments

We thank Valerie Le Sage and Payton Ottum for critical reading of the manuscript. This review was supported by Conicyt Chile through the Fondecyt Initiation Into Research Program No. 11140502 (F.V.-E.) and No. 11150611 (J.V.-O.), The International Cooperation Program of Conicyt (DRI USA2013-0005)(R.S.-R.), Proyecto P09/016-F from Iniciativa Científica Milenio del Ministerio de Economía, Fomento y Turismo (J.V.-O.) and Proyecto Anillo ACT1408 (R.S.-R. and J.V.-O.).

Author Contributions

N.P.-D., Y.P.-P., J.V.-O., R.S.-R. and F.V.-E. equally contributed to the writing of this paper.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kedersha, N.; Anderson, P. Regulation of translation by stress granules and processing bodies. Prog. Mol. Biol. Transl. Sci. 2009, 90, 155–185. [Google Scholar] [PubMed]
  2. Buchan, J.R.; Parker, R. Eukaryotic stress granules: The ins and outs of translation. Mol. Cell 2009, 36, 932–941. [Google Scholar] [CrossRef] [PubMed]
  3. Anderson, P.; Kedersha, N.; Ivanov, P. Stress granules, P-bodies and cancer. Biochim. Biophys. Acta 2015, 1849, 861–870. [Google Scholar] [CrossRef] [PubMed]
  4. Morimoto, M.; Boerkoel, C.F. The role of nuclear bodies in gene expression and disease. Biology (Basel) 2013, 2, 976–1033. [Google Scholar] [CrossRef] [PubMed]
  5. Dundr, M.; Misteli, T. Biogenesis of nuclear bodies. Cold Spring Harb. Perspect. Biol. 2010, 2, a000711. [Google Scholar] [CrossRef] [PubMed]
  6. Anderson, P.; Kedersha, N. RNA granules. J. Cell Biol. 2006, 172, 803–808. [Google Scholar] [CrossRef] [PubMed]
  7. Thomas, M.G.; Loschi, M.; Desbats, M.A.; Boccaccio, G.L. RNA granules: The good, the bad and the ugly. Cell. Signal. 2011, 23, 324–334. [Google Scholar] [CrossRef] [PubMed]
  8. Anderson, P.; Kedersha, N. Stress granules. Curr. Biol. 2009, 19, 397–398. [Google Scholar] [CrossRef] [PubMed]
  9. Buchan, J.R. mRNP granules. Assembly, function, and connections with disease. RNA Biol. 2014, 11, 1019–1030. [Google Scholar] [CrossRef] [PubMed]
  10. Valiente-Echeverria, F.; Melnychuk, L.; Mouland, A.J. Viral modulation of stress granules. Virus Res. 2012, 169, 430–437. [Google Scholar] [CrossRef] [PubMed]
  11. Han, A.P.; Yu, C.; Lu, L.; Fujiwara, Y.; Browne, C.; Chin, G.; Fleming, M.; Leboulch, P.; Orkin, S.H.; Chen, J.J. Heme-regulated eIF2alpha kinase (HRI) is required for translational regulation and survival of erythroid precursors in iron deficiency. EMBO J. 2001, 20, 6909–6918. [Google Scholar] [CrossRef] [PubMed]
  12. Williams, B.R. Signal integration via PKR. Sci. STKE Signal Transduct. Knowl. Environ. 2001, 2001, re2. [Google Scholar] [CrossRef] [PubMed]
  13. Harding, H.P.; Zhang, Y.; Bertolotti, A.; Zeng, H.; Ron, D. Perk is essential for translational regulation and cell survival during the unfolded protein response. Mol. Cell 2000, 5, 897–904. [Google Scholar] [CrossRef]
  14. Jiang, H.Y.; Wek, R.C. GCN2 phosphorylation of eIF2alpha activates NF-kappaB in response to UV irradiation. Boilchem. J. 2005, 385, 371–380. [Google Scholar]
  15. Dang, Y.; Kedersha, N.; Low, W.K.; Romo, D.; Gorospe, M.; Kaufman, R.; Anderson, P.; Liu, J.O. Eukaryotic initiation factor 2alpha-independent pathway of stress granule induction by the natural product pateamine A. J. Biol. Chem. 2006, 281, 32870–32878. [Google Scholar] [CrossRef] [PubMed]
  16. Mazroui, R.; Sukarieh, R.; Bordeleau, M.E.; Kaufman, R.J.; Northcote, P.; Tanaka, J.; Gallouzi, I.; Pelletier, J. Inhibition of ribosome recruitment induces stress granule formation independently of eukaryotic initiation factor 2alpha phosphorylation. Mol. Biol. Cell 2006, 17, 4212–4219. [Google Scholar] [CrossRef] [PubMed]
  17. Kedersha, N.; Chen, S.; Gilks, N.; Li, W.; Miller, I.J.; Stahl, J.; Anderson, P. Evidence that ternary complex (eIF2-GTP-tRNA(i)(Met))-deficient preinitiation complexes are core constituents of mammalian stress granules. Mol. Biol. Cell 2002, 13, 195–210. [Google Scholar] [CrossRef] [PubMed]
  18. Emara, M.M.; Fujimura, K.; Sciaranghella, D.; Ivanova, V.; Ivanov, P.; Anderson, P. Hydrogen peroxide induces stress granule formation independent of eIF2alpha phosphorylation. Boilchem. Biophys. Res. Commun. 2012, 423, 763–769. [Google Scholar] [CrossRef] [PubMed]
  19. Kedersha, N.L.; Gupta, M.; Li, W.; Miller, I.; Anderson, P. RNA-binding proteins TIA-1 and TIAR link the phosphorylation of eIF-2α to the assembly of mammalian stress granules. J. Cell Biol. 1999, 147, 1431–1441. [Google Scholar] [CrossRef] [PubMed]
  20. Tourriere, H.; Chebli, K.; Zekri, L.; Courselaud, B.; Blanchard, J.M.; Bertrand, E.; Tazi, J. The RasGAP-associated endoribonuclease G3BP assembles stress granules. J. Cell Biol. 2003, 160, 823–831. [Google Scholar] [CrossRef] [PubMed]
  21. Kedersha, N.; Panas, M.D.; Achorn, C.A.; Lyons, S.; Tisdale, S.; Hickman, T.; Thomas, M.; Lieberman, J.; McInerney, G.M.; Ivanov, P.; et al. G3BP-Caprin1-USP10 complexes mediate stress granule condensation and associate with 40S subunits. J. Cell Biol. 2016, 212, 845–860. [Google Scholar] [CrossRef] [PubMed]
  22. Decker, C.J.; Teixeira, D.; Parker, R. Edc3p and a glutamine/asparagine-rich domain of Lsm4p function in processing body assembly in Saccharomyces cerevisiae. J. Cell Biol. 2007, 179, 437–449. [Google Scholar] [CrossRef] [PubMed]
  23. Pilkington, G.R.; Parker, R. Pat1 contains distinct functional domains that promote P-body assembly and activation of decapping. Mol. Cell. Biol. 2008, 28, 1298–1312. [Google Scholar] [CrossRef] [PubMed]
  24. Reijns, M.A.; Alexander, R.D.; Spiller, M.P.; Beggs, J.D. A role for Q/N-rich aggregation-prone regions in P-body localization. J. Cell Sci. 2008, 121, 2463–2672. [Google Scholar] [CrossRef] [PubMed]
  25. Brengues, M.; Teixeira, D.; Parker, R. Movement of eukaryotic mRNAs between polysomes and cytoplasmic processing bodies. Science 2005, 310, 486–489. [Google Scholar] [CrossRef] [PubMed]
  26. Bhattacharyya, S.N.; Habermacher, R.; Martine, U.; Closs, E.I.; Filipowicz, W. Relief of microRNA-mediated translational repression in human cells subjected to stress. Cell 2006, 125, 1111–1124. [Google Scholar] [CrossRef] [PubMed]
  27. Kedersha, N.; Cho, M.R.; Li, W.; Yacono, P.W.; Chen, S.; Gilks, N.; Golan, D.E.; Anderson, P. Dynamic shuttling of TIA-1 accompanies the recruitment of mRNA to mammalian stress granules. J. Cell Biol. 2000, 151, 1257–1268. [Google Scholar] [CrossRef] [PubMed]
  28. Anderson, P.; Kedersha, N. Stressful initiations. J. Cell Sci. 2002, 115, 3227–3234. [Google Scholar] [PubMed]
  29. McEwen, E.; Kedersha, N.; Song, B.; Scheuner, D.; Gilks, N.; Han, A.; Chen, J.J.; Anderson, P.; Kaufman, R.J. Heme-regulated inhibitor kinase-mediated phosphorylation of eukaryotic translation initiation factor 2 inhibits translation, induces stress granule formation, and mediates survival upon arsenite exposure. J. Biol. Chem. 2005, 280, 16925–16933. [Google Scholar] [CrossRef] [PubMed]
  30. Kedersha, N.L.; Gupta, M.; Li, W.; Miller, I.; Anderson, P. RNA-binding proteins TIA-1 and TIAR link the phosphorylation of eIF-2 alpha to the assembly of mammalian stress granules. J. Cell Biol. 1999, 147, 1431–1442. [Google Scholar] [CrossRef] [PubMed]
  31. Mazroui, R. Trapping of messenger RNA by Fragile X Mental Retardation protein into cytoplasmic granules induces translation repression. Hum. Mol. Genet. 2002, 11, 3007–3017. [Google Scholar] [CrossRef] [PubMed]
  32. Wilczynska, A.; Aigueperse, C.; Kress, M.; Dautry, F.; Weil, D. The translational regulator CPEB1 provides a link between dcp1 bodies and stress granules. J. Cell Sci. 2005, 118, 981–992. [Google Scholar] [CrossRef] [PubMed]
  33. Yang, F.; Peng, Y.; Murray, E.L.; Otsuka, Y.; Kedersha, N.; Schoenberg, D.R. Polysome-bound endonuclease PMR1 is targeted to stress granules via stress-specific binding to TIA-1. Mol. Cell. Biol. 2006, 26, 8803–8813. [Google Scholar] [CrossRef] [PubMed]
  34. Hua, Y.; Zhou, J. Modulation of SMN nuclear foci and cytoplasmic localization by its C-terminus. Cell. Mol. Life Sci. CMLS 2004, 61, 2658–2663. [Google Scholar] [CrossRef] [PubMed]
  35. Thomas, M.G.; Martinez Tosar, L.J.; Desbats, M.A.; Leishman, C.C.; Boccaccio, G.L. Mammalian Staufen 1 is recruited to stress granules and impairs their assembly. J. Cell Sci. 2009, 122, 563–573. [Google Scholar] [CrossRef] [PubMed]
  36. Chalupníková, K.; Lattmann, S.; Selak, N.; Iwamoto, F.; Fujiki, Y.; Nagamine, Y. Recruitment of the RNA helicase RHAU to stress granules via a unique RNA-binding domain. J. Biol. Chem. 2008, 283, 35186–35198. [Google Scholar] [CrossRef] [PubMed]
  37. Wippich, F.; Bodenmiller, B.; Trajkovska, M.G.; Wanka, S.; Aebersold, R.; Pelkmans, L. Dual specificity kinase DYRK3 couples stress granule condensation/dissolution to mTORC1 signaling. Cell 2013, 152, 791–805. [Google Scholar] [CrossRef] [PubMed]
  38. Okonski, K.M.; Samuel, C.E. Stress granule formation induced by measles virus is protein kinase PKR dependent and impaired by RNA adenosine deaminase ADAR1. J. Virol. 2013, 87, 756–766. [Google Scholar] [CrossRef] [PubMed]
  39. Onomoto, K.; Jogi, M.; Yoo, J.S.; Narita, R.; Morimoto, S.; Takemura, A.; Sambhara, S.; Kawaguchi, A.; Osari, S.; Nagata, K.; et al. Critical role of an antiviral stress granule containing RIG-I and PKR in viral detection and innate immunity. PLoS ONE 2012, 7, e43031. [Google Scholar] [CrossRef]
  40. Ingelfinger, D.; Arndt-Jovin, D.J.; Lührmann, R.; Achsel, T. The human LSm1–7 proteins colocalize with the mRNA-degrading enzymes Dcp1/2 and Xrnl in distinct cytoplasmic foci. RNA (New York, NY) 2002, 8, 1489–1501. [Google Scholar]
  41. Yu, J.H.; Yang, W.-H.; Gulick, T.; Bloch, K.D.; Bloch, D.B. Ge-1 is a central component of the mammalian cytoplasmic mRNA processing body. RNA (New York, NY) 2005, 11, 1795–1802. [Google Scholar] [CrossRef] [PubMed]
  42. Sheth, U.; Parker, R. Targeting of aberrant mRNAs to cytoplasmic processing bodies. Cell 2006, 125, 1095–1109. [Google Scholar] [CrossRef] [PubMed]
  43. Durand, S.; Cougot, N.; Mahuteau-Betzer, F.; Nguyen, C.H.; Grierson, D.S.; Bertrand, E.; Tazi, J.; Lejeune, F. Inhibition of nonsense-mediated mRNA decay (NMD) by a new chemical molecule reveals the dynamic of NMD factors in P-bodies. J. Cell Biol. 2007, 178, 1145–1160. [Google Scholar] [CrossRef] [PubMed]
  44. Andrei, M.A.; Ingelfinger, D.; Heintzmann, R.; Achsel, T.; Rivera-Pomar, R.; Lührmann, R. A role for eIF4E and eIF4E-transporter in targeting mRNPs to mammalian processing bodies. RNA (New York, NY) 2005, 11, 717–727. [Google Scholar] [CrossRef] [PubMed]
  45. Patel, P.H.; Barbee, S.A.; Blankenship, J.T. GW-Bodies and P-bodies constitute two separate pools of sequestered Non-translating RNAs. PLoS ONE 2016, 11, e0150291. [Google Scholar] [CrossRef] [PubMed]
  46. Hebner, C.M.; Wilson, R.; Rader, J.; Bidder, M.; Laimins, L.A. Human papillomaviruses target the double-stranded RNA protein kinase pathway. J. Gen. Virol. 2006, 87, 3183–3193. [Google Scholar] [CrossRef] [PubMed]
  47. Gallois-Montbrun, S.; Kramer, B.; Swanson, C.M.; Byers, H.; Lynham, S.; Ward, M.; Malim, M.H. Antiviral protein APOBEC3G localizes to ribonucleoprotein complexes found in P bodies and stress granules. J. Virol. 2007, 81, 2165–2178. [Google Scholar] [CrossRef] [PubMed]
  48. Sen, G.L.; Blau, H.M. Argonaute 2/RISC resides in sites of mammalian mRNA decay known as cytoplasmic bodies. Nat. Cell Biol. 2005, 7, 633–636. [Google Scholar] [CrossRef] [PubMed]
  49. Chen, D.; Wilkinson, C.R.M.; Watt, S.; Penkett, C.J.; Toone, W.M.; Jones, N.; Bähler, J. Multiple pathways differentially regulate global oxidative stress responses in fission yeast. Mol. Biol. Cell 2008, 19, 308–317. [Google Scholar] [CrossRef] [PubMed]
  50. Ng, C.S.; Jogi, M.; Yoo, J.S.; Onomoto, K.; Koike, S.; Iwasaki, T.; Yoneyama, M.; Kato, H.; Fujita, T. Encephalomyocarditis virus disrupts stress granules, the critical platform for triggering antiviral innate immune responses. J. Virol. 2013, 87, 9511–9522. [Google Scholar] [CrossRef] [PubMed]
  51. Lai, M.C.; Lee, Y.H.; Tarn, W.Y. The DEAD-box RNA helicase DDX3 associates with export messenger ribonucleoproteins as well as tip-associated protein and participates in translational control. Mol. Biol. Cell 2008, 19, 3847–3858. [Google Scholar] [CrossRef] [PubMed]
  52. Kolobova, E.; Efimov, A.; Kaverina, I.; Rishi, A.K.; Schrader, J.W.; Ham, A.J.; Larocca, M.C.; Goldenring, J.R. Microtubule-dependent association of AKAP350A and CCAR1 with RNA stress granules. Exp. Cell Res. 2009, 315, 542–555. [Google Scholar] [CrossRef] [PubMed]
  53. Baguet, A.; Degot, S.; Cougot, N.; Bertrand, E.; Chenard, M.P.; Wendling, C.; Kessler, P.; Le Hir, H.; Rio, M.C.; Tomasetto, C. The exon-junction-complex-component metastatic lymph node 51 functions in stress-granule assembly. J. Cell Sci. 2007, 120, 2774–2784. [Google Scholar] [CrossRef] [PubMed]
  54. Pizzo, E.; Sarcinelli, C.; Sheng, J.; Fusco, S.; Formiggini, F.; Netti, P.; Yu, W.; D’Alessio, G.; Hu, G.-F. Ribonuclease/angiogenin inhibitor 1 regulates stress-induced subcellular localization of angiogenin to control growth and survival. J. Cell Sci. 2013, 126, 4308–4319. [Google Scholar] [CrossRef] [PubMed]
  55. Kawahara, H.; Imai, T.; Imataka, H.; Tsujimoto, M.; Matsumoto, K.; Okano, H. Neural RNA-binding protein Musashi1 inhibits translation initiation by competing with eIF4G for PABP. J. Cell Biol. 2008, 181, 639–653. [Google Scholar] [CrossRef] [PubMed]
  56. Nonhoff, U.; Ralser, M.; Welzel, F.; Piccini, I.; Balzereit, D.; Yaspo, M.L.; Lehrach, H.; Krobitsch, S. Ataxin-2 interacts with the DEAD/H-box RNA helicase DDX6 and interferes with P-bodies and stress granules. Mol. Biol. Cell 2007, 18, 1385–1396. [Google Scholar] [CrossRef] [PubMed]
  57. Swisher, K.D.; Parker, R. Localization to, and effects of Pbp1, Pbp4, Lsm12, Dhh1, and Pab1 on stress granules in Saccharomyces cerevisiae. PLoS ONE 2010, 5, e10006. [Google Scholar] [CrossRef] [PubMed]
  58. Takahara, T.; Maeda, T. Transient sequestration of TORC1 into stress granules during heat stress. Mol. Cell 2012, 47, 242–252. [Google Scholar] [CrossRef] [PubMed]
  59. Decca, M.B.; Carpio, M.A.; Bosc, C.; Galiano, M.R.; Job, D.; Andrieux, A.; Hallak, M.E. Post-translational arginylation of calreticulin: A new isospecies of calreticulin component of stress granules. J. Biol. Chem. 2007, 282, 8237–8245. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  60. Wehner, K.A.; Schutz, S.; Sarnow, P. OGFOD1, a novel modulator of eukaryotic translation initiation factor 2alpha phosphorylation and the cellular response to stress. Mol. Cell. Biol. 2010, 30, 2006–2016. [Google Scholar] [CrossRef] [PubMed]
  61. Solomon, S.; Xu, Y.; Wang, B.; David, M.D.; Schubert, P.; Kennedy, D.; Schrader, J.W. Distinct structural features of caprin-1 mediate its interaction with G3BP-1 and its induction of phosphorylation of eukaryotic translation initiation factor 2alpha, entry to cytoplasmic stress granules, and selective interaction with a subset of mRNAs. Mol. Cell. Biol. 2007, 27, 2324–2342. [Google Scholar] [CrossRef] [PubMed]
  62. Nousch, M.; Reed, V.; Bryson-Richardson, R.J.; Currie, P.D.; Preiss, T. The eIF4G-homolog p97 can activate translation independent of caspase cleavage. RNA 2007, 13, 374–384. [Google Scholar] [CrossRef] [PubMed]
  63. De Leeuw, F.; Zhang, T.; Wauquier, C.; Huez, G.; Kruys, V.; Gueydan, C. The cold-inducible RNA-binding protein migrates from the nucleus to cytoplasmic stress granules by a methylation-dependent mechanism and acts as a translational repressor. Exp. Cell Res. 2007, 313, 4130–4144. [Google Scholar] [CrossRef] [PubMed]
  64. Ohn, T.; Kedersha, N.; Hickman, T.; Tisdale, S.; Anderson, P. A functional RNAi screen links O-GlcNAc modification of ribosomal proteins to stress granule and processing body assembly. Nat. Cell Biol. 2008, 10, 1224–1231. [Google Scholar] [CrossRef] [PubMed]
  65. Fujimura, K.; Kano, F.; Murata, M. Dual localization of the RNA binding protein CUGBP-1 to stress granule and perinucleolar compartment. Exp. Cell Res. 2008, 314, 543–553. [Google Scholar] [CrossRef] [PubMed]
  66. Hofmann, I.; Casella, M.; Schnolzer, M.; Schlechter, T.; Spring, H.; Franke, W.W. Identification of the junctional plaque protein plakophilin 3 in cytoplasmic particles containing RNA-binding proteins and the recruitment of plakophilins 1 and 3 to stress granules. Mol. Biol. Cell 2006, 17, 1388–1398. [Google Scholar] [CrossRef] [PubMed]
  67. Onishi, H.; Kino, Y.; Morita, T.; Futai, E.; Sasagawa, N.; Ishiura, S. MBNL1 associates with YB-1 in cytoplasmic stress granules. J. Neurosci. Res. 2008, 86, 1994–2002. [Google Scholar] [CrossRef] [PubMed]
  68. Loschi, M.; Leishman, C.C.; Berardone, N.; Boccaccio, G.L. Dynein and kinesin regulate stress-granule and P-body dynamics. J. Cell Sci. 2009, 122, 3973–3982. [Google Scholar] [CrossRef] [PubMed]
  69. Kim, J.E.; Ryu, I.; Kim, W.J.; Song, O.K.; Ryu, J.; Kwon, M.Y.; Kim, J.H.; Jang, S.K. Proline-rich transcript in brain protein induces stress granule formation. Mol. Cell. Biol. 2008, 28, 803–813. [Google Scholar] [CrossRef] [PubMed]
  70. Ogawa, F.; Kasai, M.; Akiyama, T. A functional link between Disrupted-In-Schizophrenia 1 and the eukaryotic translation initiation factor 3. Boilchem. Biophys. Res. Commun. 2005, 338, 771–776. [Google Scholar] [CrossRef] [PubMed]
  71. Morris, A.R.; Mukherjee, N.; Keene, J.D. Ribonomic analysis of human Pum1 reveals cis-trans conservation across species despite evolution of diverse mRNA target sets. Mol. Cell. Biol. 2008, 28, 4093–4103. [Google Scholar] [CrossRef] [PubMed]
  72. Kimball, S.R.; Horetsky, R.L.; Ron, D.; Jefferson, L.S.; Harding, H.P. Mammalian stress granules represent sites of accumulation of stalled translation initiation complexes. Am. J. Physiol. Cell Physiol. 2003, 284, C273–C284. [Google Scholar] [CrossRef] [PubMed]
  73. Vessey, J.P.; Vaccani, A.; Xie, Y.; Dahm, R.; Karra, D.; Kiebler, M.A.; Macchi, P. Dendritic localization of the translational repressor Pumilio 2 and its contribution to dendritic stress granules. J. Neurosci. Off. J. Soc. Neurosci. 2006, 26, 6496–6508. [Google Scholar] [CrossRef] [PubMed]
  74. Zurla, C.; Lifland, A.W.; Santangelo, P.J. Characterizing mRNA interactions with RNA granules during translation initiation inhibition. PLoS ONE 2011, 6, e19727. [Google Scholar] [CrossRef] [PubMed]
  75. Arimoto, K.; Fukuda, H.; Imajoh-Ohmi, S.; Saito, H.; Takekawa, M. Formation of stress granules inhibits apoptosis by suppressing stress-responsive MAPK pathways. Nat. Cell. Biol. 2008, 10, 1324–1332. [Google Scholar] [CrossRef] [PubMed]
  76. Fukuda, T.; Naiki, T.; Saito, M.; Irie, K. hnRNP K interacts with RNA binding motif protein 42 and functions in the maintenance of cellular ATP level during stress conditions. Genes Cells Devot. Mol. Cell. Mech. 2009, 14, 113–128. [Google Scholar] [CrossRef] [PubMed]
  77. Li, C.H.; Ohn, T.; Ivanov, P.; Tisdale, S.; Anderson, P. eIF5A promotes translation elongation, polysome disassembly and stress granule assembly. PLoS ONE 2010, 5, e9942. [Google Scholar] [CrossRef] [PubMed]
  78. Tsai, N.P.; Ho, P.C.; Wei, L.N. Regulation of stress granule dynamics by Grb7 and FAK signalling pathway. EMBO J. 2008, 27, 715–726. [Google Scholar] [CrossRef] [PubMed]
  79. Hua, Y.; Zhou, J. Rpp20 interacts with SMN and is re-distributed into SMN granules in response to stress. Boilchem. Biophys. Res. Commun. 2004, 314, 268–276. [Google Scholar] [CrossRef]
  80. Rothe, F.; Gueydan, C.; Bellefroid, E.; Huez, G.; Kruys, V. Identification of FUSE-binding proteins as interacting partners of TIA proteins. Boilchem. Biophys. Res. Commun. 2006, 343, 57–68. [Google Scholar] [CrossRef] [PubMed]
  81. Eisinger-Mathason, T.S.; Andrade, J.; Groehler, A.L.; Clark, D.E.; Muratore-Schroeder, T.L.; Pasic, L.; Smith, J.A.; Shabanowitz, J.; Hunt, D.F.; Macara, I.G.; et al. Codependent functions of RSK2 and the apoptosis-promoting factor TIA-1 in stress granule assembly and cell survival. Mol. Cell 2008, 31, 722–736. [Google Scholar] [CrossRef] [PubMed]
  82. Andersson, M.K.; Stahlberg, A.; Arvidsson, Y.; Olofsson, A.; Semb, H.; Stenman, G.; Nilsson, O.; Aman, P. The multifunctional FUS, EWS and TAF15 proto-oncoproteins show cell type-specific expression patterns and involvement in cell spreading and stress response. BMC Cell Biol. 2008, 9, 37. [Google Scholar] [CrossRef] [PubMed]
  83. Henao-Mejia, J.; He, J.J. Sam68 relocalization into stress granules in response to oxidative stress through complexing with TIA-1. Exp. Cell Res. 2009, 315, 3381–3395. [Google Scholar] [CrossRef] [PubMed]
  84. Goulet, I.; Boisvenue, S.; Mokas, S.; Mazroui, R.; Cote, J. TDRD3, a novel Tudor domain-containing protein, localizes to cytoplasmic stress granules. Hum. Mol. Genet. 2008, 17, 3055–3074. [Google Scholar] [CrossRef] [PubMed]
  85. Zhu, C.H.; Kim, J.; Shay, J.W.; Wright, W.E. SGNP: An essential Stress Granule/Nucleolar Protein potentially involved in 5.8s rRNA processing/transport. PloS ONE 2008, 3, e3716. [Google Scholar] [CrossRef] [PubMed]
  86. Kobayashi, T.; Winslow, S.; Sunesson, L.; Hellman, U.; Larsson, C. PKCalpha binds G3BP2 and regulates stress granule formation following cellular stress. PloS ONE 2012, 7, e35820. [Google Scholar] [CrossRef] [PubMed]
  87. Colombrita, C.; Zennaro, E.; Fallini, C.; Weber, M.; Sommacal, A.; Buratti, E.; Silani, V.; Ratti, A. TDP-43 is recruited to stress granules in conditions of oxidative insult. J. Neurochem. 2009, 111, 1051–1061. [Google Scholar] [CrossRef] [PubMed]
  88. Kwon, S.; Zhang, Y.; Matthias, P. The deacetylase HDAC6 is a novel critical component of stress granules involved in the stress response. Genes Dev. 2007, 21, 3381–3394. [Google Scholar] [CrossRef] [PubMed]
  89. Guil, S.; Long, J.C.; Caceres, J.F. hnRNP A1 relocalization to the stress granules reflects a role in the stress response. Mol. Cell. Biol. 2006, 26, 5744–5758. [Google Scholar] [CrossRef] [PubMed]
  90. Martina, J.A.; Diab, H.I.; Brady, O.A.; Puertollano, R. TFEB and TFE3 are novel components of the integrated stress response. EMBO J. 2016, 35, 479–495. [Google Scholar] [CrossRef] [PubMed]
  91. Gilks, N.; Kedersha, N.; Ayodele, M.; Shen, L.; Stoecklin, G.; Dember, L.M.; Anderson, P. Stress granule assembly is mediated by prion-like aggregation of TIA-1. Mol. Biol. Cell 2004, 15, 5383–5398. [Google Scholar] [CrossRef] [PubMed]
  92. Kim, W.J.; Back, S.H.; Kim, V.; Ryu, I.; Jang, S.K. Sequestration of TRAF2 into stress granules interrupts tumor necrosis factor signaling under stress conditions. Mol. Cell. Biol. 2005, 25, 2450–2462. [Google Scholar] [CrossRef] [PubMed]
  93. Burry, R.W.; Smith, C.L. HuD distribution changes in response to heat shock but not neurotrophic stimulation. J. Histochem. Cytochem. Off. J. Histochem. Soc. 2006, 54, 1129–1138. [Google Scholar] [CrossRef] [PubMed]
  94. Simpson-Holley, M.; Kedersha, N.; Dower, K.; Rubins, K.H.; Anderson, P.; Hensley, L.E.; Connor, J.H. Formation of antiviral cytoplasmic granules during orthopoxvirus infection. J. Virol. 2011, 85, 1581–1593. [Google Scholar] [CrossRef] [PubMed]
  95. Gallouzi, I.E.; Brennan, C.M.; Stenberg, M.G.; Swanson, M.S.; Eversole, A.; Maizels, N.; Steitz, J.A. HuR binding to cytoplasmic mRNA is perturbed by heat shock. Proc. Natl. Acad. Sci. USA 2000, 97, 3073–3078. [Google Scholar] [CrossRef] [PubMed]
  96. Chang, Y.W.; Huang, Y.S. Arsenite-activated JNK signaling enhances CPEB4-Vinexin interaction to Facilitate stress granule assembly and cell survival. PloS ONE 2014, 9, e107961. [Google Scholar] [CrossRef] [PubMed]
  97. Rothenburg, S.; Deigendesch, N.; Dittmar, K.; Koch-Nolte, F.; Haag, F.; Lowenhaupt, K.; Rich, A. A PKR-like eukaryotic initiation factor 2alpha kinase from zebrafish contains Z-DNA binding domains instead of dsRNA binding domains. Proc. Natl. Acad. Sci. USA 2005, 102, 1602–1607. [Google Scholar] [CrossRef] [PubMed]
  98. Brehm, M.A.; Schenk, T.M.; Zhou, X.; Fanick, W.; Lin, H.; Windhorst, S.; Nalaskowski, M.M.; Kobras, M.; Shears, S.B.; Mayr, G.W. Intracellular localization of human Ins(1,3,4,5,6)P5 2-kinase. Boilchem. J. 2007, 408, 335–345. [Google Scholar]
  99. Goodier, J.L.; Zhang, L.; Vetter, M.R.; Kazazian, H.H., Jr. LINE-1 ORF1 protein localizes in stress granules with other RNA-binding proteins, including components of RNA interference RNA-induced silencing complex. Mol. Cell. Biol. 2007, 27, 6469–6483. [Google Scholar] [CrossRef] [PubMed]
  100. Sheth, U.; Parker, R. Decapping and decay of messenger RNA occur in cytoplasmic processing bodies. Science 2003, 300, 805–808. [Google Scholar] [CrossRef] [PubMed]
  101. Savas, J.N.; Makusky, A.; Ottosen, S.; Baillat, D.; Then, F.; Krainc, D.; Shiekhattar, R.; Markey, S.P.; Tanese, N. Huntington’s disease protein contributes to RNA-mediated gene silencing through association with Argonaute and P bodies. Proc. Natl. Acad. Sci. USA 2008, 105, 10820–10825. [Google Scholar] [CrossRef] [PubMed]
  102. Luke, B.; Azzalin, C.M.; Hug, N.; Deplazes, A.; Peter, M.; Lingner, J. Saccharomyces cerevisiae Ebs1p is a putative ortholog of human Smg7 and promotes nonsense-mediated mRNA decay. Nucleic Acids Res. 2007, 35, 7688–7697. [Google Scholar] [CrossRef] [PubMed]
  103. Zheng, D.; Ezzeddine, N.; Chen, C.Y.; Zhu, W.; He, X.; Shyu, A.B. Deadenylation is prerequisite for P-body formation and mRNA decay in mammalian cells. J. Cell Biol. 2008, 182, 89–101. [Google Scholar] [CrossRef] [PubMed]
  104. Lippincott-Schwartz, J.; Patterson, G.H. Development and use of fluorescent protein markers in living cells. Science 2003, 300, 87–91. [Google Scholar] [CrossRef] [PubMed]
  105. Kshirsagar, M.; Parker, R. Identification of Edc3p as an enhancer of mRNA decapping in Saccharomyces cerevisiae. Genetics 2004, 166, 729–739. [Google Scholar] [CrossRef] [PubMed]
  106. Teixeira, D.; Parker, R. Analysis of P-body assembly in Saccharomyces cerevisiae. Mol. Biol. Cell 2007, 18, 2274–2287. [Google Scholar] [CrossRef] [PubMed]
  107. Ferraiuolo, M.A.; Basak, S.; Dostie, J.; Murray, E.L.; Schoenberg, D.R.; Sonenberg, N. A role for the eIF4E-binding protein 4E-T in P-body formation and mRNA decay. J. Cell Biol. 2005, 170, 913–924. [Google Scholar] [CrossRef] [PubMed]
  108. Matsumoto, K.; Nakayama, H.; Yoshimura, M.; Masuda, A.; Dohmae, N.; Matsumoto, S.; Tsujimoto, M. PRMT1 is required for RAP55 to localize to processing bodies. RNA Biol. 2012, 9, 610–623. [Google Scholar] [CrossRef] [PubMed]
  109. Meister, G.; Landthaler, M.; Peters, L.; Chen, P.Y.; Urlaub, H.; Luhrmann, R.; Tuschl, T. Identification of novel argonaute-associated proteins. Curr. Biol. CB 2005, 15, 2149–2155. [Google Scholar] [CrossRef] [PubMed]
  110. Baillat, D.; Shiekhattar, R. Functional dissection of the human TNRC6 (GW182-related) family of proteins. Mol. Cell. Biol. 2009, 29, 4144–4155. [Google Scholar] [CrossRef] [PubMed]
  111. Buchet-Poyau, K.; Courchet, J.; Le Hir, H.; Seraphin, B.; Scoazec, J.Y.; Duret, L.; Domon-Dell, C.; Freund, J.N.; Billaud, M. Identification and characterization of human Mex-3 proteins, a novel family of evolutionarily conserved RNA-binding proteins differentially localized to processing bodies. Nucleic Acids Res. 2007, 35, 1289–1300. [Google Scholar] [CrossRef] [PubMed]
  112. Katahira, J.; Miki, T.; Takano, K.; Maruhashi, M.; Uchikawa, M.; Tachibana, T.; Yoneda, Y. Nuclear RNA export factor 7 is localized in processing bodies and neuronal RNA granules through interactions with shuttling hnRNPs. Nucleic Acids Res. 2008, 36, 616–628. [Google Scholar] [CrossRef] [PubMed]
  113. Leung, A.K.; Calabrese, J.M.; Sharp, P.A. Quantitative analysis of Argonaute protein reveals microRNA-dependent localization to stress granules. Proc. Natl. Acad. Sci. USA 2006, 103, 18125–18130. [Google Scholar] [CrossRef] [PubMed]
  114. Hoyle, N.P.; Castelli, L.M.; Campbell, S.G.; Holmes, L.E.; Ashe, M.P. Stress-dependent relocalization of translationally primed mRNPs to cytoplasmic granules that are kinetically and spatially distinct from P-bodies. J. Cell Biol. 2007, 179, 65–74. [Google Scholar] [CrossRef] [PubMed]
  115. Kozak, S.L.; Marin, M.; Rose, K.M.; Bystrom, C.; Kabat, D. The anti-HIV-1 editing enzyme APOBEC3G binds HIV-1 RNA and messenger RNAs that shuttle between polysomes and stress granules. J. Biol. Chem. 2006, 281, 29105–29119. [Google Scholar] [CrossRef] [PubMed]
  116. Fujimura, K.; Kano, F.; Murata, M. Identification of PCBP2, a facilitator of IRES-mediated translation, as a novel constituent of stress granules and processing bodies. Rna 2008, 14, 425–431. [Google Scholar] [CrossRef] [PubMed]
  117. Kedersha, N.; Stoecklin, G.; Ayodele, M.; Yacono, P.; Lykke-Andersen, J.; Fitzler, M.J.; Scheuner, D.; Kaufman, R.J.; Golan, D.E.; Anderson, P. Stress granules and processing bodies are dynamically linked sites of mRNP remodeling. J. Cell Biol. 2005, 169, 871–884. [Google Scholar] [CrossRef] [PubMed]
  118. Yang, W.-H.; Yu, J.H.; Gulick, T.; Bloch, K.D.; Bloch, D.B. RNA-associated protein 55 (RAP55) localizes to mRNA processing bodies and stress granules. RNA (New York, NY) 2006, 12, 547–554. [Google Scholar] [CrossRef] [PubMed]
  119. Athanasopoulos, V.; Barker, A.; Yu, D.; Tan, A.H.; Srivastava, M.; Contreras, N.; Wang, J.; Lam, K.P.; Brown, S.H.; Goodnow, C.C.; et al. The ROQUIN family of proteins localizes to stress granules via the ROQ domain and binds target mRNAs. FEBS J. 2010, 277, 2109–2127. [Google Scholar] [CrossRef] [PubMed]
  120. Eulalio, A.; Behm-Ansmant, I.; Izaurralde, E. P bodies: At the crossroads of post-transcriptional pathways. Nat. Rev. Mol. Cell Biol. 2007, 8, 9–22. [Google Scholar] [CrossRef] [PubMed]
  121. Baez, M.V.; Boccaccio, G.L. Mammalian smaug is a translational repressor that forms cytoplasmic foci similar to stress granules. J. Biol. Chem. 2005, 280, 43131–43140. [Google Scholar] [CrossRef] [PubMed]
  122. Buchan, J.R.; Muhlrad, D.; Parker, R. P bodies promote stress granule assembly in Saccharomyces cerevisiae. J. Cell Biol. 2008, 183, 441–455. [Google Scholar] [CrossRef] [PubMed]
  123. Courchet, J.; Buchet-Poyau, K.; Potemski, A.; Bres, A.; Jariel-Encontre, I.; Billaud, M. Interaction with 14–3–3 adaptors regulates the sorting of hMex-3B RNA-binding protein to distinct classes of RNA granules. J. Biol. Chem. 2008, 283, 32131–32142. [Google Scholar] [CrossRef] [PubMed]
  124. Quaresma, A.J.; Bressan, G.C.; Gava, L.M.; Lanza, D.C.; Ramos, C.H.; Kobarg, J. Human hnRNP Q re-localizes to cytoplasmic granules upon PMA, thapsigargin, arsenite and heat-shock treatments. Exp. Cell Res. 2009, 315, 968–980. [Google Scholar] [CrossRef] [PubMed]
  125. Weinmann, L.; Hock, J.; Ivacevic, T.; Ohrt, T.; Mutze, J.; Schwille, P.; Kremmer, E.; Benes, V.; Urlaub, H.; Meister, G. Importin 8 is a gene silencing factor that targets argonaute proteins to distinct mRNAs. Cell 2009, 136, 496–507. [Google Scholar] [CrossRef] [PubMed]
  126. Yang, W.-H.; Bloch, D.B. Probing the mRNA processing body using protein macroarrays and ”autoantigenomics”. RNA (New York, NY) 2007, 13, 704–712. [Google Scholar] [CrossRef] [PubMed]
  127. Wasserman, T.; Katsenelson, K.; Daniliuc, S.; Hasin, T.; Choder, M.; Aronheim, A. A Novel c-Jun N-terminal Kinase (JNK)-binding Protein WDR62 Is Recruited to Stress Granules and Mediates a Nonclassical JNK Activation. Mol. Biol. Cell 2010, 21, 117–130. [Google Scholar] [CrossRef] [PubMed]
  128. Balzer, E.; Moss, E.G. Localization of the developmental timing regulator Lin28 to mRNP complexes, P-bodies and stress granules. RNA Biol. 2007, 4, 16–25. [Google Scholar] [CrossRef] [PubMed]
  129. Sciortino, M.T.; Parisi, T.; Siracusano, G.; Mastino, A.; Taddeo, B.; Roizman, B. The virion host shutoff RNase plays a key role in blocking the activation of protein kinase R in cells infected with herpes simplex virus 1. J. Virol. 2013, 87, 3271–3276. [Google Scholar] [CrossRef] [PubMed]
  130. Cassady, K.A.; Gross, M. The herpes simplex virus type 1 U(S)11 protein interacts with protein kinase R in infected cells and requires a 30-amino-acid sequence adjacent to a kinase substrate domain. J. Virol. 2002, 76, 2029–2035. [Google Scholar] [CrossRef] [PubMed]
  131. He, B.; Gross, M.; Roizman, B. The gamma(1)34.5 protein of herpes simplex virus 1 complexes with protein phosphatase 1alpha to dephosphorylate the alpha subunit of the eukaryotic translation initiation factor 2 and preclude the shutoff of protein synthesis by double-stranded RNA-activated protein kinase. Proc. Natl. Acad. Sci. USA 1997, 94, 843–848. [Google Scholar] [PubMed]
  132. Mulvey, M.; Arias, C.; Mohr, I. Maintenance of endoplasmic reticulum (ER) homeostasis in herpes simplex virus type 1-infected cells through the association of a viral glycoprotein with PERK, a cellular ER stress sensor. J. Virol. 2007, 81, 3377–3390. [Google Scholar] [CrossRef] [PubMed]
  133. Esclatine, A.; Taddeo, B.; Roizman, B. Herpes simplex virus 1 induces cytoplasmic accumulation of TIA-1/TIAR and both synthesis and cytoplasmic accumulation of tristetraprolin, two cellular proteins that bind and destabilize AU-rich RNAs. J. Virol. 2004, 78, 8582–8592. [Google Scholar] [CrossRef] [PubMed]
  134. Dauber, B.; Pelletier, J.; Smiley, J.R. The herpes simplex virus 1 vhs protein enhances translation of viral true late mRNAs and virus production in a cell type-dependent manner. J. Virol. 2011, 85, 5363–5373. [Google Scholar] [CrossRef] [PubMed]
  135. Dauber, B.; Poon, D.; Dos Santos, T.; Duguay, B.A.; Mehta, N.; Saffran, H.A.; Smiley, J.R. The herpes simplex virus virion host shutoff protein enhances translation of viral true late mrnas independently of suppressing protein kinase r and stress granule formation. J. Virol. 2016, 90, 6049–6057. [Google Scholar] [CrossRef] [PubMed]
  136. Finnen, R.L.; Pangka, K.R.; Banfield, B.W. Herpes simplex virus 2 infection impacts stress granule accumulation. J. Virol. 2012, 86, 8119–8130. [Google Scholar] [CrossRef] [PubMed]
  137. Finnen, R.L.; Hay, T.J.; Dauber, B.; Smiley, J.R.; Banfield, B.W. The herpes simplex virus 2 virion-associated ribonuclease vhs interferes with stress granule formation. J. Virol. 2014, 88, 12727–12739. [Google Scholar] [CrossRef] [PubMed]
  138. Isler, J.A.; Skalet, A.H.; Alwine, J.C. Human cytomegalovirus infection activates and regulates the unfolded protein response. J. Virol. 2005, 79, 6890–6899. [Google Scholar] [CrossRef] [PubMed]
  139. Isler, J.A.; Maguire, T.G.; Alwine, J.C. Production of infectious human cytomegalovirus virions is inhibited by drugs that disrupt calcium homeostasis in the endoplasmic reticulum. J. Virol. 2005, 79, 15388–15397. [Google Scholar] [CrossRef] [PubMed]
  140. Katsafanas, G.C.; Moss, B. Vaccinia virus intermediate stage transcription is complemented by Ras-GTPase-activating protein SH3 domain-binding protein (G3BP) and cytoplasmic activation/proliferation-associated protein (p137) individually or as a heterodimer. J. Biol. Chem. 2004, 279, 52210–52217. [Google Scholar] [CrossRef] [PubMed]
  141. Katsafanas, G.C.; Moss, B. Colocalization of transcription and translation within cytoplasmic poxvirus factories coordinates viral expression and subjugates host functions. Cell Host Microbe 2007, 2, 221–228. [Google Scholar] [CrossRef] [PubMed]
  142. Zaborowska, I.; Kellner, K.; Henry, M.; Meleady, P.; Walsh, D. Recruitment of host translation initiation factor eIF4G by the Vaccinia Virus ssDNA-binding protein I3. Virology 2012, 425, 11–22. [Google Scholar] [CrossRef] [PubMed]
  143. Walsh, D.; Arias, C.; Perez, C.; Halladin, D.; Escandon, M.; Ueda, T.; Watanabe-Fukunaga, R.; Fukunaga, R.; Mohr, I. Eukaryotic translation initiation factor 4F architectural alterations accompany translation initiation factor redistribution in poxvirus-infected cells. Mol. Cell. Biol. 2008, 28, 2648–2658. [Google Scholar] [CrossRef] [PubMed]
  144. Montero, H.; Rojas, M.; Arias, C.F.; Lopez, S. Rotavirus infection induces the phosphorylation of eIF2alpha but prevents the formation of stress granules. J. Virol. 2008, 82, 1496–1504. [Google Scholar] [CrossRef] [PubMed]
  145. Rojas, M.; Arias, C.F.; Lopez, S. Protein kinase R is responsible for the phosphorylation of eIF2alpha in rotavirus infection. J. Virol. 2010, 84, 10457–10466. [Google Scholar] [CrossRef] [PubMed]
  146. Qin, Q.; Hastings, C.; Miller, C.L. Mammalian orthoreovirus particles induce and are recruited into stress granules at early times postinfection. J. Virol. 2009, 83, 11090–11101. [Google Scholar] [CrossRef] [PubMed]
  147. Qin, Q.; Carroll, K.; Hastings, C.; Miller, C.L. Mammalian orthoreovirus escape from host translational shutoff correlates with stress granule disruption and is independent of eIF2alpha phosphorylation and PKR. J. Virol. 2011, 85, 8798–8810. [Google Scholar] [CrossRef] [PubMed]
  148. Carroll, K.; Hastings, C.; Miller, C.L. Amino acids 78 and 79 of Mammalian Orthoreovirus protein microNS are necessary for stress granule localization, core protein lambda2 interaction, and de novo virus replication. Virology 2014, 448, 133–145. [Google Scholar] [CrossRef] [PubMed]
  149. Dougherty, J.D.; Tsai, W.C.; Lloyd, R.E. Multiple poliovirus proteins repress cytoplasmic RNA granules. Viruses 2015, 7, 6127–6140. [Google Scholar] [CrossRef] [PubMed]
  150. White, J.P.; Cardenas, A.M.; Marissen, W.E.; Lloyd, R.E. Inhibition of cytoplasmic mRNA stress granule formation by a viral proteinase. Cell Host Microbe 2007, 2, 295–305. [Google Scholar] [CrossRef] [PubMed]
  151. Piotrowska, J.; Hansen, S.J.; Park, N.; Jamka, K.; Sarnow, P.; Gustin, K.E. Stable formation of compositionally unique stress granules in virus-infected cells. J. Virol. 2010, 84, 3654–3665. [Google Scholar] [CrossRef] [PubMed]
  152. White, J.P.; Lloyd, R.E. Poliovirus unlinks TIA1 aggregation and mRNA stress granule formation. J. Virol. 2011, 85, 12442–12454. [Google Scholar] [CrossRef] [PubMed]
  153. Fung, G.; Ng, C.S.; Zhang, J.; Shi, J.; Wong, J.; Piesik, P.; Han, L.; Chu, F.; Jagdeo, J.; Jan, E.; et al. Production of a dominant-negative fragment due to G3BP1 cleavage contributes to the disruption of mitochondria-associated protective stress granules during CVB3 infection. PLoS ONE 2013, 8, e79546. [Google Scholar] [CrossRef] [PubMed]
  154. Borghese, F.; Michiels, T. The leader protein of cardioviruses inhibits stress granule assembly. J. Virol. 2011, 85, 9614–9622. [Google Scholar] [CrossRef] [PubMed]
  155. Langereis, M.A.; Feng, Q.; van Kuppeveld, F.J. MDA5 localizes to stress granules, but this localization is not required for the induction of type I interferon. J. Virol. 2013, 87, 6314–6325. [Google Scholar] [CrossRef] [PubMed]
  156. Reineke, L.C.; Kedersha, N.; Langereis, M.A.; van Kuppeveld, F.J.; Lloyd, R.E. Stress granules regulate double-stranded RNA-dependent protein kinase activation through a complex containing G3BP1 and Caprin1. mBio 2015, 6, e02486. [Google Scholar] [CrossRef] [PubMed]
  157. McInerney, G.; Kedersha, N.; Kaufman, R.J.; Anderson, P.; Liljestrom, P. Importance of eIF2a Phosphorylation and Stress Granule Assembly in Alphavirus Translation Regulation. Mol. Biol. Cell 2005, 16, 3753–3763. [Google Scholar] [CrossRef] [PubMed]
  158. Panas, M.D.; Varjak, M.; Lulla, A.; Eng, K.E.; Merits, A.; Karlsson Hedestam, G.B.; McInerney, G.M. Sequestration of G3BP coupled with efficient translation inhibits stress granules in Semliki Forest virus infection. Mol. Biol. Cell 2012, 23, 4701–4712. [Google Scholar] [CrossRef] [PubMed]
  159. Fros, J.J.; Domeradzka, N.E.; Baggen, J.; Geertsema, C.; Flipse, J.; Vlak, J.M.; Pijlman, G.P. Chikungunya virus nsP3 blocks stress granule assembly by recruitment of G3BP into cytoplasmic foci. J. Virol. 2012, 86, 10873–10879. [Google Scholar] [CrossRef] [PubMed]
  160. Matsuki, H.; Takahashi, M.; Higuchi, M.; Makokha, G.N.; Oie, M.; Fujii, M. Both G3BP1 and G3BP2 contribute to stress granule formation. Genes Cells Devot. Mol. Cell. Mech. 2013, 18, 135–146. [Google Scholar] [CrossRef] [PubMed]
  161. Scholte, F.E.; Tas, A.; Albulescu, I.C.; Zusinaite, E.; Merits, A.; Snijder, E.J.; van Hemert, M.J. Stress granule components G3BP1 and G3BP2 play a proviral role early in chikungunya virus replication. J. Virol. 2015, 89, 4457–4469. [Google Scholar] [CrossRef] [PubMed]
  162. Matthews, J.D.; Frey, T.K. Analysis of subcellular G3BP redistribution during rubella virus infection. J. Gen. Virol. 2012, 93, 267–274. [Google Scholar] [CrossRef] [PubMed]
  163. Gorchakov, R.; Garmashova, N.; Frolova, E.; Frolov, I. Different types of nsP3-containing protein complexes in Sindbis virus-infected cells. J. Virol. 2008, 82, 10088–10101. [Google Scholar] [CrossRef] [PubMed]
  164. Frolova, E.; Gorchakov, R.; Garmashova, N.; Atasheva, S.; Vergara, L.A.; Frolov, I. Formation of nsP3-specific protein complexes during Sindbis virus replication. J. Virol. 2006, 80, 4122–4134. [Google Scholar] [CrossRef] [PubMed]
  165. Cristea, I.M.; Rozjabek, H.; Molloy, K.R.; Karki, S.; White, L.L.; Rice, C.M.; Rout, M.P.; Chait, B.T.; MacDonald, M.R. Host factors associated with the Sindbis virus RNA-dependent RNA polymerase: Role for G3BP1 and G3BP2 in virus replication. J. Virol. 2010, 84, 6720–6732. [Google Scholar] [CrossRef] [PubMed]
  166. Venticinque, L.; Meruelo, D. Sindbis viral vector induced apoptosis requires translational inhibition and signaling through Mcl-1 and Bak. Mol. Cancer 2010, 9, 37. [Google Scholar] [CrossRef] [PubMed]
  167. Li, W.; Li, Y.; Kedersha, N.; Anderson, P.; Emara, M.; Swiderek, K.M.; Moreno, G.T.; Brinton, M.A. Cell proteins TIA-1 and TIAR interact with the 3′ stem-loop of the West Nile virus complementary minus-strand RNA and facilitate virus replication. J. Virol. 2002, 76, 11989–12000. [Google Scholar] [CrossRef] [PubMed]
  168. Emara, M.M.; Brinton, M.A. Interaction of TIA-1/TIAR with West Nile and dengue virus products in infected cells interferes with stress granule formation and processing body assembly. Proc. Natl. Acad. Sci. USA 2007, 104, 9041–9046. [Google Scholar] [CrossRef] [PubMed]
  169. Courtney, S.C.; Scherbik, S.V.; Stockman, B.M.; Brinton, M.A. West nile virus infections suppress early viral RNA synthesis and avoid inducing the cell stress granule response. J. Virol. 2012, 86, 3647–3657. [Google Scholar] [CrossRef] [PubMed]
  170. Xia, J.; Chen, X.; Xu, F.; Wang, Y.; Shi, Y.; Li, Y.; He, J.; Zhang, P. Dengue virus infection induces formation of G3BP1 granules in human lung epithelial cells. Arch. Virol. 2015, 160, 2991–2999. [Google Scholar] [CrossRef] [PubMed]
  171. Ward, A.M.; Bidet, K.; Yinglin, A.; Ler, S.G.; Hogue, K.; Blackstock, W.; Gunaratne, J.; Garcia-Blanco, M.A. Quantitative mass spectrometry of DENV-2 RNA-interacting proteins reveals that the DEAD-box RNA helicase DDX6 binds the DB1 and DB2 3′ UTR structures. RNA Biol. 2011, 8, 1173–1186. [Google Scholar] [CrossRef] [PubMed]
  172. Bidet, K.; Dadlani, D.; Garcia-Blanco, M.A. G3BP1, G3BP2 and CAPRIN1 are required for translation of interferon stimulated mRNAs and are targeted by a dengue virus non-coding RNA. PLoS Pathog. 2014, 10, e1004242. [Google Scholar] [CrossRef] [PubMed]
  173. Albornoz, A.; Carletti, T.; Corazza, G.; Marcello, A. The stress granule component TIA-1 binds tick-borne encephalitis virus RNA and is recruited to perinuclear sites of viral replication to inhibit viral translation. J. Virol. 2014, 88, 6611–6622. [Google Scholar] [CrossRef] [PubMed]
  174. Katoh, H.; Okamoto, T.; Fukuhara, T.; Kambara, H.; Morita, E.; Mori, Y.; Kamitani, W.; Matsuura, Y. Japanese encephalitis virus core protein inhibits stress granule formation through an interaction with Caprin-1 and facilitates viral propagation. J. Virol. 2013, 87, 489–502. [Google Scholar] [CrossRef] [PubMed]
  175. Jefferson, M.; Donaszi-Ivanov, A.; Pollen, S.; Dalmay, T.; Saalbach, G.; Powell, P.P. Host factors that interact with the pestivirus N-terminal protease, Npro, are components of the ribonucleoprotein complex. J. Virol. 2014, 88, 10340–10353. [Google Scholar] [CrossRef] [PubMed]
  176. Garaigorta, U.; Heim, M.H.; Boyd, B.; Wieland, S.; Chisari, F.V. Hepatitis C virus (HCV) induces formation of stress granules whose proteins regulate HCV RNA replication and virus assembly and egress. J. Virol. 2012, 86, 11043–11056. [Google Scholar] [CrossRef] [PubMed]
  177. Ruggieri, A.; Dazert, E.; Metz, P.; Hofmann, S.; Bergeest, J.P.; Mazur, J.; Bankhead, P.; Hiet, M.S.; Kallis, S.; Alvisi, G.; et al. Dynamic oscillation of translation and stress granule formation mark the cellular response to virus infection. Cell Host Microbe 2012, 12, 71–85. [Google Scholar] [CrossRef] [PubMed]
  178. Li, Q.; Pene, V.; Krishnamurthy, S.; Cha, H.; Liang, T.J. Hepatitis C virus infection activates an innate pathway involving IKK-alpha in lipogenesis and viral assembly. Nat. Med. 2013, 19, 722–729. [Google Scholar] [CrossRef] [PubMed]
  179. Pene, V.; Li, Q.; Sodroski, C.; Hsu, C.S.; Liang, T.J. Dynamic interaction of stress granule, DDX3X and IKK-alpha mediates multiple functions in hepatitis C virus infection. J. Virol. 2015, 89, 5462–5477. [Google Scholar] [CrossRef] [PubMed]
  180. Valiente-Echeverría, F.; Hermoso, M.A.; Soto-Rifo, R. RNA helicase DDX3: At the crossroad of viral replication and antiviral immunity. Rev. Med. Virol. 2015, 25, 286–299. [Google Scholar] [CrossRef] [PubMed]
  181. Ariumi, Y.; Kuroki, M.; Kushima, Y.; Osugi, K.; Hijikata, M.; Maki, M.; Ikeda, M.; Kato, N. Hepatitis C virus hijacks P-body and stress granule components around lipid droplets. J. Virol. 2011, 85, 6882–6892. [Google Scholar] [CrossRef] [PubMed]
  182. Pager, C.T.; Schutz, S.; Abraham, T.M.; Luo, G.; Sarnow, P. Modulation of hepatitis C virus RNA abundance and virus release by dispersion of processing bodies and enrichment of stress granules. Virology 2013, 435, 472–484. [Google Scholar] [CrossRef] [PubMed]
  183. Khong, A.; Jan, E. Modulation of stress granules and P bodies during dicistrovirus infection. J. Virol. 2011, 85, 1439–1451. [Google Scholar] [CrossRef] [PubMed]
  184. Fields, B.N.; Knipe, D.M.; Howley, P.M. Fields Virology, 6th ed.; Wolters Kluwer Health/Lippincott Williams & Wilkins: Philadelphia, PA, USA, 2013. [Google Scholar]
  185. Raaben, M.; Groot Koerkamp, M.J.; Rottier, P.J.; de Haan, C.A. Mouse hepatitis coronavirus replication induces host translational shutoff and mRNA decay, with concomitant formation of stress granules and processing bodies. Cell. Microbiol. 2007, 9, 2218–2229. [Google Scholar] [CrossRef] [PubMed]
  186. Sola, I.; Galan, C.; Mateos-Gomez, P.A.; Palacio, L.; Zuniga, S.; Cruz, J.L.; Almazan, F.; Enjuanes, L. The polypyrimidine tract-binding protein affects coronavirus RNA accumulation levels and relocalizes viral RNAs to novel cytoplasmic domains different from replication-transcription sites. J. Virol. 2011, 85, 5136–5149. [Google Scholar] [CrossRef] [PubMed]
  187. Humoud, M.N.; Doyle, N.; Royall, E.; Willcocks, M.M.; Sorgeloos, F.; van Kuppeveld, F.; Roberts, L.O.; Goodfellow, I.G.; Langereis, M.A.; Locker, N. Feline Calicivirus infection disrupts the assembly of cytoplasmic stress granules and induces G3BP1 cleavage. J. Virol. 2016. [Google Scholar] [CrossRef] [PubMed]
  188. Royall, E.; Doyle, N.; Abdul-Wahab, A.; Emmott, E.; Morley, S.J.; Goodfellow, I.; Roberts, L.O.; Locker, N. Murine Norovirus 1 (MNV1) replication induces translational control of the host by regulating eIF4E activity during infection. J. Biol. Chem. 2015. [Google Scholar] [CrossRef] [PubMed]
  189. Khaperskyy, D.A.; Hatchette, T.F.; McCormick, C. Influenza A virus inhibits cytoplasmic stress granule formation. FASEB J. Off. Publ. Fed. Am. Soc. Exp. Biol. 2012, 26, 1629–1639. [Google Scholar] [CrossRef] [PubMed]
  190. Mok, B.W.-Y.; Song, W.; Wang, P.; Tai, H.; Chen, Y.; Zheng, M.; Wen, X.; Lau, S.-Y.; Wu, W.L.; Matsumoto, K.; et al. The NS1 protein of influenza A virus interacts with cellular processing bodies and stress granules through RNA-associated protein 55 (RAP55) during virus infection. J. Virol. 2012, 86, 12695–12707. [Google Scholar] [CrossRef] [PubMed]
  191. Khaperskyy, D.A.; Emara, M.M.; Johnston, B.P.; Anderson, P.; Hatchette, T.F.; McCormick, C. Influenza a virus host shutoff disables antiviral stress-induced translation arrest. PLoS Pathog. 2014, 10, e1004217. [Google Scholar] [CrossRef] [PubMed]
  192. Thulasi Raman, S.N.; Liu, G.; Pyo, H.M.; Cui, Y.C.; Xu, F.; Ayalew, L.E.; Tikoo, S.K.; Zhou, Y. DDX3 interacts with influenza A NS1 and NP proteins and exerts antiviral function through regulation of stress granule formation. J. Virol. 2016, 90, 3661–3675. [Google Scholar] [CrossRef] [PubMed]
  193. Linero, F.N.; Thomas, M.G.; Boccaccio, G.L.; Scolaro, L.A. Junin virus infection impairs stress-granule formation in Vero cells treated with arsenite via inhibition of eIF2alpha phosphorylation. J. Gen. Virol. 2011, 92, 2889–2899. [Google Scholar] [CrossRef] [PubMed]
  194. Baird, N.L.; York, J.; Nunberg, J.H. Arenavirus infection induces discrete cytosolic structures for RNA replication. J. Virol. 2012, 86, 11301–11310. [Google Scholar] [CrossRef] [PubMed]
  195. Dinh, P.X.; Beura, L.K.; Das, P.B.; Panda, D.; Das, A.; Pattnaik, A.K. Induction of stress granule-like structures in vesicular stomatitis virus-infected cells. J. Virol. 2013, 87, 372–383. [Google Scholar] [CrossRef] [PubMed]
  196. Lindquist, M.E.; Lifland, A.W.; Utley, T.J.; Santangelo, P.J.; Crowe, J.E., Jr. Respiratory syncytial virus induces host RNA stress granules to facilitate viral replication. J. Virol. 2010, 84, 12274–12284. [Google Scholar] [CrossRef] [PubMed]
  197. Lindquist, M.E.; Mainou, B.A.; Dermody, T.S.; Crowe, J.E., Jr. Activation of protein kinase R is required for induction of stress granules by respiratory syncytial virus but dispensable for viral replication. Virology 2011, 413, 103–110. [Google Scholar] [CrossRef] [PubMed]
  198. Lifland, A.W.; Jung, J.; Alonas, E.; Zurla, C.; Crowe, J.E., Jr.; Santangelo, P.J. Human respiratory syncytial virus nucleoprotein and inclusion bodies antagonize the innate immune response mediated by MDA5 and MAVS. J. Virol. 2012, 86, 8245–8258. [Google Scholar] [CrossRef] [PubMed]
  199. Fricke, J.; Koo, L.Y.; Brown, C.R.; Collins, P.L. p38 and OGT sequestration into viral inclusion bodies in cells infected with human respiratory syncytial virus suppresses MK2 activities and stress granule assembly. J. Virol. 2013, 87, 1333–1347. [Google Scholar] [CrossRef] [PubMed]
  200. Randall, R.E.; Goodbourn, S. Interferons and viruses: An interplay between induction, signalling, antiviral responses and virus countermeasures. J. Gen. Virol. 2008, 89, 1–47. [Google Scholar] [CrossRef] [PubMed]
  201. Iseni, F.; Garcin, D.; Nishio, M.; Kedersha, N.; Anderson, P.; Kolakofsky, D. Sendai virus trailer RNA binds TIAR, a cellular protein involved in virus-induced apoptosis. EMBO J. 2002, 21, 5141–5150. [Google Scholar] [CrossRef] [PubMed]
  202. Yoshida, A.; Kawabata, R.; Honda, T.; Tomonaga, K.; Sakaguchi, T.; Irie, T. IFN-beta-inducing, unusual viral RNA species produced by paramyxovirus infection accumulated into distinct cytoplasmic structures in an RNA-type-dependent manner. Front. Microbiol. 2015, 6, 804. [Google Scholar] [CrossRef] [PubMed]
  203. Garcin, D.; Lezzi, M.; Dobbs, M.; Elliott, R.M.; Schmaljohn, C.; Kang, C.Y.; Kolakofsky, D. The 5′ ends of Hantaan virus (Bunyaviridae) RNAs suggest a prime-and-realign mechanism for the initiation of RNA synthesis. J. Virol. 1995, 69, 5754–5762. [Google Scholar] [PubMed]
  204. Mir, M.A.; Duran, W.A.; Hjelle, B.L.; Ye, C.; Panganiban, A.T. Storage of cellular 5′ mRNA caps in P bodies for viral cap-snatching. Proc. Natl. Acad. Sci. USA 2008, 105, 19294–19299. [Google Scholar] [CrossRef] [PubMed]
  205. Cheng, E.; Mir, M.A. Signatures of host mRNA 5′ terminus for efficient hantavirus cap snatching. J. Virol. 2012, 86, 10173–10185. [Google Scholar] [CrossRef] [PubMed]
  206. Hopkins, K.C.; Tartell, M.A.; Herrmann, C.; Hackett, B.A.; Taschuk, F.; Panda, D.; Menghani, S.V.; Sabin, L.R.; Cherry, S. Virus-induced translational arrest through 4EBP1/2-dependent decay of 5′-TOP mRNAs restricts viral infection. Proc. Natl. Acad. Sci. USA 2015, 112, E2920–E2929. [Google Scholar] [CrossRef] [PubMed]
  207. Mir, M.A.; Panganiban, A.T. A protein that replaces the entire cellular eIF4F complex. EMBO J. 2008, 27, 3129–3139. [Google Scholar] [CrossRef] [PubMed]
  208. Ganaie, S.S.; Haque, A.; Cheng, E.; Bonny, T.S.; Salim, N.N.; Mir, M.A. Ribosomal protein S19-binding domain provides insights into hantavirus nucleocapsid protein-mediated translation initiation mechanism. Boilchem. J. 2014, 464, 109–121. [Google Scholar] [CrossRef] [PubMed]
  209. Blakqori, G.; van Knippenberg, I.; Elliott, R.M. Bunyamwera orthobunyavirus S-segment untranslated regions mediate poly(A) tail-independent translation. J. Virol. 2009, 83, 3637–3646. [Google Scholar] [CrossRef] [PubMed]
  210. Vera-Otarola, J.; Soto-Rifo, R.; Ricci, E.P.; Ohlmann, T.; Darlix, J.L.; Lopez-Lastra, M. The 3′ untranslated region of the Andes hantavirus small mRNA functionally replaces the poly(A) tail and stimulates cap-dependent translation initiation from the viral mRNA. J. Virol. 2010, 84, 10420–10424. [Google Scholar] [CrossRef] [PubMed]
  211. Wang, X.; Li, M.; Zheng, H.; Muster, T.; Palese, P.; Beg, A.A.; Garcia-Sastre, A. Influenza A virus NS1 protein prevents activation of NF-kappaB and induction of alpha/beta interferon. J. Virol. 2000, 74, 11566–11573. [Google Scholar] [CrossRef] [PubMed]
  212. Dalrymple, N.A.; Cimica, V.; Mackow, E.R. Dengue virus NS proteins inhibit RIG-I/MAVS signaling by blocking TBK1/IRF3 phosphorylation: Dengue Virus serotype 1 NS4A Is a unique interferon-regulating virulence determinant. mBio 2015, 6. [Google Scholar] [CrossRef] [PubMed]
  213. Jaaskelainen, K.M.; Kaukinen, P.; Minskaya, E.S.; Plyusnina, A.; Vapalahti, O.; Elliott, R.M.; Weber, F.; Vaheri, A.; Plyusnin, A. Tula and Puumala hantavirus NSs ORFs are functional and the products inhibit activation of the interferon-beta promoter. J. Med. Virol. 2007, 79, 1527–1536. [Google Scholar] [CrossRef] [PubMed]
  214. Kohl, A.; Clayton, R.F.; Weber, F.; Bridgen, A.; Randall, R.E.; Elliott, R.M. Bunyamwera virus nonstructural protein NSs counteracts interferon regulatory factor 3-mediated induction of early cell death. J. Virol. 2003, 77, 7999–8008. [Google Scholar] [CrossRef] [PubMed]
  215. Weber, F.; Bridgen, A.; Fazakerley, J.K.; Streitenfeld, H.; Kessler, N.; Randall, R.E.; Elliott, R.M. Bunyamwera bunyavirus nonstructural protein NSs counteracts the induction of alpha/beta interferon. J. Virol. 2002, 76, 7949–7955. [Google Scholar] [CrossRef] [PubMed]
  216. Billecocq, A.; Spiegel, M.; Vialat, P.; Kohl, A.; Weber, F.; Bouloy, M.; Haller, O. NSs protein of Rift Valley fever virus blocks interferon production by inhibiting host gene transcription. J. Virol. 2004, 78, 9798–9806. [Google Scholar] [CrossRef] [PubMed]
  217. Alff, P.J.; Gavrilovskaya, I.N.; Gorbunova, E.; Endriss, K.; Chong, Y.; Geimonen, E.; Sen, N.; Reich, N.C.; Mackow, E.R. The pathogenic NY-1 hantavirus G1 cytoplasmic tail inhibits RIG-I- and TBK-1-directed interferon responses. J. Virol. 2006, 80, 9676–9686. [Google Scholar] [CrossRef] [PubMed]
  218. Cimica, V.; Dalrymple, N.A.; Roth, E.; Nasonov, A.; Mackow, E.R. An innate immunity-regulating virulence determinant is uniquely encoded by the Andes virus nucleocapsid protein. mBio 2014, 5. [Google Scholar] [CrossRef] [PubMed]
  219. Matthys, V.S.; Cimica, V.; Dalrymple, N.A.; Glennon, N.B.; Bianco, C.; Mackow, E.R. Hantavirus GnT elements mediate TRAF3 binding and inhibit RIG-I/TBK1-directed beta interferon transcription by blocking IRF3 phosphorylation. J. Virol. 2014, 88, 2246–2259. [Google Scholar] [CrossRef] [PubMed]
  220. Wang, Z.; Mir, M.A. Andes virus nucleocapsid protein interrupts protein kinase R dimerization to counteract host interference in viral protein synthesis. J. Virol. 2015, 89, 1628–1639. [Google Scholar] [CrossRef] [PubMed]
  221. Ikegami, T.; Narayanan, K.; Won, S.; Kamitani, W.; Peters, C.J.; Makino, S. Rift Valley fever virus NSs protein promotes post-transcriptional downregulation of protein kinase PKR and inhibits eIF2alpha phosphorylation. PLoS Pathog. 2009, 5, e1000287. [Google Scholar] [CrossRef] [PubMed]
  222. Habjan, M.; Pichlmair, A.; Elliott, R.M.; Overby, A.K.; Glatter, T.; Gstaiger, M.; Superti-Furga, G.; Unger, H.; Weber, F. NSs protein of rift valley fever virus induces the specific degradation of the double-stranded RNA-dependent protein kinase. J. Virol. 2009, 83, 4365–4375. [Google Scholar] [CrossRef] [PubMed]
  223. Vialat, P.; Bouloy, M. Germiston virus transcriptase requires active 40S ribosomal subunits and utilizes capped cellular RNAs. J. Virol. 1992, 66, 685–693. [Google Scholar] [PubMed]
  224. Barr, J.N. Bunyavirus mRNA synthesis is coupled to translation to prevent premature transcription termination. RNA 2007, 13, 731–736. [Google Scholar] [CrossRef] [PubMed]
  225. Nelson, E.V.; Schmidt, K.M.; Deflube, L.R.; Doganay, S.; Banadyga, L.; Olejnik, J.; Hume, A.J.; Ryabchikova, E.; Ebihara, H.; Kedersha, N.; et al. Ebola virus does not induce stress granule formation during infection and sequesters stress granule proteins within viral inclusions. J. Virol. 2016, in press. [Google Scholar] [CrossRef] [PubMed]
  226. Legros, S.; Boxus, M.; Gatot, J.S.; Van Lint, C.; Kruys, V.; Kettmann, R.; Twizere, J.C.; Dequiedt, F. The HTLV-1 Tax protein inhibits formation of stress granules by interacting with histone deacetylase 6. Oncogene 2011, 30, 4050–4062. [Google Scholar] [CrossRef] [PubMed]
  227. Takahashi, M.; Higuchi, M.; Makokha, G.N.; Matsuki, H.; Yoshita, M.; Tanaka, Y.; Fujii, M. HTLV-1 Tax oncoprotein stimulates ROS production and apoptosis in T cells by interacting with USP10. Blood 2013, 122, 715–725. [Google Scholar] [CrossRef] [PubMed]
  228. Abrahamyan, L.G.; Chatel-Chaix, L.; Ajamian, L.; Milev, M.P.; Monette, A.; Clement, J.F.; Song, R.; Lehmann, M.; DesGroseillers, L.; Laughrea, M.; et al. Novel Staufen1 ribonucleoproteins prevent formation of stress granules but favour encapsidation of HIV-1 genomic RNA. J. Cell Sci. 2010, 123, 369–383. [Google Scholar] [CrossRef] [PubMed]
  229. Valiente-Echeverria, F.; Melnychuk, L.; Vyboh, K.; Ajamian, L.; Gallouzi, I.E.; Bernard, N.; Mouland, A.J. eEF2 and Ras-GAP SH3 domain-binding protein (G3BP1) modulate stress granule assembly during HIV-1 infection. Nat. Commun. 2014, 5, 4819. [Google Scholar] [CrossRef] [PubMed]
  230. Cobos Jimenez, V.; Martinez, F.O.; Booiman, T.; van Dort, K.A.; van de Klundert, M.A.; Gordon, S.; Geijtenbeek, T.B.; Kootstra, N.A. G3BP1 restricts HIV-1 replication in macrophages and T-cells by sequestering viral RNA. Virology 2015, 486, 94–104. [Google Scholar] [CrossRef] [PubMed]
  231. Soto-Rifo, R.; Valiente-Echeverria, F.; Rubilar, P.S.; Garcia-de-Gracia, F.; Ricci, E.P.; Limousin, T.; Decimo, D.; Mouland, A.J.; Ohlmann, T. HIV-2 genomic RNA accumulates in stress granules in the absence of active translation. Nucleic Acids Res. 2014, 42, 12861–12875. [Google Scholar] [CrossRef] [PubMed]
  232. Fujimura, K.; Sasaki, A.T.; Anderson, P. Selenite targets eIF4E-binding protein-1 to inhibit translation initiation and induce the assembly of non-canonical stress granules. Nucleic Acids Res. 2012, 40, 8099–8110. [Google Scholar] [CrossRef] [PubMed]
  233. Cinti, A.; Le Sage, V.; Ghanem, M.; Mouland, A.J. HIV-1 gag blocks selenite-induced stress granule assembly by altering the mRNA cap-binding complex. mBio 2016, 7. [Google Scholar] [CrossRef] [PubMed]
  234. Smith, J.A.; Schmechel, S.C.; Raghavan, A.; Abelson, M.; Reilly, C.; Katze, M.G.; Kaufman, R.J.; Bohjanen, P.R.; Schiff, L.A. Reovirus induces and benefits from an integrated cellular stress response. J. Virol. 2006, 80, 2019–2033. [Google Scholar] [CrossRef] [PubMed]
  235. Hanley, L.L.; McGivern, D.R.; Teng, M.N.; Djang, R.; Collins, P.L.; Fearns, R. Roles of the respiratory syncytial virus trailer region: Effects of mutations on genome production and stress granule formation. Virology 2010, 406, 241–252. [Google Scholar] [CrossRef] [PubMed]
  236. Soto-Rifo, R.; Rubilar, P.S.; Ohlmann, T. The DEAD-box helicase DDX3 substitutes for the cap-binding protein eIF4E to promote compartmentalized translation initiation of the HIV-1 genomic RNA. Nucleic Acids Res. 2013, 41, 6286–6299. [Google Scholar] [CrossRef] [PubMed]
  237. Greer, A.E.; Hearing, P.; Ketner, G. The adenovirus E4 11k protein binds and relocalizes the cytoplasmic P-body component Ddx6 to aggresomes. Virology 2011, 417, 161–168. [Google Scholar] [CrossRef] [PubMed]
  238. Seto, E.; Inoue, T.; Nakatani, Y.; Yamada, M.; Isomura, H. Processing bodies accumulate in human cytomegalovirus-infected cells and do not affect viral replication at high multiplicity of infection. Virology 2014, 458–459, 151–161. [Google Scholar] [CrossRef] [PubMed]
  239. Bhowmick, R.; Mukherjee, A.; Patra, U.; Chawla-Sarkar, M. Rotavirus disrupts cytoplasmic P bodies during infection. Virus Res. 2015, 210, 344–354. [Google Scholar] [CrossRef] [PubMed]
  240. Chahar, H.S.; Chen, S.; Manjunath, N. P-body components LSM1, GW182, DDX3, DDX6 and XRN1 are recruited to WNV replication sites and positively regulate viral replication. Virology 2013, 436, 1–7. [Google Scholar] [CrossRef] [PubMed]
  241. Silva, P.A.; Pereira, C.F.; Dalebout, T.J.; Spaan, W.J.; Bredenbeek, P.J. An RNA pseudoknot is required for production of yellow fever virus subgenomic RNA by the host nuclease XRN1. J. Virol. 2010, 84, 11395–11406. [Google Scholar] [CrossRef] [PubMed]
  242. Pijlman, G.P.; Funk, A.; Kondratieva, N.; Leung, J.; Torres, S.; van der Aa, L.; Liu, W.J.; Palmenberg, A.C.; Shi, P.Y.; Hall, R.A.; et al. A highly structured, nuclease-resistant, noncoding RNA produced by flaviviruses is required for pathogenicity. Cell Host Microbe 2008, 4, 579–591. [Google Scholar] [CrossRef] [PubMed]
  243. Moon, S.L.; Anderson, J.R.; Kumagai, Y.; Wilusz, C.J.; Akira, S.; Khromykh, A.A.; Wilusz, J. A noncoding RNA produced by arthropod-borne flaviviruses inhibits the cellular exoribonuclease XRN1 and alters host mRNA stability. RNA 2012, 18, 2029–2040. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  244. Ariumi, Y.; Kuroki, M.; Abe, K.; Dansako, H.; Ikeda, M.; Wakita, T.; Kato, N. DDX3 DEAD-box RNA helicase is required for hepatitis C virus RNA replication. J. Virol. 2007, 81, 13922–13926. [Google Scholar] [CrossRef] [PubMed]
  245. Scheller, N.; Mina, L.B.; Galão, R.P.; Chari, A.; Giménez-Barcons, M.; Noueiry, A.; Fischer, U.; Meyerhans, A.; Díez, J. Translation and replication of hepatitis C virus genomic RNA depends on ancient cellular proteins that control mRNA fates. Proc. Natl. Acad. Sci. USA 2009, 106, 13517–13522. [Google Scholar] [CrossRef] [PubMed]
  246. Jangra, R.K.; Yi, M.; Lemon, S.M. DDX6 (Rck/p54) is required for efficient hepatitis C virus replication but not for internal ribosome entry site-directed translation. J. Virol. 2010, 84, 6810–6824. [Google Scholar] [CrossRef] [PubMed]
  247. Perez-Vilaro, G.; Scheller, N.; Saludes, V.; Diez, J. Hepatitis C virus infection alters P-body composition but is independent of P-body granules. J. Virol. 2012, 86, 8740–8749. [Google Scholar] [CrossRef] [PubMed]
  248. Dougherty, J.D.; White, J.P.; Lloyd, R.E. Poliovirus-mediated disruption of cytoplasmic processing bodies. J. Virol. 2011, 85, 64–75. [Google Scholar] [CrossRef] [PubMed]
  249. Sokoloski, K.J.; Dickson, A.M.; Chaskey, E.L.; Garneau, N.L.; Wilusz, C.J.; Wilusz, J. Sindbis virus usurps the cellular HuR protein to stabilize its transcripts and promote productive infections in mammalian and mosquito cells. Cell Host Microbe 2010, 8, 196–207. [Google Scholar] [CrossRef] [PubMed]
  250. Marin, M.; Golem, S.; Rose, K.M.; Kozak, S.L.; Kabat, D. Human immunodeficiency virus type 1 Vif functionally interacts with diverse APOBEC3 cytidine deaminases and moves with them between cytoplasmic sites of mRNA metabolism. J. Virol. 2008, 82, 987–998. [Google Scholar] [CrossRef] [PubMed]
  251. Chable-Bessia, C.; Meziane, O.; Latreille, D.; Triboulet, R.; Zamborlini, A.; Wagschal, A.; Jacquet, J.M.; Reynes, J.; Levy, Y.; Saib, A.; et al. Suppression of HIV-1 replication by microRNA effectors. RetroVirology 2009, 6, 26. [Google Scholar] [CrossRef] [PubMed]
  252. Martin, K.L.; Johnson, M.; D’Aquila, R.T. APOBEC3G complexes decrease human immunodeficiency virus type 1 production. J. Virol. 2011, 85, 9314–9326. [Google Scholar] [CrossRef] [PubMed]
  253. Nathans, R.; Chu, C.Y.; Serquina, A.K.; Lu, C.C.; Cao, H.; Rana, T.M. Cellular microRNA and P bodies modulate host-HIV-1 interactions. Mol. Cell 2009, 34, 696–709. [Google Scholar] [CrossRef] [PubMed]
  254. Phalora, P.K.; Sherer, N.M.; Steven, M.; Swanson, C.M.; Malim, M.H.; Phalora, P.K.; Sherer, N.M.; Wolinsky, S.M.; Swanson, C.M.; Malim, M.H. HIV-1 replication and APOBEC3 antiviral activity are not. J. Virol. 2012, 86, 11712. [Google Scholar] [CrossRef] [PubMed]
  255. Bouttier, M.; Saumet, A.; Peter, M.; Courgnaud, V.; Schmidt, U.; Cazevieille, C.; Bertrand, E.; Lecellier, C.H. Retroviral GAG proteins recruit AGO2 on viral RNAs without affecting RNA accumulation and translation. Nucleic Acids Res. 2012, 40, 775–786. [Google Scholar] [CrossRef] [PubMed]
  256. Reed, J.C.; Molter, B.; Geary, C.D.; McNevin, J.; McElrath, J.; Giri, S.; Klein, K.C.; Lingappa, J.R. HIV-1 Gag co-opts a cellular complex containing DDX6, a helicase that facilitates capsid assembly. J. Cell Biol. 2012, 198, 439–456. [Google Scholar] [CrossRef] [PubMed]
  257. Lingappa, J.R.; Hill, R.L.; Wong, M.L.; Hegde, R.S. A multistep, ATP-dependent pathway for assembly of human immunodeficiency virus capsids in a cell-free system. J. Cell Biol. 1997, 136, 567–581. [Google Scholar] [CrossRef] [PubMed]
  258. Burdick, R.; Smith, J.L.; Chaipan, C.; Friew, Y.; Chen, J.; Venkatachari, N.J.; Delviks-Frankenberry, K.A.; Hu, W.S.; Pathak, V.K. P body-associated protein Mov10 inhibits HIV-1 replication at multiple stages. J. Virol. 2010, 84, 10241–10253. [Google Scholar] [CrossRef] [PubMed]
  259. Wang, X.; Han, Y.; Dang, Y.; Fu, W.; Zhou, T.; Ptak, R.G.; Zheng, Y.H. Moloney leukemia virus 10 (MOV10) protein inhibits retrovirus replication. J. Biol. Chem. 2010, 285, 14346–14355. [Google Scholar] [CrossRef] [PubMed]
  260. Izumi, T.; Burdick, R.; Shigemi, M.; Plisov, S.; Hu, W.S.; Pathak, V.K. Mov10 and APOBEC3G localization to processing bodies is not required for virion incorporation and antiviral activity. J. Virol. 2013, 87, 11047–11062. [Google Scholar] [CrossRef] [PubMed]
  261. Panas, M.D.; Kedersha, N.; McInerney, G.M. Methods for the characterization of stress granules in virus infected cells. Methods 2015, 90, 57–64. [Google Scholar] [CrossRef] [PubMed]
  262. Lin, Y.; Protter, D.S.; Rosen, M.K.; Parker, R. Formation and maturation of phase-separated liquid droplets by RNA-binding proteins. Mol. Cell 2015. [Google Scholar] [CrossRef] [PubMed]
  263. Kroschwald, S.; Maharana, S.; Mateju, D.; Malinovska, L.; Nuske, E.; Poser, I.; Richter, D.; Alberti, S. Promiscuous interactions and protein disaggregases determine the material state of stress-inducible RNP granules. Elife 2015, 4. [Google Scholar] [CrossRef] [PubMed]
  264. Hilliker, A. Analysis of RNA helicases in P-bodies and stress granules. Methods Enzymol. 2012, 511, 323–346. [Google Scholar] [PubMed]
  265. Kedersha, N.; Anderson, P. Mammalian stress granules and processing bodies. Methods Enzymol. 2007, 431, 61–81. [Google Scholar] [PubMed]
  266. Kedersha, N.; Tisdale, S.; Hickman, T.; Anderson, P. Real-time and quantitative imaging of mammalian stress granules and processing bodies. Methods Enzymol. 2008, 448, 521–552. [Google Scholar] [PubMed]
  267. Lin, R.-J. RNA-Protein Complexes and Interactions; Springer: New York, NY, USA, 2016; Volume 1421. [Google Scholar]
  268. Kato, M.; Han, T.W.; Xie, S.; Shi, K.; Du, X.; Wu, L.C.; Mirzaei, H.; Goldsmith, E.J.; Longgood, J.; Pei, J.; et al. Cell-free formation of RNA granules: Low complexity sequence domains form dynamic fibers within hydrogels. Cell 2012, 149, 753–767. [Google Scholar] [CrossRef] [PubMed]
  269. Han, T.W.; Kato, M.; Xie, S.; Wu, L.C.; Mirzaei, H.; Pei, J.; Chen, M.; Xie, Y.; Allen, J.; Xiao, G.; et al. Cell-free formation of RNA granules: Bound RNAs identify features and components of cellular assemblies. Cell 2012, 149, 768–779. [Google Scholar] [CrossRef] [PubMed]
  270. Jain, S.; Wheeler, J.R.; Walters, R.W.; Agrawal, A.; Barsic, A.; Parker, R. ATPase-modulated stress granules contain a diverse proteome and substructure. Cell 2016, 164, 487–498. [Google Scholar] [CrossRef] [PubMed]
Table 1. Components of stress granules (SGs) and processing bodies (PBs).
Table 1. Components of stress granules (SGs) and processing bodies (PBs).
SG ComponentReferenceSG ComponentReference
ADAR[38,50]MEX67[51]
AKAP350[52]MLN51[53]
ANG[54]MSI1[55]
ATXN2/pbp1[56,57]mTOR[37,58]
CALR[59]OGFOD1[60]
Caprin1[61]P97/NAT1[62]
CCAR1[52]PABP1[30]
CIRP[63]PHB2[64]
CUGBP1[65]PKP1/3[66]
DDX1[67]PKR[38,50]
DDX3/Ded1[51]PMR1[33]
DHC1[68]PRTB[69]
DISC1[70]PUM1[71]
eIF2B[72]PUM2[73]
eIF2α[74]RACK1[75]
eIF3[17]RBM42[76]
eIF4A1[17]RHAU/DHX36[36]
eIF4G[17]RIG-I[38,50]
eIF5A[77]RNH1[54]
FAK[78]Rpp20[79]
FBP/KSRP[80]RSK-2[81]
FUS[82]Sam68[83]
FXR1/FXMR[31]SERBP1[84]
FXR2P[31]SGNP[85]
G3BP1[20]SMN[34]
G3BP2[86]Staufen1[35]
Grb7[78]TDP43[87]
HDAC6[88]TDRD3[84]
hnRNP A1[89]TFE3[90]
hnRNP K[76]TFEB[90]
Hsp27[91]TRAF2[92]
HuD[93]USP10[21,94]
HuR[95]Vinexin[96]
IFIH1/ MDA-5[38,50]ZBP-1[97]
IP5K[98]
KHC/KLC[68]
LINE1 ORF1p[99]
MBNL1[67]
PB ComponentReferencePB ComponentReference
Ccr4[100]Htt[101]
Dcp1/Dcp2[100]LSM1[40]
Ebs1[102]Pan2/3[103]
Edc1-2[104]Pat1/PatL1[100]
Edc3[105]Pop2/Caf1[106]
eIF4E-T[107]PRMT1[108]
Ge-1/Hedls[41]TNRC6B[109]
GW182[110]UPF1[42]
hMex3A[111]UPF2[42]
hnRNP A3[112]UPF3[42]
SG and PB ComponentReferenceSG and PB ComponentReference
Ago1[48]NXF7[112]
Ago2[113]PABP/Pab1[30,114]
APOBEC3G[47,115]PCBP2[116]
BRF1[117]Rap55/Scd6[118]
CPEB[32]RCK/Dhh1/DDX6[32]
Dcp1/Dcp1a[32]Roquin[119]
eIF4E[17]Smaug 1[120,121]
FAST[117]TIA1/TIAR[30,91,122]
hMex3B[123]TTP/BRF1[117]
hnRNP Q[124]Xrn1[117]
IPO8[125]YB-1[126]
JNK[127]
Lin28[128]
Table 2. Virus families that modulate SGs.
Table 2. Virus families that modulate SGs.
GenomeVirus FamilyVirusSG InductionSG BlockadeMechanismReference
IdsDNAHerpesviridaeHerpes simplex virus type 1 (HSV-1)NoYes(-)RNA stem loop captures TIA-1/TIAR to favor replication[133]
Herpes simplex virus type 2 (HSV-2)YesYesInhibits SG assembly dependent of eIF2α-P[136]
Induces SG independent of eIF2α-P[136]
Cytomegalovirus (HCMV)NoYesInduces UPR with intact viral translation[138]
PoxviridaeVaccinia virus (VV)YesYesReplication factories (RF) sequester G3BP1, Caprin1[140]
RF sequester eIF4G, eIF4E, PABP[141]
VV lacking of E3L induces antiviral granules (AVGs)[94]
IIIdsRNAReoviridaeRotavirusNoYesNSP2, VP2 and NSP5 translocate PABP to the nucleus[144]
Mammalian orthoreovirus (MRV)YesYesInduces eIF2α-P[234]
uNS is recruited to SGs[147,148]
IV(+)ssRNAPicornaviridaePoliovirus (PV)YesYesEarly PV-infection induces SG assembly[16]
viral C3 protease cleaves G3BP1[150]
PV-infection induces TIA-1 aggregates[151]
Encephalomyocarditis virus (EMCV)NoYesCleavage of G3BP1[50]
Coxsackievirus B3 (CVB3)NoYesCleavage of G3BP1[153]
Theiler’s murine encephalomyelitis virus (TMEV)NoYesLeader protein (L) inhibits SG assembly[154]
Mengovirus, a strain of EMCVNoYesLeader protein (L) inhibits SG assembly[155,156]
TogaviridaeSemliki Forest virus (SFV)YesYesInduces eIF2α-P[157]
nsP3 protein captures G3BP1[158]
Chikungunya virus (CHIKV)NoYesnsp3 protein recruits G3BP1 to replication foci[159]
G3BP2 colocalize with nsP3/nsP2[161]
Rubella virus (RUBV)YesNoAccumulation of G3BP[162]
Sindbis virus (SINV)YesYesNsp4 interacts with G3BP1[165]
Induces PKR-mediated SG assembly[166]
FlaviviridaeWest Nile Virus (WNV)NoYes3′-end viral genome captures TIA-1/TIAR[167]
Dengue virus (DENV)NoYes3′-end viral genome captures TIA-1/TIAR[168]
3′ UTR interacts with G3BP1, G3BP2, Caprin1 and USP1[171]
Tick-borne encephalitis virus (TBEV)YesNoInduces eIF2α-P[173]
Japanese encephalitis virus (JEV)NoYesCore protein interacts with Caprin1[174]
Bovine viral diarrhea virus (BVDV)NoYesImpairs the Ars-mediated SG assembly[175]
Hepatitis C virus (HCV)YesYesG3BP1, ataxin-2 and PABP localized to lipid droplets[181]
Induces PKR[176]
SG disassembly mediated by GADD34[177]
DDX3 binds 3′ UTR[178]
DDX3 and G3BP1 localize with HCV core protein[179]
DicistroviridaeCricket paralysis virus (CrPV)NoYes3Cpro sequesters to SG[183]
CoronaviridaeMouse hepatitis coronavirus (MHV)YesNoInduces eIF2α-P[185]
Transmissible gastroenteritis virus (TGEV)YesNoPTB localizes to SG and correlates with replication increase[186]
CaliciviridaeMurine Norovirus 1 (MNV1) ndNoeIF4E phosphorylation[188]
Feline Calicivirus (FCV)NoYesCleavage of G3BP1 by FCV NS6[187]
V(-)ssRNAOrthomyxoviridaeInfluenza A virus (FLUA)NoYesNS1 inhibits PKR and eIF2α-P[189,190]
NP and PA-X block SGs[191]
DDX3 colocalize with NP[192]
ArenaviridaeJunin virus (JUNV)NoYesN and GPC proteins impairs SG assembly[193,194]}
RhabdoviridaeVesicular stomatitis virus (VSV)YesNoInduces SG-like structures recruiting TIA-1, TIAR y PCBP2[195]
ParamyxoviridaeRespiratory syncytial virus (RSV)YesYesInduces PKR[196,197]
5′ trailer region induces eIF3-aggregates[235]
RSV sequesters p38-P and OGN[199]
Measles virus (MeV)YesndInduces PKR[38]
viral C protein and ADAR1 modulate SGs[38]
Sendai virus (SV)YesYesTrailer RNA captures TIAR from SGs[201]
Forms antiviral stress granules (avSG)[202]
BunyaviridaeRift Valley fever virus (RVFV)YesYesattenuate Akt/mTOR signaling[206]
Andes hantavirus (ANDV)ndYesN protein inhibits PKR activation[220]
FiloviridaeEbola virusNoYesVP35 prevents SG formation by blocking PKR activation[225]
VIssRNA-RTRetroviridaeHuman T-cell leukemia virus (HTLV-1)NoYesTax protein interacts with HDAC6[226]
Tax protein interacts with USP10[227]
Human immunodeficiency virus type 1 (HIV-1)NoYesStaufen 1 and Gag block SG assembly[228];
EEF2 interacts with Gag to blocks SG assembly[229]
G3BP1 interacts with Gag to disassembly SG-preformed[229]
gRNA promote pre-translation initiation complex[236]
Gag interacts with eIF4E to promote disassembly of SGs[233]
Human immunodeficiency virus type 2(HIV-2)YesNogRNA and TIAR aggregates in SG[231]
nd = not determined; dsDNA = double-stranded DNA; dsRNA = double-stranded RNA; UPR = unfolded protein response; (+)ssRNA = positive-sense single strand RNA; (-)ssRNA = negative-sense single stand RNA; ssRNA-RT = single strand RNA retroviruses
Table 3. Virus families that modulate PBs.
Table 3. Virus families that modulate PBs.
GenomeVirus FamilyVirusPB InductionPB BlockadeMechanismReference
IdsDNAAdenoviridaeAdenovirusNoYesDecreased PB by redistribution of E4 11K[237]
PapillomaviridaeHuman papilloma virus (HPV)Yes *NoRe-colocalization of PKR in PBs[46]
HerpesviridaeCytomegalovirus (HCMV)YesNoIncreased of Dcp1a, EDC4, Rck/p54/DDX6 and Rap55 protein levels[238]
IIIdsRNAReoviridaeRotavirusNoYesDecrease of XRN1, DCP1 and Pan3, but not GW182 protein levels[239]
IV(+)ssRNAFlaviviridaeWest Nile virus (WNV)NoYesCaptures of Lsm1, GW182, DDX6, DDX3 and Xrn1 to viral replication factories (RF)[168,240]
Dengue virus (DENV)NoYesCaptures of Lsm1, GW182, DDX6, DDX3 and Xrn1 to viral replication factories (RF)[168,240]
Yellow fever virus (YFV)Yes *NosfRNA stalls Xrn1 and co-localizes at PB[241]
Kunjin virus (KUNV), Australian strain of DENVYes *NosfRNA stalls Xrn1 and co-localizes at PB[242]
Hepatitis C virus (HCV)Yes *YesDDX6, Lsm1, Xrn1, PATL1 and Ago2 localize to lipid droplets[181]
Dcp2 not localize to viral factories[181]
PicornaviridaePoliovirus (PV)NoYesCleavage of Xrn1, Dcp1a and Pan3[248]
Protease 2A blocks PB formation[149]
Coxsackievirus B3 (CVB3)NoYesCleavage of Xrn1, Dcp1a and Pan3[248]
DicistroviridaeCricket paralysis virus (CrPV)YesYesDisrupts only GW182/Dcp1 aggregate, but not Ago1/Ago2[183]
TogaviridaeSindbis virus (SINV)NoYesHuR-translocation out of the nucleus[249]
V(-)ssRNAOrthomyxoviridaeInfluenza virus A (IAV)NoYesInteraction of RAP55 and NSP1[190]
BunyaviridaeHanta virusYes *NoCap snatching occurs in PBs[204]
VIssRNA-RTRetroviridaeHuman immunodeficiency virus type 1 (HIV-1)ndYesHIV-1 mRNA interacts with DDX6, Ago 2 and APOBE3G and displaces from the PB[253]
Relocalization of PB during the HIV-1 infection[228]
Assembly intermediates (AIs) recruits DDX6 and ABCE1[256]
Overexpression of MOV10 inhibits HIV-1 replication[258]
*maintains PB endogenously; nd = not determined; dsDNA = double-stranded DNA; dsRNA = double-stranded RNA; UPR = unfolded protein response; (+)ssRNA = positive-sense single strand RNA; (-)ssRNA = negative-sense single stand RNA; ssRNA-RT = single strand RNA retroviruses.

Share and Cite

MDPI and ACS Style

Poblete-Durán, N.; Prades-Pérez, Y.; Vera-Otarola, J.; Soto-Rifo, R.; Valiente-Echeverría, F. Who Regulates Whom? An Overview of RNA Granules and Viral Infections. Viruses 2016, 8, 180. https://0-doi-org.brum.beds.ac.uk/10.3390/v8070180

AMA Style

Poblete-Durán N, Prades-Pérez Y, Vera-Otarola J, Soto-Rifo R, Valiente-Echeverría F. Who Regulates Whom? An Overview of RNA Granules and Viral Infections. Viruses. 2016; 8(7):180. https://0-doi-org.brum.beds.ac.uk/10.3390/v8070180

Chicago/Turabian Style

Poblete-Durán, Natalia, Yara Prades-Pérez, Jorge Vera-Otarola, Ricardo Soto-Rifo, and Fernando Valiente-Echeverría. 2016. "Who Regulates Whom? An Overview of RNA Granules and Viral Infections" Viruses 8, no. 7: 180. https://0-doi-org.brum.beds.ac.uk/10.3390/v8070180

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop