Open access peer-reviewed chapter

Cellular Control of Dengue Virus Replication: Role of Interferon-Inducible Genes

Written By

Hirotaka Takahashi and Youichi Suzuki

Submitted: 13 October 2016 Reviewed: 20 February 2017 Published: 26 July 2017

DOI: 10.5772/67984

From the Edited Volume

Dengue - Immunopathology and Control Strategies

Edited by Márcia Aparecida Sperança

Chapter metrics overview

1,767 Chapter Downloads

View Full Metrics

Abstract

Dengue, one of the most common mosquito‐borne viral infectious diseases in the world, is caused by the dengue virus (DENV). This enveloped RNA virus has immunologically distinct serotypes that increase the risk of life‐threatening diseases, such as dengue haemorrhagic fever. However, no effective antiviral therapy against DENV infection has yet been established. As seen in other RNA viruses, various cellular factors have been reported to participate in efficient DENV replication. On the other hand, increasing recent evidence demonstrates that host cells harbour inhibitory factors that limit the DENV replication. In particular, it is well known that the response of interferons (IFNs), the first line of a host defence system against invading pathogens, evokes the expression of a number of genes that negatively regulate various steps of virus replication. This set of inhibitory genes, called interferon‐stimulated genes (ISGs), is considered to be a central force in IFN‐mediated antiviral responses. In this chapter, we focus our attention on the cellular factors involved in DENV infection, particularly to those that modulate DENV replication through their association with viral RNA. In addition, we also summarize general experimental approaches for identifying the host factors of RNA viruses, including DENV.

Keywords

  • dengue virus
  • cellular factors
  • RNA untranslated regions
  • interferon‐stimulated genes
  • identification systems

1. Introduction

Dengue virus (DENV) is an enveloped and positive‐strand RNA virus that belongs to the genus Flavivirus of the Flaviviridae family [1]. An important characteristic of the Flavivirus is that this genus consists of a large number of arthropod‐borne viruses, many of which are transmitted by mosquitoes and ticks. In addition, flavivirus infection often causes life‐threatening diseases in humans, such as haemorrhagic fever, encephalitis, and meningitis [2]. Recently, the Zika virus (ZIKV), a member of the flavivirus family that has spread explosively throughout the Americas, is reported to be associated with neurological complications [3, 4]. Flaviviruses, therefore, have significant clinical as well as economic impacts on modern society.

DENV is a mosquito‐borne virus widely distributed in the tropical and subtropical areas of the world. This flavivirus infection is transmitted to humans via the bite of infected mosquitos. The primary vector of DENV infection is Aedes aegypti, while Aedes albopictus, which originated in Asia but has extended its range to other regions of the world, is also capable of spreading a dengue outbreak. DENV has four antigenically distinct serotypes (from DENV‐1 to DENV‐4). Primary infection with one of the serotypes is often asymptomatic or causes self‐limiting dengue fever (DF). However, secondary infection of different serotypes increases the risk of more serious forms of DENV infection, such as dengue haemorrhagic fever (DHF) and dengue shock syndrome (DSS), due to the presence of non‐ or sub‐neutralizing antibodies generated during the primary infection. Therefore, dengue is a significant threat to humans, yet there is currently no specific antiviral available for DENV infection [1, 5]. However, it should be noted that a live attenuated vaccine against DENV developed by Sanofi Pasteur (Dengvaxia) has been licensed for use in a limited number of countries, including Mexico and the Philippines, although the efficacy of the DENV vaccine in endemic countries is still under investigation [6].

Advertisement

2. Brief overview of DENV replication

DENV infection begins with its entry into a permissive cell via receptor‐mediated endocytosis (Figure 1). So far, various types of human cells, such as macrophages, lymphocytes, hepatocytes, and endothelial cells, are reportedly susceptible to DENV infection. Among them, monocyte lineage cells (i.e., dendritic cells [DCs] and macrophages) are thought to be the primary targets of DENV in humans. As the entry receptors of DENV, several cellular proteins, including C‐type lectin receptors (e.g., DC‐SIGN/CD209, mannose receptor/CD206) and phosphatidylserine receptors (e.g., TIM, TAM), have been demonstrated [7].

Figure 1.

The genome structure and replication cycle of DENV.

Upon entry into the cell, a membrane fusion between DENV envelope (E) glycoprotein and endosomal vesicle occurs, leading to the release of viral RNA into the cytoplasm. The DENV genome is a single‐stranded positive‐sense RNA and is approximately 10‐kb long. The viral RNA contains a single long open reading frame (ORF) flanked by highly structured 5′ and 3′ untranslated regions (UTRs), with a type 1 cap (m7GpppAmp) at the 5′ terminus and no poly(A) tract at the 3’ terminus [8]. The single ORF is translated to a large polyprotein, which is subsequently cleaved co‐ and post‐translationally into three structural (capsid [C], pre‐membrane [prM], and E) and seven non‐structural (NS) proteins (NS1, NS2A, NS2B, NS3, NS4A, NS4B, and NS5) by viral (NS3) and host proteases (Figure 1A). Of the NS proteins, NS5, the largest viral protein, functions as an RNA‐dependent RNA polymerase (RdRp), which synthesizes a complementary minus‐strand RNA template and, in turn, produces many copies of positive‐strand viral genomic RNA. Viral RNA replication takes place in the so‐called replication complexes (RCs) composed of viral RNA and proteins as well as hypothetical cellular proteins, which are formed on the endoplasmic reticulum (ER) membrane. This membrane structure, rearranged by DENV infection, is also the place of viral RNA translation, protein processing, and progeny virion assembly [9]. The immature particles then bud into the ER lumen and are transported through the secretion pathway of the trans‐Golgi network, in which progeny virions undergo a maturation process via a conformational change of prM and E proteins on the virion surface. Finally, mature, infectious virions egress from infected cells by exocytosis (Figure 1B) [5, 8].

Advertisement

3. Cellular factors involved in DENV replication: interaction with viral RNA untranslated regions

As seen in other RNA and DNA viruses, flaviviruses, including DENV, have been shown to utilize biological processes of the host to replicate efficiently in infected cells through the interaction between viral and cellular proteins. Although the biological relevance of many of the host interactors in in vivo replication and the pathogenicity of DENV remain unclear, those virus‐host interactions could serve as attractive targets of antiviral drugs [2]. In addition to the entry receptors, intracellular factors implicated in the replication cycle of DENV have been comprehensively reviewed elsewhere [2, 8, 10, 11]. Meanwhile, it is becoming apparent that various cellular cofactors are recruited into the 5′ and 3′ UTRs of DENV RNA, which results in enhanced virus replication [12].

The DENV 5′ and 3′ UTRs are approximately 100 and 350–700 nucleotides long, respectively (Figure 1A). The primary sequences located within or adjacent to the UTRs have been shown to be essential for virus replication [8]. For instance, complementary sequences, termed CS (cyclization sequence) elements, found in the ORF of capsid and the 3′ UTR and UAR (upstream of the AUG region) elements, found in the 5′ and 3′ UTRs, are reported to physically anneal to mediate DENV genome cyclization. This cyclization is likely to recruit the DENV RdRp at the 5′ end of the viral genome and then likely to facilitate viral RNA amplification [13]. Additionally, DENV UTRs form stable secondary and tertiary structures, and these structural integrities are thought to be important for the regulation of viral RNA synthesis and translation process. Therefore, albeit the sequences are diverse, UTR secondary structures are highly conserved among flaviviruses [8]. Furthermore, these regions have been demonstrated to contain several cellular proteins [12].

An early study that employed an in vitro‐formed nucleoprotein complex of synthesized DENV 3′ UTR and mosquito cell extracts identified some RNA‐binding proteins—the La autoantigen, translation elongation factor‐1α (EF‐1α), and polypyrimidine tract‐binding protein (PTB)—as the DENV UTR‐associated cellular factors in cells [14]. Subsequent studies have also revealed that human La protein binds not only with the DENV 3′ UTR but also with the 5′ UTR [15, 16]. Interestingly, interactions of La protein with viral proteins NS3 and NS5 were shown, suggesting that the La protein is somehow involved in the function of RC in infected cells [15]. The La autoantigen is reported to associate with RNA polymerase III transcription; interestingly, this RNA‐binding protein has also been shown to stimulate the translation of viral and cellular mRNAs by binding with their UTRs [17]. As with mosquito EF‐1α, its human homologue (i.e., EF1A) has been reported to specifically recognize the conserved 3′‐terminal stem‐loop (SL) in the 3′ UTR of West Nile virus (WNV) RNA [18]. Given the colocalization of EF‐1α with the DENV RC in the infected cell [19], it can be suggested that this cellular protein also plays an important role in the function of flaviviral RNA. However, the precise step of flaviviral replication in which EF‐1α is involved remains to be elucidated [19].

PTB is a ubiquitous RNA‐binding protein known to be involved in splicing, polyadenylation, stability, and translation of cellular mRNA [20]. With regard to its role in virus infection, PTB has been shown to bind to the UTRs of picornaviruses [21, 22] and the hepatitis C virus (HCV) [23]; it functions as an internal ribosome entry site, (IRES)‐trans‐acting factor, to activate viral translation [24, 25]. On the other hand, several studies have raised questions about the involvement of PTB in the translation process of those RNA viruses [2628]. Interestingly, both La and PTB were implicated in HCV replication [29]. In the case of DENV infection, PTB is reportedly required for efficient replication. Furthermore, PTB interacted with a DENV protein, NS4, as well as viral RNA, suggesting that PTB associates with DENV RCs [30]. Although it remains unclear whether PTB regulates DENV RNA simplification or the translation process (or another step of virus infection), PTB may function as a molecular chaperone to stabilize the structured viral RNA [30, 31]. In addition to the host factors mentioned above, recent studies using tobramycin RNA aptamer affinity chromatography identified DDX6 (DEAD‐box RNA helicase) and ERI3 (putative 3′–5′ RNA exonuclease) as DENV UTR‐binding cellular proteins that promote DENV replication [32, 33].

The genome of flaviviruses, including DENV, encodes enzymes required for viral RNA synthesis (i.e., RdRp and helicase) and viral protein processing (i.e., protease); however, the viral protein translation process must rely fully on the translational machinery of the host [34], except for the methyltransferase activity conferred by NS5, which adds a type 1 7‐methylguanosine cap to the 5′ terminus of viral RNA. It is, therefore, not surprising that cellular factors are associated with DENV RNA during viral translation [8, 35]. At the initiation step of eukaryotic mRNA translation, a 5′ cap structure is first recognized by a eukaryotic initiation factor, eIF4E, which, in turn, leads to the recruitment of eIF4G. eIF4G serves as a scaffold protein that binds the DEAD‐box RNA helicase eIF4A and also the poly(A)‐binding protein (PABP), resulting in the circularization of mRNA [8]. This complex formation induces the association of the 43S ribosomal subunit through the binding of the eIF3 complex. Finally, the 60S ribosomal subunit joins the initiation complex that enables the elongation process of translation [8]. In the case of DENV, its genomic RNA is 5′ capped but lacks 3′‐end poly(A) sequences [8]. Nevertheless, PABP is shown to interact with DENV 3′ UTR in vitro by recognizing the A‐rich regions upstream of 3′ SL in the 3′ UTR [36]. Indeed, a study using chimeras reported that mRNA based on a cellular (globin) gene and DENV RNA revealed that DENV 3′ UTR exhibits functions similar to those of polyadenylated non‐viral UTRs in enhancing translation rather than RNA stabilization [37]. Thus, it is plausible that as with cellular mRNA, translated DENV RNA also forms a closed‐loop structure via association with host translation factors [38]. In addition, it is demonstrated that DENV RNA may be able to produce proteins by a 5′ cap‐independent translation mechanism in certain cellular situations in which the eIF4E is starved and, thereby, canonical host translation is inhibited [39].

Advertisement

4. Experimental procedures for identifying the cellular factors

In the following sections, we summarize several general approaches to search for the cellular factors involved in virus infections.

4.1. Conventional methods using living cells

4.1.1. Gene expression analysis and proteomic analysis

Studies using a microarray system, quantitative RT‐PCR analysis, and GeneChip analysis have revealed that the expression levels of many mRNAs in host cells are dramatically altered upon DENV infection [4042]. It is highly possible that some of these DENV‐responsible genes and their products function as crucial positive or negative regulators of DENV replication in cells. Indeed, tumour necrosis factor–related apoptosis‐inducing ligand (TRAIL), whose mRNA expression level significantly increased with DENV infection, was found to be a negative regulator of DENV replication [42]. In addition to the analysis of the gene expression profile, the global host protein expression profile upon DENV infection was also investigated by comprehensive proteomic analysis. Conventionally, the total host proteins extracted from infected cells and control cells were separated by two‐dimensional PAGE (2‐D PAGE), and the proteins whose expression levels were altered by the viral infection were picked up, and their amino acid sequences were determined with mass spectrometry (MS) analysis. Previous studies have identified many host factors that respond to DENV infection [43, 44]. More recently, a proteomic analysis based on stable isotope labelling by amino acids in cell culture (SILAC), which overcame the limitations of sensitivity and resolution of 2‐D PAGE, was developed and applied to DENV research [45, 46]. The technologies of high‐throughput gene expression analysis and proteomic analysis are thought to be powerful tools for understanding the global cellular expression profile of host cells upon DENV infection, both at the gene and at the protein levels. From the results of these assays alone, however, it is difficult to distinguish whether DENV‐responsible host factors that are identified are indeed involved in the regulation of viral replication. To understand the roles and functions of these hit factors requires further functional analysis.

4.1.2. RNAi screening

RNA interference (RNAi) is a well‐known approach for identifying novel host factors in virus‐infected cells. When the knockdown of a host gene by RNAi alters the efficiency of viral replication, it is highly possible that this gene works as a host factor of the virus. Currently, commercial siRNA pools that cover most human genes are available, and many genome‐wide comprehensive screenings for many kinds of viruses have been performed thus far [4749]. In studies of flaviviruses, hundreds of host factor proteins involved with the early steps of WNV infection have been identified by silencing more than 20,000 human genes from a small interfering RNA (siRNA) pool [50]. An additional bioinformatic study, followed by the siRNA screening, revealed that many of these hits were involved in the ubiquitin‐proteasome pathway and the ER‐associated degradation (ERAD) pathway, both of which are essential for many steps of viral replication. In addition to host factors in human cells, one study aimed to identify host factors from insect cells, another host of DENV [51]. In that study of DENV‐adapted cells from Drosophila melanogaster and an siRNA pool targeting more than 20,000 genes of Drosophila, the proteins that accelerate or inhibit the replication of DENV in Drosophila cells were identified [51]. Then human analogues of hit genes in Drosophila cells were subsequently silenced in DENV‐infected human cultured cells, and 42 of these were found to be common host factors in human and Drosophila cells [51]. It is noteworthy that RNAi screening can identify host factors that can affect viral replication either by direct interaction with viral components or by indirect interaction, such as the regulation of the IFN pathway.

4.1.3. Yeast two‐hybrid analysis

For researchers attempting to identify host factors that directly interact with viral proteins, yeast two‐hybrid (Y2H) analysis is a common and effective way; numerous host factors of many viruses, including DENV, have been identified using this system [5258]. Taking advantage of good throughput, several large screenings were performed to build DENV host interactomes using human cDNA libraries. One study reported that a Y2H analysis using either partial or full‐length DNA fragments encoding each DENV protein and the human liver cDNA library identified 105 viral‐host interactions; further knockdown experiments using siRNAs revealed that six proteins were essential for the efficient viral replication of DENV [59]. Another study also carried out a Y2H assay using NS3 and NS5 proteins from DENV and other flaviviruses as bait, and 108 human proteins were identified as interacting with NS3 or NS5 or both [60]. In addition, Y2H assays were performed to identify DENV proteins interacting with host factors from human and mosquito cDNA libraries. They identified several common host factors conserved in both humans and mosquitos [61]. These studies identified many host proteins that interact with DENV proteins; however, few of the hits overlap in independent studies. Although this might be caused mainly by the fact that the assay in each study was performed in different conditions and with different cDNA libraries, it is also possible that each study includes many false positives and false negatives.

4.2. Wheat germ–based protein array system

4.2.1. Overview of the protein array technology

As described above, the living cell‐based methods are powerful tools for identifying viral‐host interactions since the assays could be carried out under physiological conditions, at least partially. However, these methods have several disadvantages. First, proteins whose expression levels are quite low or that show cytotoxicity are hard to analyse. Second, immunoprecipitation assays are commonly used to detect the interaction between a protein and a protein; however, the number of interactions that can be detected at one time is limited. Therefore, it is highly possible that many researchers have potentially overlooked important but difficult‐to‐detect interactions in their first screenings using living cells. To solve these problems, we recently developed a novel biochemical screening method based on a wheat germ cell–free protein synthesis system (wheat cell‐free system) and high‐throughput binding assay. The wheat cell‐free system enables the synthesis of various kinds of eukaryotic proteins in a 96‐well format [62]. So far, proteins having several transmembrane domains, relatively large molecular weights, and cytotoxic activity were successfully synthesized by this system [63, 64]. This robust protein synthesis system allows us to establish a “protein array”, from which tens to thousands of recombinant proteins sorted by their functions were arrayed into each well of a 96‐well plate [65]. The protein arrays currently available in our research group are shown in Table 1 [6570]. Based on these protein arrays, we have established a high‐throughput binding assay for identifying proteins directly bound to target proteins from the array. To perform hundreds to tens of thousands of binding reactions, a luminescent‐based binding assay, called AlphaScreen, was employed [65, 66, 71]. This assay is able to use crude translation products of wheat cell‐free synthesis to detect binding reactions by mixing these crude proteins in a 384‐well plate, followed by adding two beads and the antibody for detection. Because of its flexibility, AlphaScreen can be used not only as a binding assay but also as a protein cleavage assay for a viral protease. The principle of the assays is shown in Figure 2.

Figure 2.

A schematic diagram of binding assay and protease assay based on AlphaScreen technology.

Protein arrayOriginNumber of proteinsReferences
Protein kinaseHuman and mouse400Masaki et al. [66], Miyakawa et al. [67], Kudoh et al. [68]
E3 ligaseHuman and mouse250Takahashi et al. [65], Tan et al. [69]
Deubiquitinating enzymeHuman85Unpublished
Single transmembrane proteinHuman and mouse730Unpublished
Auto‐antigen proteinHuman and mouse2100Matsuoka et al. [70]

Table 1.

Wheat cell–free‐based protein arrays currently available in our research group.

4.2.2. Practical applications for viral research

We and other research groups have done several assays using this technology to identify host proteins that interacted with viral proteins. In researching the HCV (a virus related to DENV), several protein kinases were found to bind directly with the HCV non‐structural protein NS5A using the protein kinase array; by phosphorylating NS5A through additional functional analysis using a cultured cell system, one protein kinase, Casein kinase I‐α, was found to regulate viral replication [66]. The protein kinase array was used to identify other host factors that functionally interacted with human immunodeficiency virus (HIV) proteins [67, 68]. In the case of the protease assay, one research article demonstrated that a protease from xenotropic murine leukaemia virus‐related virus (XMRV) and 24 cellular proteins that were target candidates of the viral protease were synthesized; in vitro cleavage assay revealed several novel substrates of XMRV protease [72], indicating the feasibility of using wheat cell–free‐based protein array technology and high‐throughput biochemical assay based on AlphaScreen.

4.2.3. Wheat cell–free protein array for DENV research

Thus far, NS3 and NS5 proteins of DENV have been considered difficult to synthesize as full‐length active recombinant proteins using a conventional protein expression system, such as for Escherichia coli. We previously reported that the wheat cell‐free system successfully synthesized full‐length NS3 and NS5 proteins in a soluble form; some biochemical analyses revealed that both recombinant proteins possessed enzymatic activities [63], indicating the usefulness of the expression system for the preparation of DENV proteins. In addition to these soluble proteins, NS4B, a protein with at least three transmembrane domains, was synthesized with the wheat cell‐free system (Figure 3A). Our preliminary study demonstrated that NS4B was highly ubiquitinated when NS4B was overexpressed in HEK293T cells (Figure 3B), suggesting that the amount of NS4B expressed was regulated in host cells in a ubiquitin/proteasome‐dependent manner. When a protein is ubiquitinated and, subsequently, degraded by the 26S proteasome, E3 is a determinant of the ubiquitination, as E3 specifically binds to the target protein and transfers activated ubiquitin from the ubiquitin conjugation enzyme, E2 [73]. Therefore, we screened NS4B‐binding E3s from the E3 protein array using AlphaScreen, as we recently reported [65]. NS3 was used as control to determine the NS4B‐specific E3s. As shown in Figure 3C, many E3s were found to bind with recombinant NS4B. Currently, additional functional analysis, such as in vitro ubiquitination assay and protein degradation assay in cells, is ongoing.

Figure 3.

Identification of E3 ligases targeting DENV NS4B by wheat cell‐free‐based protein array system. (A) Expression of biotinylated NS4B with wheat cell‐free system. The total translation products of NS4B (T) and supernatant (S) after centrifugation of the total translation product were subjected to SDS‐PAGE, followed by immunoblot analysis using anti‐biotin antibody. M is the molecular marker. (B) Ubiquitination of NS4B. HEK293T cells overexpressing FLAG‐tagged NS4B and HA‐tagged ubiquitin were treated with proteasomal inhibitor MG132, and FLAG‐tagged NS4B was precipitated with anti‐FLAG antibody. The NS4B and ubiquitin were detected by immunoblot analysis using anti‐FLAG antibody and anti‐HA antibody, respectively. (C) AlphaScreen assay to identify the E3s targeting DENV NS3 and NS4B proteins using the E3 protein array. Biotinylated NS3 and NS4B were used as bait, and biotinylated DHFR was used as negative control of NS proteins. The relative luminescent signal was calculated as follows: the value from E3 and NS4B/value from E3 and DHFR.

Advertisement

5. Restriction of DENV infection by cellular inhibitors

As previously mentioned, DENV hijacks the host's biological process for its efficient replication. Meanwhile, it has become apparent that DENV infection can be limited by cellular factors. In this sense, the innate immune response induced by IFN is considered to be the first line of defence against DENV [8]

Generally, RNA viruses that infect target cells are sensed by the pattern recognition receptors (PRRs), which specifically recognize a component of invading viruses. As for DENV infection, the membrane‐bound Toll‐like receptors (TLR3, TLR7, TLR8) and the cytosolic receptors (retinoic acid‐inducible gene I [RIG‐I], melanoma differentiation‐associated gene 5 [MDA5]) are reported to be the PRRs for viral RNA [7477]. These recognitions in turn activate adaptor molecules of the PRR, leading to the activation of a downstream phosphorylation cascade and the subsequent production of IFN and pro‐inflammatory cytokines (Figure 4) [78]. Among the IFNs produced, type I IFNs, including IFNα, IFN‐β, and IFN‐ω, play an important role in antiviral immunity [79]. The type I IFN then binds to its receptors (IFNAR) on neighbouring cells and signals to induce the phosphorylation of signal transducers and the activators of transcriptions 1 and 2 (STAT1 and 2) in cytoplasm. This phosphorylation of STAT1/2 triggers the formation of IFN‐stimulated gene factor 3 (ISGF3) with IFN‐regulatory factor 9 (IRF9). Finally, the ISGF3 complex translocates to the nucleus and acts as a transcription activator for the expression of a number of genes by binding to the IFN‐stimulated response elements (ISREs) on chromosomes (Figure 4) [80].

Figure 4.

Intracellular recognition of RNA viruses and IFN responses (left). PRRs‐sensing virus‐specific nucleic acid structures (ssRNA and dsRNA) activate signalling cascades via phosphorylation of transcription factors (IRF3 and IRF7), which leads to the expression of IFN genes together with NF‐κB (right). Binding of IFNs produced from virus‐infected cells to type I IFN receptor (IFNAR) activates JAK1 and TYK2, which, in turn, phosphorylate STAT1 and STAT2. Phosphorylation of STAT1/2 induces the formation of ISGF3 complex with IRF9, and finally, ISGF3 translocated in nucleus binds to ISREs to induce ISG expressions. ssRNA, single‐stranded RNA; dsRNA, double‐stranded RNA; MAVS, mitochondrial antiviral‐signalling protein; JAK1, Janus kinase 1; and TYK2, tyrosine kinase 2.

As seen above, IFN is considered to be an inducer of the antiviral state, and it has been well demonstrated that actual antiviral effector molecules in the IFN response are a subset of the genes upregulated by IFN, which are called IFN‐stimulated genes (ISGs). Thus far, hundreds of genes have been classified as ISGs, and many of them are reported to have inhibitory effects on divergent families of viruses, including flaviviruses [8083]. Importantly, several ISGs have also been shown to restrict DENV infection, and their suppressive effects are likely to be exerted at the multiple steps of virus replication, including virus entry (IFITMs [84, 85], ADAP2 [86]), viral RNA/protein synthesis (ISG20 [87], viperin [88, 89]), and infectious virion production (tetherin [90], ISG15 [91, 92]). In addition, a comprehensive study using an overexpression of a cDNA library derived from known ISGs demonstrates the involvement of many more ISGs in the restriction of DENV [93]. However, the precise mechanisms of these ISG‐mediated anti‐DENV activities remain unclear, and we will have to wait for future studies for detailed analysis.

Advertisement

6. Identification of RyDEN/C19orf66 as novel anti‐DENV ISG

It has been well demonstrated that a gain‐of‐function (i.e., overexpression) screen of cDNA is a powerful approach to identifying antiviral ISGs [82, 83, 93, 94]. In a recent study, we employed the gain‐of‐function strategy using a pool of cDNA library derived from type 1 IFN‐treated cells to search for ISGs suppressing DENV replication in human cells [95]. In this approach, a human cervical carcinoma cell line, HeLa, was treated with type I IFN (a mixture of human IFN‐α and ‐ω) for 24 hours at a concentration that had been reported to inhibit DENV infection in vitro [96], and mRNA from the IFN‐treated HeLa cells was converted into a library of cDNA and transferred to a HIV‐based lentiviral vector. Then, a human hepatoma cell line, Huh7.5, was exposed to the HIV vector carrying the IFN cDNA library. Huh7.5 cells are highly susceptible to DENV and, therefore, exhibit massive cell death upon DENV infection. Therefore, we expected that if anti‐DENV genes derived from IFN‐treated HeLa cells were introduced into DENV‐permissive Huh7.5 cells, those cells should be non‐permissive and survive DENV‐induced cell death. As anticipated, even with the DENV challenge, surviving cell clones were obtained; subsequent sequencing analysis revealed that approximately half of DENV‐resistant clones harboured an ORF of a gene on chromosome 19, C19orf66, in the integrated HIV vector genome [95]. Since C19orf66 was a previously uncharacterized gene, we named this repressor of yield of DENV (RyDEN).

The ORF of RyDEN (C19orf66) encodes a 291 amino acid protein, and the secondary structure prediction suggested that the RyDEN protein contained a nuclear localization signal (NLS) in the middle region and a nuclear export signal (NES) in the C‐terminal region. Additionally, a characteristic glutamic acid (E)‐rich domain was found in the C‐terminus.

The anti‐DENV activity of RyDEN was confirmed by creating stable cell lines that expressed epitope tag‐fused human hepatoma cell lines, and all DENV serotypes (i.e., DENV‐1–4) were found to be inhibited by RyDEN expressions. In line with the fact that this gene was first identified by the gain‐of‐function approach using an IFN‐derived cDNA library, expressions of RyDEN in various human cell lines were upregulated by IFN treatment to a greater or lesser extent, indicating that RyDEN is a bona fide anti‐DENV ISG. More importantly, when the endogenous expression of RyDEN mRNA was knocked down by RNAi, the suppressive activity of type I IFNs against DENV became less effective, suggesting that RyDEN was a major contributor of the IFN‐mediated anti‐DENV response [95].

One question to ponder: what is the molecular mechanism by which RyDEN suppresses DENV replication? Affinity purification‐mass spectrometry analysis with affinity tag‐fused RyDEN found that RyDEN interacted with two other cellular proteins, poly(A)‐binding protein cytoplasmic 1 (PABPC1) and La motif‐related protein 1 (LARP1). PABPC1 is a member of the PABP family; as described above, this protein bridges the 5′ and 3′ ends of mRNA by binding both the eIF4G and the poly(A) tail, which stimulates the initiation of translation [97]. LARP1 is also an RNA‐binding protein and one of the LARPS, which shares the signature motif with the La autoantigen, called the La motif (LM). Therefore, Larp and La proteins are categorized as being in the same family [98]. Intriguingly, LARP1 is shown to interact with PABP to stimulate the mRNA translation process [99, 100]. Considering the positive effect of PABP on translation [97], the association of PABPC1 and LARP1 with RyDEN suggests that RyDEN might interfere with the translation of DENV RNA by inhibiting PABPC1 and LARP1 functions. Indeed, this speculation was supported by the following findings: (i) an RNAi‐mediated knockdown of PABPC1 and LARP1 significantly reduced the level of DENV replication, (ii) PABPC1 interacted with DENV RNA in infected cells, (iii) the expression of RyDEN suppressed the expression of the reporter protein from a DENV‐based sub‐genomic RNA replicon that lacked structural (C, prM, E) genes, and (iv) the recruitment of RyDEN to the DENV RNA 3’ UTR was enhanced by the presence of PABPC1 in vitro [95]. Therefore, one could envisage that the anti‐DENV activity of RyDEN, an antiviral ISG, is exerted during the translation of viral RNA by associating PABPC1/LARP1 with the 3′ UTR and somehow interfering with the function of the translation factors (Figure 5). Or another possibility is that RyDEN may facilitate the degradation of viral RNA, since PABPC1 and LARP1 have been reported to be involved in eukaryotic mRNA decay as processing body (P‐body) and stress granule (SG) components (Figure 5) [99, 101]. In accordance with this notion, a recent study revealed that RyDEN (also referred to as FLJ11286 or IRAV) was colocalized with cytoplasmic P‐bodies in IFN‐treated cells [102]. Furthermore, it was noteworthy that RyDEN expression limited a diverse range of RNA and DNA viruses [83, 95, 102], indicating that RyDEN is a broad‐spectrum antiviral ISG.

Figure 5.

Possible mechanisms by which RyDEN restricts the function of DENV RNA. RyDEN, whose expression is upregulated by IFN, associates with DENV RNA through interaction with PABPC1 and LARP1 that are required for DENV replication. This association may result in (i) translational suppression or (ii) degradation of viral RNA in the cytoplasm.

Advertisement

7. Concluding remarks

This chapter highlighted the molecular interactions between DENV and host factors, particularly focused on the cellular regulation of DENV replication. As is well known, IFN response is one of the host controls of DENV infection and pathogenesis [96, 103, 104]. Although the effector molecules in the IFN response that actually interfere with virus replication remain fully clarified, profound efforts have been made to identify the IFN‐inducible cellular factors restricting RNA virus replication, including DENV [80]. It should be noted that antagonistic effects of DENV infection on IFN signalling and production, which are mediated by several means using viral factors, are observed [105107]. However, the characterization of anti‐DENV ISGs is to illuminate the “heel of Achilles” of DENV, which will provide the underpinnings for the development of antivirals against dengue.

References

  1. 1. Guzman MG, Harris E. Dengue. Lancet. 2015;385:453-465. doi:10.1016/S0140‐6736(14)60572‐9
  2. 2. Pastorino B, Nougairede A, Wurtz N, Gould E, de Lamballerie X. Role of host cell factors in flavivirus infection: implications for pathogenesis and development of antiviral drugs. Antiviral Research. 2010;87:281-294. doi:10.1016/j.antiviral.2010.04.014
  3. 3. Basarab M, Bowman C, Aarons EJ, Cropley I. Zika virus. BMJ. 2016;352:i1049. doi:10.1136/bmj.i1049
  4. 4. Becker R. Missing link: animal models to study whether Zika causes birth defects. Nature Medicine. 2016;22:225-227. doi:10.1038/nm0316‐225
  5. 5. Diamond MS, Pierson TC. Molecular insight into dengue virus pathogenesis and its implications for disease control. Cell. 2015;162:488-492. doi:10.1016/j.cell.2015.07.005
  6. 6. Villar L, Dayan GH, Arredondo‐Garcia JL, Rivera DM, Cunha R, Deseda C, Reynales H, Costa MS, Morales‐Ramirez JO, Carrasquilla G, Rey LC, Dietze R, Luz K, Rivas E, Miranda Montoya MC, Cortes Supelano M, Zambrano B, Langevin E, Boaz M, Tornieporth N, Saville M, Noriega F, Group CYDS. Efficacy of a tetravalent dengue vaccine in children in Latin America. New England Journal of Medicine. 2015;372:113-123. doi:10.1056/NEJMoa1411037
  7. 7. Perera‐Lecoin M, Meertens L, Carnec X, Amara A. Flavivirus entry receptors: an update. Viruses. 2013;6:69-88. doi:10.3390/v6010069
  8. 8. Clyde K, Kyle JL, Harris E. Recent advances in deciphering viral and host determinants of dengue virus replication and pathogenesis. Journal of Virology. 2006;80:11418-11431. doi:10.1128/JVI.01257‐06
  9. 9. den Boon JA, Diaz A, Ahlquist P. Cytoplasmic viral replication complexes. Cell Host Microbe. 2010;8:77-85. doi:10.1016/j.chom.2010.06.010
  10. 10. Fischl W, Bartenschlager R. Exploitation of cellular pathways by Dengue virus. Current Opinion in Microbiology. 2011;14:470-475. doi:10.1016/j.mib.2011.07.012
  11. 11. Krishnan MN, Garcia‐Blanco MA. Targeting host factors to treat West Nile and dengue viral infections. Viruses. 2014;6:683-708. doi:10.3390/v6020683
  12. 12. Bidet K, Garcia‐Blanco MA. Flaviviral RNAs: weapons and targets in the war between virus and host. Biochemical Journal. 2014;462:215-230. doi:10.1042/BJ20140456
  13. 13. Villordo SM, Gamarnik AV. Genome cyclization as strategy for flavivirus RNA replication. Virus Research. 2009;139:230-239. doi:10.1016/j.virusres.2008.07.016
  14. 14. De Nova‐Ocampo M, Villegas‐Sepulveda N, del Angel RM. Translation elongation factor‐1alpha, La, and PTB interact with the 3′ untranslated region of dengue 4 virus RNA. Virology. 2002;295:337-347. doi:10.1006/viro.2002.1407
  15. 15. Garcia‐Montalvo BM, Medina F, del Angel RM. La protein binds to NS5 and NS3 and to the 5′ and 3′ ends of Dengue 4 virus RNA. Virus Research. 2004;102:141-150. doi:10.1016/j.virusres.2004.01.024
  16. 16. Yocupicio‐Monroy RM, Medina F, Reyes‐del Valle J, del Angel RM. Cellular proteins from human monocytes bind to dengue 4 virus minus‐strand 3′ untranslated region RNA. Journal of Virology. 2003;77:3067-3076.
  17. 17. Bousquet‐Antonelli C, Deragon JM. A comprehensive analysis of the La‐motif protein superfamily. RNA. 2009;15:750-764. doi:10.1261/rna.1478709
  18. 18. Blackwell JL, Brinton MA. Translation elongation factor‐1 alpha interacts with the 3′ stem‐loop region of West Nile virus genomic RNA. Journal of Virology. 1997;71:6433-6444.
  19. 19. Davis WG, Blackwell JL, Shi PY, Brinton MA. Interaction between the cellular protein eEF1A and the 3′‐terminal stem‐loop of West Nile virus genomic RNA facilitates viral minus‐strand RNA synthesis. Journal of Virology. 2007;81:10172-10187. doi:10.1128/JVI.00531‐07
  20. 20. Romanelli MG, Diani E, Lievens PM. New insights into functional roles of the polypyrimidine tract‐binding protein. International Journal of Molecular Sciences. 2013;14:22906-22932. doi:10.3390/ijms141122906
  21. 21. Hellen CU, Witherell GW, Schmid M, Shin SH, Pestova TV, Gil A, Wimmer E. A cytoplasmic 57‐kDa protein that is required for translation of picornavirus RNA by internal ribosomal entry is identical to the nuclear pyrimidine tract‐binding protein. Proceedings of the National Academy of Sciences of the United States of America. 1993;90:7642-7646.
  22. 22. Jang SK, Wimmer E. Cap‐independent translation of encephalomyocarditis virus RNA: structural elements of the internal ribosomal entry site and involvement of a cellular 57‐kD RNA‐binding protein. Genes and Development. 1990;4:1560-1572.
  23. 23. Ito T, Lai MM. An internal polypyrimidine‐tract‐binding protein‐binding site in the hepatitis C virus RNA attenuates translation, which is relieved by the 3′‐untranslated sequence. Virology. 1999;254:288-296. doi:10.1006/viro.1998.9541
  24. 24. Jang SK. Internal initiation: IRES elements of picornaviruses and hepatitis c virus. Virus Research. 2006;119:2-15. doi:10.1016/j.virusres.2005.11.003
  25. 25. Martinez‐Salas E, Pacheco A, Serrano P, Fernandez N. New insights into internal ribosome entry site elements relevant for viral gene expression. Journal of General Virology. 2008;89:611-626. doi:10.1099/vir.0.83426‐0
  26. 26. Aizaki H, Choi KS, Liu M, Li YJ, Lai MM. Polypyrimidine‐tract‐binding protein is a component of the HCV RNA replication complex and necessary for RNA synthesis. Journal of Biomedical Science. 2006;13:469-480. doi:10.1007/s11373‐006‐9088‐4
  27. 27. Choi KS, Huang P, Lai MM. Polypyrimidine‐tract‐binding protein affects transcription but not translation of mouse hepatitis virus RNA. Virology. 2002;303:58-68.
  28. 28. Tischendorf JJ, Beger C, Korf M, Manns MP, Kruger M. Polypyrimidine tract‐binding protein (PTB) inhibits Hepatitis C virus internal ribosome entry site (HCV IRES)‐mediated translation, but does not affect HCV replication. Archives of Virology. 2004;149:1955-1970.
  29. 29. Domitrovich AM, Diebel KW, Ali N, Sarker S, Siddiqui A. Role of La autoantigen and polypyrimidine tract‐binding protein in HCV replication. Virology. 2005;335:72-86. doi:10.1016/j.virol.2005.02.009
  30. 30. Anwar A, Leong KM, Ng ML, Chu JJ, Garcia‐Blanco MA. The polypyrimidine tract‐binding protein is required for efficient dengue virus propagation and associates with the viral replication machinery. Journal of Biological Chemistry. 2009;284:17021-17029. doi:10.1074/jbc.M109.006239
  31. 31. Karakasiliotis I, Chaudhry Y, Roberts LO, Goodfellow IG. Feline calicivirus replication: requirement for polypyrimidine tract‐binding protein is temperature‐dependent. Journal of General Virology. 2006;87:3339-3347. doi:10.1099/vir.0.82153‐0
  32. 32. Ward AM, Bidet K, Yinglin A, Ler SG, Hogue K, Blackstock W, Gunaratne J, Garcia‐Blanco MA. 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 Biology. 2011;8:1173-1186. doi:10.4161/rna.8.6.17836
  33. 33. Ward AM, Calvert ME, Read LR, Kang S, Levitt BE, Dimopoulos G, Bradrick SS, Gunaratne J, Garcia‐Blanco MA. The Golgi associated ERI3 is a flavivirus host factor. Scientific Reports. 2016;6:34379. doi:10.1038/srep34379
  34. 34. Jackson RJ, Hellen CU, Pestova TV. The mechanism of eukaryotic translation initiation and principles of its regulation. Nature Reviews: Molecular Cell Biology. 2010;11:113-127. doi:10.1038/nrm2838
  35. 35. Li MM, MacDonald MR, Rice CM. To translate, or not to translate: viral and host mRNA regulation by interferon‐stimulated genes. Trends in Cell Biology. 2015;25:320-329. doi:10.1016/j.tcb.2015.02.001
  36. 36. Polacek C, Friebe P, Harris E. Poly(A)‐binding protein binds to the non‐polyadenylated 3′ untranslated region of dengue virus and modulates translation efficiency. Journal of General Virology. 2009;90:687-692. doi:10.1099/vir.0.007021‐0
  37. 37. Chiu WW, Kinney RM, Dreher TW. Control of translation by the 5′‐ and 3′‐terminal regions of the dengue virus genome. Journal of Virology. 2005;79:8303-8315. doi:10.1128/JVI.79.13.8303‐8315.2005
  38. 38. Edgil D, Harris E. End‐to‐end communication in the modulation of translation by mammalian RNA viruses. Virus Research. 2006;119:43-51. doi:10.1016/j.virusres.2005.10.012
  39. 39. Edgil D, Polacek C, Harris E. Dengue virus utilizes a novel strategy for translation initiation when cap‐dependent translation is inhibited. Journal of Virology. 2006;80:2976-2986. doi:10.1128/JVI.80.6.2976‐2986.2006
  40. 40. Liew KJ, Chow VT. Differential display RT‐PCR analysis of ECV304 endothelial‐like cells infected with dengue virus type 2 reveals messenger RNA expression profiles of multiple human genes involved in known and novel roles. Journal of Medical Virology. 2004;72:597-609. doi:10.1002/jmv.20034
  41. 41. Warke RV, Xhaja K, Martin KJ, Fournier MF, Shaw SK, Brizuela N, de Bosch N, Lapointe D, Ennis FA, Rothman AL, Bosch I. Dengue virus induces novel changes in gene expression of human umbilical vein endothelial cells. Journal of Virology. 2003;77:11822-11832.
  42. 42. Warke RV, Martin KJ, Giaya K, Shaw SK, Rothman AL, Bosch I. TRAIL is a novel antiviral protein against dengue virus. Journal of Virology. 2008;82:555-564. doi:10.1128/JVI.01694‐06
  43. 43. Higa LM, Caruso MB, Canellas F, Soares MR, Oliveira‐Carvalho AL, Chapeaurouge DA, Almeida PM, Perales J, Zingali RB, Da Poian AT. Secretome of HepG2 cells infected with dengue virus: implications for pathogenesis. Biochimica et Biophysica Acta. 2008;1784:1607-1616. doi:10.1016/j.bbapap.2008.06.015
  44. 44. Kanlaya R, Pattanakitsakul SN, Sinchaikul S, Chen ST, Thongboonkerd V. The ubiquitin‐proteasome pathway is important for dengue virus infection in primary human endothelial cells. Journal of Proteome Research. 2010;9:4960-4971. doi:10.1021/pr100219y
  45. 45. Chiu HC, Hannemann H, Heesom KJ, Matthews DA, Davidson AD. High‐throughput quantitative proteomic analysis of dengue virus type 2 infected A549 cells. PLoS One. 2014;9:e93305. doi:10.1371/journal.pone.0093305
  46. 46. Viktorovskaya OV, Greco TM, Cristea IM, Thompson SR. Identification of RNA binding proteins associated with dengue virus RNA in infected cells reveals temporally distinct host factor requirements. PLoS Neglected Tropical Diseases. 2016;10:e0004921. doi:10.1371/journal.pntd.0004921
  47. 47. Zhou H, Xu M, Huang Q, Gates AT, Zhang XD, Castle JC, Stec E, Ferrer M, Strulovici B, Hazuda DJ, Espeseth AS. Genome‐scale RNAi screen for host factors required for HIV replication. Cell Host Microbe. 2008;4:495-504. doi:10.1016/j.chom.2008.10.004
  48. 48. Brass AL, Dykxhoorn DM, Benita Y, Yan N, Engelman A, Xavier RJ, Lieberman J, Elledge SJ. Identification of host proteins required for HIV infection through a functional genomic screen. Science. 2008;319:921-926. doi:10.1126/science.1152725
  49. 49. Tai AW, Benita Y, Peng LF, Kim SS, Sakamoto N, Xavier RJ, Chung RT. A functional genomic screen identifies cellular cofactors of hepatitis C virus replication. Cell Host Microbe. 2009;5:298-307. doi:10.1016/j.chom.2009.02.001
  50. 50. Krishnan MN, Ng A, Sukumaran B, Gilfoy FD, Uchil PD, Sultana H, Brass AL, Adametz R, Tsui M, Qian F, Montgomery RR, Lev S, Mason PW, Koski RA, Elledge SJ, Xavier RJ, Agaisse H, Fikrig E. RNA interference screen for human genes associated with West Nile virus infection. Nature. 2008;455:242-245. doi:10.1038/nature07207
  51. 51. Sessions OM, Barrows NJ, Souza‐Neto JA, Robinson TJ, Hershey CL, Rodgers MA, Ramirez JL, Dimopoulos G, Yang PL, Pearson JL, Garcia‐Blanco MA. Discovery of insect and human dengue virus host factors. Nature. 2009;458:1047-1050. doi:10.1038/nature07967
  52. 52. Duan X, Lu X, Li J, Liu Y. Novel binding between pre‐membrane protein and vacuolar ATPase is required for efficient dengue virus secretion. Biochemical and Biophysical Research Communications. 2008;373:319-324. doi:10.1016/j.bbrc.2008.06.041
  53. 53. Shapira SD, Gat‐Viks I, Shum BO, Dricot A, de Grace MM, Wu L, Gupta PB, Hao T, Silver SJ, Root DE, Hill DE, Regev A, Hacohen N. A physical and regulatory map of host‐influenza interactions reveals pathways in H1N1 infection. Cell. 2009;139:1255-1267. doi:10.1016/j.cell.2009.12.018
  54. 54. Falcon AM, Fortes P, Marion RM, Beloso A, Ortin J. Interaction of influenza virus NS1 protein and the human homologue of Staufen in vivo and in vitro. Nucleic Acids Research. 1999;27:2241-2247.
  55. 55. Gao L, Tu H, Shi ST, Lee KJ, Asanaka M, Hwang SB, Lai MM. Interaction with a ubiquitin‐like protein enhances the ubiquitination and degradation of hepatitis C virus RNA‐dependent RNA polymerase. Journal of Virology. 2003;77:4149-4159.
  56. 56. Taguwa S, Okamoto T, Abe T, Mori Y, Suzuki T, Moriishi K, Matsuura Y. Human butyrate‐induced transcript 1 interacts with hepatitis C virus NS5A and regulates viral replication. Journal of Virology. 2008;82:2631-2641. doi:10.1128/JVI.02153‐07
  57. 57. Ambrosino C, Palmieri C, Puca A, Trimboli F, Schiavone M, Olimpico F, Ruocco MR, di Leva F, Toriello M, Quinto I, Venuta S, Scala G. Physical and functional interaction of HIV‐1 Tat with E2F‐4, a transcriptional regulator of mammalian cell cycle. Journal of Biological Chemistry. 2002;277:31448-31458. doi:10.1074/jbc.M112398200
  58. 58. Kalpana GV, Marmon S, Wang W, Crabtree GR, Goff SP. Binding and stimulation of HIV‐1 integrase by a human homolog of yeast transcription factor SNF5. Science. 1994;266:2002-2006.
  59. 59. Khadka S, Vangeloff AD, Zhang C, Siddavatam P, Heaton NS, Wang L, Sengupta R, Sahasrabudhe S, Randall G, Gribskov M, Kuhn RJ, Perera R, LaCount DJ. A physical interaction network of dengue virus and human proteins. Molecular & Cellular Proteomics. 2011;10:M111.012187. doi:10.1074/mcp.M111.012187
  60. 60. Le Breton M, Meyniel‐Schicklin L, Deloire A, Coutard B, Canard B, de Lamballerie X, Andre P, Rabourdin‐Combe C, Lotteau V, Davoust N. Flavivirus NS3 and NS5 proteins interaction network: a high‐throughput yeast two‐hybrid screen. BMC Microbiology. 2011;11:234. doi:10.1186/1471‐2180‐11‐234
  61. 61. Mairiang D, Zhang H, Sodja A, Murali T, Suriyaphol P, Malasit P, Limjindaporn T, Finley RL, Jr. Identification of new protein interactions between dengue fever virus and its hosts, human and mosquito. PLoS One. 2013;8:e53535. doi:10.1371/journal.pone.0053535
  62. 62. Sawasaki T, Ogasawara T, Morishita R, Endo Y. A cell‐free protein synthesis system for high‐throughput proteomics. Proceedings of the National Academy of Sciences of the United States of America. 2002;99:14652-14657. doi:10.1073/pnas.232580399
  63. 63. Takahashi H, Takahashi C, Moreland NJ, Chang YT, Sawasaki T, Ryo A, Vasudevan SG, Suzuki Y, Yamamoto N. Establishment of a robust dengue virus NS3‐NS5 binding assay for identification of protein‐protein interaction inhibitors. Antiviral Research. 2012;96:305-314. doi:10.1016/j.antiviral.2012.09.023
  64. 64. Takeda H, Ogasawara T, Ozawa T, Muraguchi A, Jih PJ, Morishita R, Uchigashima M, Watanabe M, Fujimoto T, Iwasaki T, Endo Y, Sawasaki T. Production of monoclonal antibodies against GPCR using cell‐free synthesized GPCR antigen and biotinylated liposome‐based interaction assay. Scientific Reports. 2015;5:11333. doi:10.1038/srep11333
  65. 65. Takahashi H, Uematsu A, Yamanaka S, Imamura M, Nakajima T, Doi K, Yasuoka S, Takahashi C, Takeda H, Sawasaki T. Establishment of a wheat cell‐free synthesized protein array containing 250 human and mouse E3 ubiquitin ligases to identify novel interaction between E3 ligases and substrate proteins. PLoS One. 2016;11:e0156718. doi:10.1371/journal.pone.0156718
  66. 66. Masaki T, Matsunaga S, Takahashi H, Nakashima K, Kimura Y, Ito M, Matsuda M, Murayama A, Kato T, Hirano H, Endo Y, Lemon SM, Wakita T, Sawasaki T, Suzuki T. Involvement of hepatitis C virus NS5A hyperphosphorylation mediated by casein kinase I‐alpha in infectious virus production. Journal of Virology. 2014;88:7541-7555. doi:10.1128/JVI.03170‐13
  67. 67. Miyakawa K, Sawasaki T, Matsunaga S, Tokarev A, Quinn G, Kimura H, Nomaguchi M, Adachi A, Yamamoto N, Guatelli J, Ryo A. Interferon‐induced SCYL2 limits release of HIV‐1 by triggering PP2A‐mediated dephosphorylation of the viral protein Vpu. Science Signaling 2012;5:ra73. doi:10.1126/scisignal.2003212
  68. 68. Kudoh A, Takahama S, Sawasaki T, Ode H, Yokoyama M, Okayama A, Ishikawa A, Miyakawa K, Matsunaga S, Kimura H, Sugiura W, Sato H, Hirano H, Ohno S, Yamamoto N, Ryo A. The phosphorylation of HIV‐1 Gag by atypical protein kinase C facilitates viral infectivity by promoting Vpr incorporation into virions. Retrovirology. 2014;11:9. doi:10.1186/1742‐4690‐11‐9
  69. 69. Tan BH, Suzuki Y, Takahashi H, Ying PH, Takahashi C, Han Q, Chin WX, Chao SH, Sawasaki T, Yamamoto N, Suzuki Y. Identification of RFPL3 protein as a novel E3 ubiquitin ligase modulating the integration activity of human immunodeficiency virus, type 1 preintegration complex using a microtiter plate‐based assay. Journal of Biological Chemistry. 2014;289:26368-26382. doi:10.1074/jbc.M114.561662
  70. 70. Matsuoka K, Komori H, Nose M, Endo Y, Sawasaki T. Simple screening method for autoantigen proteins using the N‐terminal biotinylated protein library produced by wheat cell‐free synthesis. Journal of Proteome Research. 2010;9:4264-4273. doi:10.1021/pr9010553
  71. 71. Ullman EF, Kirakossian H, Singh S, Wu ZP, Irvin BR, Pease JS, Switchenko AC, Irvine JD, Dafforn A, Skold CN, et al. Luminescent oxygen channeling immunoassay: measurement of particle binding kinetics by chemiluminescence. Proceedings of the National Academy of Sciences of the United States of America. 1994;91:5426-5430.
  72. 72. Matsunaga S, Sawasaki T, Ode H, Morishita R, Furukawa A, Sakuma R, Sugiura W, Sato H, Katahira M, Takaori‐Kondo A, Yamamoto N, Ryo A. Molecular and enzymatic characterization of XMRV protease by a cell‐free proteolytic analysis. Journal of Proteomics. 2012;75:4863-4873. doi:10.1016/j.jprot.2012.05.047
  73. 73. Hershko A, Ciechanover A. The ubiquitin system for protein degradation. Annual Review of Biochemistry. 1992;61:761-807. doi:10.1146/annurev.bi.61.070192.003553
  74. 74. Loo YM, Fornek J, Crochet N, Bajwa G, Perwitasari O, Martinez‐Sobrido L, Akira S, Gill MA, Garcia‐Sastre A, Katze MG, Gale M, Jr. Distinct RIG‐I and MDA5 signaling by RNA viruses in innate immunity. Journal of Virology. 2008;82:335-345. doi:10.1128/JVI.01080‐07
  75. 75. Nasirudeen AM, Wong HH, Thien P, Xu S, Lam KP, Liu DX. RIG‐I, MDA5 and TLR3 synergistically play an important role in restriction of dengue virus infection. PLoS Neglected Tropical Diseases. 2011;5:e926. doi:10.1371/journal.pntd.0000926
  76. 76. Tsai YT, Chang SY, Lee CN, Kao CL. Human TLR3 recognizes dengue virus and modulates viral replication in vitro. Cellular Microbiology. 2009;11:604-615. doi:10.1111/j.1462‐5822.2008.01277.x
  77. 77. Liang Z, Wu S, Li Y, He L, Wu M, Jiang L, Feng L, Zhang P, Huang X. Activation of toll‐like receptor 3 impairs the dengue virus serotype 2 replication through induction of IFN‐beta in cultured hepatoma cells. PLoS One. 2011;6:e23346. doi:10.1371/journal.pone.0023346
  78. 78. Takeuchi O, Akira S. Pattern recognition receptors and inflammation. Cell. 2010;140:805-820. doi:10.1016/j.cell.2010.01.022
  79. 79. Honda K, Takaoka A, Taniguchi T. Type I interferon [corrected] gene induction by the interferon regulatory factor family of transcription factors. Immunity. 2006;25:349-360. doi:10.1016/j.immuni.2006.08.009
  80. 80. Schneider WM, Chevillotte MD, Rice CM. Interferon‐stimulated genes: a complex web of host defenses. Annual Review of Immunology. 2014;32:513-545. doi:10.1146/annurev‐immunol‐032713‐120231
  81. 81. Sen GC. Viruses and interferons. Annual Review of Microbiology. 2001;55:255-281. doi:10.1146/annurev.micro.55.1.255
  82. 82. Kane M, Zang TM, Rihn SJ, Zhang F, Kueck T, Alim M, Schoggins J, Rice CM, Wilson SJ, Bieniasz PD. Identification of interferon‐stimulated genes with antiretroviral activity. Cell Host Microbe. 2016;20:392-405. doi:10.1016/j.chom.2016.08.005
  83. 83. Schoggins JW, Wilson SJ, Panis M, Murphy MY, Jones CT, Bieniasz P, Rice CM. A diverse range of gene products are effectors of the type I interferon antiviral response. Nature. 2011;472:481-485. doi:10.1038/nature09907
  84. 84. Brass AL, Huang IC, Benita Y, John SP, Krishnan MN, Feeley EM, Ryan BJ, Weyer JL, van der Weyden L, Fikrig E, Adams DJ, Xavier RJ, Farzan M, Elledge SJ. The IFITM proteins mediate cellular resistance to influenza A H1N1 virus, West Nile virus, and dengue virus. Cell. 2009;139:1243-1254. doi:10.1016/j.cell.2009.12.017
  85. 85. Chan YK, Huang IC, Farzan M. IFITM proteins restrict antibody‐dependent enhancement of dengue virus infection. PLoS One. 2012;7:e34508. doi:10.1371/journal.pone.0034508
  86. 86. Shu Q, Lennemann NJ, Sarkar SN, Sadovsky Y, Coyne CB. ADAP2 is an interferon stimulated gene that restricts RNA virus entry. PLoS Pathogens. 2015;11:e1005150. doi:10.1371/journal.ppat.1005150
  87. 87. Jiang D, Weidner JM, Qing M, Pan XB, Guo H, Xu C, Zhang X, Birk A, Chang J, Shi PY, Block TM, Guo JT. Identification of five interferon‐induced cellular proteins that inhibit West Nile virus and dengue virus infections. Journal of Virology. 2010;84:8332-8341. doi:10.1128/JVI.02199‐09
  88. 88. Helbig KJ, Carr JM, Calvert JK, Wati S, Clarke JN, Eyre NS, Narayana SK, Fiches GN, McCartney EM, Beard MR. Viperin is induced following dengue virus type‐2 (DENV‐2) infection and has anti‐viral actions requiring the C‐terminal end of viperin. PLoS Neglected Tropical Diseases. 2013;7:e2178. doi:10.1371/journal.pntd.0002178
  89. 89. Fink J, Gu F, Ling L, Tolfvenstam T, Olfat F, Chin KC, Aw P, George J, Kuznetsov VA, Schreiber M, Vasudevan SG, Hibberd ML. Host gene expression profiling of dengue virus infection in cell lines and patients. PLoS Neglected Tropical Diseases. 2007;1:e86. doi:10.1371/journal.pntd.0000086
  90. 90. Pan XB, Han JC, Cong X, Wei L. BST2/tetherin inhibits dengue virus release from human hepatoma cells. PLoS One. 2012;7:e51033. doi:10.1371/journal.pone.0051033
  91. 91. Hishiki T, Han Q, Arimoto K, Shimotohno K, Igarashi T, Vasudevan SG, Suzuki Y, Yamamoto N. Interferon‐mediated ISG15 conjugation restricts dengue virus 2 replication. Biochemical and Biophysical Research Communications. 2014;448:95-100. doi:10.1016/j.bbrc.2014.04.081
  92. 92. Dai J, Pan W, Wang P. ISG15 facilitates cellular antiviral response to dengue and West Nile virus infection in vitro. Virology Journal. 2011;8:468. doi:10.1186/1743‐422X‐8‐468
  93. 93. Schoggins JW, Dorner M, Feulner M, Imanaka N, Murphy MY, Ploss A, Rice CM. Dengue reporter viruses reveal viral dynamics in interferon receptor‐deficient mice and sensitivity to interferon effectors in vitro. Proceedings of the National Academy of Sciences of the United States of America. 2012;109:14610-14615. doi:10.1073/pnas.1212379109
  94. 94. Yi Z, Sperzel L, Nurnberger C, Bredenbeek PJ, Lubick KJ, Best SM, Stoyanov CT, Law LM, Yuan Z, Rice CM, MacDonald MR. Identification and characterization of the host protein DNAJC14 as a broadly active flavivirus replication modulator. PLoS Pathogens. 2011;7:e1001255. doi:10.1371/journal.ppat.1001255
  95. 95. Suzuki Y, Chin WX, Han Q, Ichiyama K, Lee CH, Eyo ZW, Ebina H, Takahashi H, Takahashi C, Tan BH, Hishiki T, Ohba K, Matsuyama T, Koyanagi Y, Tan YJ, Sawasaki T, Chu JJ, Vasudevan SG, Sano K, Yamamoto N. Characterization of RyDEN (C19orf66) as an interferon‐stimulated cellular inhibitor against dengue virus replication. PLoS Pathogens. 2016;12:e1005357. doi:10.1371/journal.ppat.1005357
  96. 96. Diamond MS, Roberts TG, Edgil D, Lu B, Ernst J, Harris E. Modulation of dengue virus infection in human cells by alpha, beta, and gamma interferons. Journal of Virology. 2000;74:4957-4966.
  97. 97. Mangus DA, Evans MC, Jacobson A. Poly(A)‐binding proteins: multifunctional scaffolds for the post‐transcriptional control of gene expression. Genome Biology. 2003;4:223. doi:10.1186/gb‐2003‐4‐7‐223
  98. 98. Wolin SL, Cedervall T. The La protein. Annual Review of Biochemistry. 2002;71:375-403. doi:10.1146/annurev.biochem.71.090501.150003
  99. 99. Merret R, Descombin J, Juan YT, Favory JJ, Carpentier MC, Chaparro C, Charng YY, Deragon JM, Bousquet‐Antonelli C. XRN4 and LARP1 are required for a heat‐triggered mRNA decay pathway involved in plant acclimation and survival during thermal stress. Cell Reports. 2013;5:1279-1293. doi:10.1016/j.celrep.2013.11.019
  100. 100. Tcherkezian J, Cargnello M, Romeo Y, Huttlin EL, Lavoie G, Gygi SP, Roux PP. Proteomic analysis of cap‐dependent translation identifies LARP1 as a key regulator of 5'TOP mRNA translation. Genes and Development. 2014;28:357-371. doi:10.1101/gad.231407.113
  101. 101. Kozlov G, Safaee N, Rosenauer A, Gehring K. Structural basis of binding of P‐body‐associated proteins GW182 and ataxin‐2 by the Mlle domain of poly(A)‐binding protein. Journal of Biological Chemistry. 2010;285:13599-13606. doi:10.1074/jbc.M109.089540
  102. 102. Balinsky CA, Schmeisser H, Wells AI, Ganesan S, Jin T, Singh K, Zoon KC. IRAV (FLJ11286), an interferon stimulated gene with antiviral activity against dengue virus, interacts with MOV10. Journal of Virology. 2016. doi:10.1128/JVI.01606‐16
  103. 103. Kurane I, Innis BL, Nimmannitya S, Nisalak A, Meager A, Ennis FA. High levels of interferon alpha in the sera of children with dengue virus infection. American Journal of Tropical Medicine and Hygiene. 1993;48:222-229.
  104. 104. Ajariyakhajorn C, Mammen MP, Jr., Endy TP, Gettayacamin M, Nisalak A, Nimmannitya S, Libraty DH. Randomized, placebo‐controlled trial of non‐pegylated and pegylated forms of recombinant human alpha interferon 2a for suppression of dengue virus viremia in rhesus monkeys. Antimicrobial Agents and Chemotherapy. 2005;49:4508-4514. doi:10.1128/AAC.49.11.4508‐4514.2005
  105. 105. Green AM, Beatty PR, Hadjilaou A, Harris E. Innate immunity to dengue virus infection and subversion of antiviral responses. Journal of Molecular Biology. 2014;426:1148-1160. doi:10.1016/j.jmb.2013.11.023
  106. 106. Manokaran G, Finol E, Wang C, Gunaratne J, Bahl J, Ong EZ, Tan HC, Sessions OM, Ward AM, Gubler DJ, Harris E, Garcia‐Blanco MA, Ooi EE. Dengue subgenomic RNA binds TRIM25 to inhibit interferon expression for epidemiological fitness. Science. 2015;350:217-221. doi:10.1126/science.aab3369
  107. 107. Bidet K, Dadlani D, Garcia‐Blanco MA. G3BP1, G3BP2 and CAPRIN1 are required for translation of interferon stimulated mRNAs and are targeted by a dengue virus non‐coding RNA. PLoS Pathogens. 2014;10:e1004242. doi:10.1371/journal.ppat.1004242

Written By

Hirotaka Takahashi and Youichi Suzuki

Submitted: 13 October 2016 Reviewed: 20 February 2017 Published: 26 July 2017