Next Article in Journal
The Emergence of H7N7 Highly Pathogenic Avian Influenza Virus from Low Pathogenicity Avian Influenza Virus Using an in ovo Embryo Culture Model
Next Article in Special Issue
Components and Architecture of the Rhabdovirus Ribonucleoprotein Complex
Previous Article in Journal
X69R Is a Non-Essential Gene That, When Deleted from African Swine Fever, Does Not Affect Virulence in Swine
Previous Article in Special Issue
How HIV-1 Gag Manipulates Its Host Cell Proteins: A Focus on Interactors of the Nucleocapsid Domain
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Structures and Functions of Viral 5′ Non-Coding Genomic RNA Domain-I in Group-B Enterovirus Infections

1
EA-4684 CardioVir, Faculty of Medicine, University of Reims Champagne-Ardenne (URCA), 51097 Reims, France
2
Laboratoire de Biopathologie, Centre Hospitalier Universitaire Reims, 51097 Reims, France
3
Service de Médecine Interne, Immunologie clinique et maladies infectieuses, Centre Hospitalier Universitaire Reims, 51097 Reims, France
4
Laboratoire de Virologie Médicale et Moléculaire, Centre Hospitalier Universitaire Reims, 51097 Reims, France
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Viruses 2020, 12(9), 919; https://doi.org/10.3390/v12090919
Submission received: 29 July 2020 / Revised: 17 August 2020 / Accepted: 19 August 2020 / Published: 21 August 2020
(This article belongs to the Special Issue Function and Structure of Viral Ribonucleoproteins Complexes)

Abstract

:
Group-B enteroviruses (EV-B) are ubiquitous naked single-stranded positive RNA viral pathogens that are responsible for common acute or persistent human infections. Their genome is composed in the 5′ end by a non-coding region, which is crucial for the initiation of the viral replication and translation processes. RNA domain-I secondary structures can interact with viral or cellular proteins to form viral ribonucleoprotein (RNP) complexes regulating viral genomic replication, whereas RNA domains-II to -VII (internal ribosome entry site, IRES) are known to interact with cellular ribosomal subunits to initiate the viral translation process. Natural 5′ terminally deleted viral forms lacking some genomic RNA domain-I secondary structures have been described in EV-B induced murine or human infections. Recent in vitro studies have evidenced that the loss of some viral RNP complexes in the RNA domain-I can modulate the viral replication and infectivity levels in EV-B infections. Moreover, the disruption of secondary structures of RNA domain-I could impair viral RNA sensing by RIG-I (Retinoic acid inducible gene I) or MDA5 (melanoma differentiation-associated protein 5) receptors, a way to overcome antiviral innate immune response. Overall, natural 5′ terminally deleted viral genomes resulting in the loss of various structures in the RNA domain-I could be major key players of host–cell interactions driving the development of acute or persistent EV-B infections.

1. Introduction

Group-B enteroviruses (EV-B) are common human pathogens responsible for a wide range of pathologies such as myocarditis, pancreatitis, hepatitis, or meningitis resulting from acute infections in children and young adults [1,2]. Some part of acute EV-B infections can evolve toward persistent infections responsible for chronic pathologies such as dilated cardiomyopathy (DCM), type 1 diabetes, and inflammatory myopathy [3,4,5,6,7].
EV-B belong to the Picornaviridae family [8] and the Enterovirus genus is composed of 15 species, classified according to their biological properties and phylogenetic analyses, seven of which are human viruses: EV-A to -D species and three rhinoviruses species (RV-A to -C) (Figure 1) [9,10,11]. EV-B are small naked single-stranded positive RNA viruses of approximately 7400 nucleotides (nt) [12]. Enteroviral genome is functionally divided in four parts (Figure 2A) with a 5′ non-coding region (5′ NCR), two open reading frames (one monocistronic coding region), and a 3′ non-coding region (3′ NCR) [13]. The 5′ NCR is crucial for the initiation of the replication and translation of the viral genome. This enteroviral 5′ NCR is about 750 nt in length and composed of two secondary structure complexes: the first one is involved in the replication of both viral positive- and negative-strands and is organized in one stem “a” and three stem-loops “b”, “c”, and “d” resulting in a 5′ end Cloverleaf-like structure commonly named viral RNA domain-I (Figure 2B). The second one is composed of six stem-loops (domains-II to -VII) forming an internal ribosome entry site (IRES) involved in cap-independent polyprotein translation.
The viral replication process is well-conserved among EV species [14]. Following the viral capsid attachment to various cellular co-receptors or receptors defining the cell tropism, the positive-strand RNA is released from the viral particle into endocytoplasmic vesicles [14]. The capsid-free viral positive single-stranded RNA is directly translated in a cap-independent mRNA structure according to the cellular ribosomes’ recruitment to its large type 1 IRES (domains-II to -VII) into two open reading frames (ORFs). The viral polyprotein is then cis-cleaved by its autocatalytic 2A (2Apro) and 3C proteases (3Cpro) into structural viral proteins 1–4 (VP1–4) and non-structural proteins including the RNA polymerase [15]. The single positive-stranded RNA replicates through a negative-strand intermediate. Generation of the negative-strand RNA is dependent on RNA sequences and secondary structure domain-I interacting with viral and cellular proteins to form viral ribonucleoprotein (RNP) complexes. Enteroviral genomic replication mechanisms result in the efficient synthesis of negative single-strand RNA used as a matrix to then generate positive single-strand RNA.
Enteroviral RNA genomes are known to mutate more easily than DNA viruses, thus leading to mutated and deleted forms of the viral RNA including the 5′-terminally deleted (5′ TD) viral genomes. Natural 5′ TD viral genomes resulting in the loss of different structures ranging from stem “a”, to stem-loop “b” or “c” or part of stem-loop “d” in the 5′ terminal Cloverleaf RNA structure (domain-I) were described in in vitro and in vivo experimental models of EV-B infections. Recently, EV-B populations with 5′ terminal genomic RNA deletions (5′ TD) similar to those described in experimental models were associated with the development of myocarditis or DCM human cases [16,17]. These 5′ TD EV-B populations were characterized as low replicating RNA forms with viral-encoded proteinase activities related with dystrophin cleavage in human cardiac tissues [16]. Through the transfection of synthetic 5′ TD EV-B RNA forms into murine and human cardiac cells, experimental in vitro results evidenced that the loss of some RNP units in viral RNA domain-I could modulate the viral replication and infectivity levels during early phases of EV-B target cell infections [18]. Altogether these results indicate that the loss of some secondary structures of RNA domain-I could significantly impair the viral genomic RNA replication activities resulting in a significant decrease of viral fitness and infectivity levels in target tissues [18].
Reported natural deletions within the 5′ NCR of EV-B forms are known to affect functional secondary-structural elements of the RNA domain-I (Cloverleaf-like structure) [18]. The loss of some secondary structures of RNA domain-I could impair the viral genomic RNA recognition by retinoic acid inducible gene I (RIG-I)-like receptors (RLRs: RIG-I and melanoma differentiation-associated protein 5, MDA5) which are critical cellular sensors for innate immune defense against Picornaviruses [19,20]. To date, the specific contributions of RIG-I or MDA5 to the anti-EV-B innate immune response remain to be investigated in various infected cells. Through transfection of synthetic complete or various 5′ TD EV-B RNA forms into cultured human cardiac cells, it was demonstrated that natural deletions into the 5′ RNA domain-I of EV-B can modify the RNA secondary-structural elements recognized by cytoplasmic RLRs such as RIG-I or MDA5 and modulate type I interferon (IFN) activation pathway induction [17]. Interestingly, similar natural 5′ terminally deletions capable of modulating activation type I IFN response were characterized in patients with acute myocarditis. Overall, these findings evidenced that natural 5′ TD RNA forms in EV-B populations can modulate type I IFN induction, a way to overcome antiviral innate immune response and to promote persistent infection in EV-B human target tissues such as heart and pancreas.
The present article reviews the importance of secondary structural elements of 5′ terminal EV-B RNA domain-I for the formation of viral RNP complexes regulating viral genome replication and also for the activation of type I IFN signaling pathway by RLRs (RIG-I, MAD5) sensors. The present review deciphers functional consequences of the natural loss of some viral 5′ terminal RNA domain-I secondary structures on the viral RNA replication and on the type I IFN pathway activation in EV-B infected human cells. Overall, natural 5′ TD viral genomes resulting in the loss of different structures in the RNA domain-I could be major key players of host–cell interactions driving the development of acute or persistent infections in human target cells.

2. Functional Organization of 5′ Terminal EV-B RNA Secondary Structures

EV-B replication activities are dependent on the 5′ NCR of about 750 nt in length divided into two highly conserved and well-known structured functional regions. The first one is a ≈90 nt predicted Cloverleaf-like secondary structure also known as stem-loop-I (SL-I) or domain-I, which is crucial for the initiation of the genomic replication for both viral positive- and negative-strands RNA. Enteroviral RNA domain-I presents a conserved structural pattern among EV-B organized in one stem “a and three stem-loops “b”, “c”, and “d” and commonly named Cloverleaf-like structure or RNA domain-I (Figure 2B,C). The second two-dimensional structured pattern from the 5′ NCR is composed of six stem-loops (SL II–VII) and is an IRES involved in the cap-independent viral polyprotein translation (Figure 2A) [21,22,23,24], the host range [25], the replication efficiency [26,27], the cell tropism, and the virulence [26,28,29]. Two regions of around 40 and 100 nt space, respectively, the Cloverleaf from IRES and then the IRES from the translation initiation codon into the VP4 coding sequence [30] (Figure 2A). Enteroviral genome has a principal ORF encoding a long and single monocistronic autocatalytic polyprotein of ≈250 kDa resulting in viral capsid proteins and non-structural proteins. Moreover, a second upstream ORF (uORF) of the principal has been recently identified and could encode for a protein promoting the viral growth in initial viral replication sites (human epithelial cells) and facilitating viral particle release [13,31]. These two ORFs are flanked by 5′ and 3′ NCRs, also called untranslated regions (UTR). The genomic RNA of picornaviruses is linked to a viral protein genome-linked (VPg) at its 5′ NCR, which acts as a primer during viral RNA synthesis [32,33,34,35]

3. Enteroviral 5′ Terminal Domain-I RNA Sequences Involved in Viral Genomic Replication Through the Formation of Viral Ribonucleoprotein Complexes

Enteroviral RNA secondary structure domain-I interacts with viral and cellular proteins to form viral RNP complexes mediating viral genome replication mechanisms in infected cells (Figure 3). The Cloverleaf structure, first described in polioviruses, is a cis-acting RNA replication element required for initiation of negative- and positive-strand synthesis [36,37,38,39,40,41].
Following the attachment and the decapsidation phases of enteroviral cycle, the viral RNA of positive-polarity is directly translated into the cytoplasm and allows the synthesis of proteins necessary for its own genomic replication and the production of new virions. After translation of viral proteins, the first step of viral genome replication consists in the transcription of the positive-polarity RNA by 3D polymerase (3Dpol) into an RNA of negative polarity (also called antigenomic or replication intermediate), which will be used as a template for the synthesis of new viral RNA of positive-polarity that can be translated in turn (Figure 3).
Negative-strand RNA synthesis involves the formation of a functional RNP complex at the 5′ end of the viral genomic RNA. The host cellular poly(C) binding proteins 1 and 2 (PCBP1/2) and poly(A) binding proteins (PABP) are major key factors of the viral minus-strands RNA synthesis [36]. Briefly to construct this RNP, cellular cytoplasmic PCBP1/2 interact with the stem-loop “b” of the Cloverleaf structure whereas the viral 3CD proteinase (3CDpro) interacts with the stem-loop “d”, while PABP binds to the poly(A) located at the end of the 3′ UTR [42,43] (Figure 3B and Table 1). Moreover, PCBP2 interacts with the spacer 1 at the end of the Cloverleaf-like structure (Figure 3B). The binding of PABP to PCBP1/2 by protein–protein interactions allows viral RNA folding which results in the circularization of the genomic RNA promoting interactions between the 5′ and 3′ ends for picornaviruses negative-strand RNA synthesis [44] (Figure 3C). Moreover, the 3CDpro interacts with the cre, which is a RNA stem-loop element in the 2C coding region [45,46]. This regulating RNA structure enhances 3Dpol uridylylation of the viral protein VPg [38,47], which then acts as a primer in negative- and positive-strand RNA synthesis initiation [45,48]. Following the construction of this viral RNP complex, 3Dpol exercises its RNA-dependent RNA polymerase activity using the uridylylated VPg as a primer of negative-strand RNA synthesis (Figure 3D). The negative-strand synthesis results in a double-strand RNA complex called replication form (Figure 3E). The protein components of the RNP and their roles fully depend on their binding and interaction sites (Figure 3B, Table 1). The preservation of the Cloverleaf, a cis-acting replication element at the 5′ end of the positive genomic RNA strand through strong selection pressure might be necessary to maintain the primary, secondary, or tertiary structures of this genomic region. Moreover, the existence of a conserved RNA structure has been hypothesized allowing for the replication complex to specifically recognize enteroviral genome RNA among cellular mRNAs [49].
The initiation of positive-strand RNA synthesis involves the formation RNP complexes on the replicative form, which unwinds at both ends allowing to the Cloverleaf to interact with cellular and viral proteins. Cellular factors PCBP1/2 and viral 3CDpro interact, respectively, on the stem-loop “b” and the stem-loop “d” of the Cloverleaf-like structure in the 5′ NCR of the positive-strand viral RNA [41,74], while nuclear protein Heterogeneous nuclear ribonucleoprotein C (hnRNP-C) binds to both ends of the negative-strand [69,96] (Figure 3F and Table 1). Interactions at both ends of the viral double-strand RNA complex with proteins prevent base-pairing with complementary sequence and preserve a single-stranded structure at both ends of the replication form [41]. hnRNP-C oligomerization could, by drawing both ends into proximity, induce negative-strand circularization (Figure 3G). This process results in a stabilized tertiary structure of RNP complex, which allows the initiation of the positive-strand synthesis by the 3D polymerase with uridylylated VPg as primer (Figure 3H). By these successive molecular phases, neo-synthetized positive-strands RNA and viral double-strand RNA complexes are synthesized (Figure 3A,E). During positive-strand RNA synthesis, a large excess of viral positive-strand RNA is simultaneously synthesized from only one negative-strand (60–70-fold increased levels).
Interestingly, hnRNP-C interactions with the 3′ end of EV negative-strand RNA promote the viral proteins recruitment to an initiation complex for positive-strand RNA synthesis, increasing the amplification of genomic RNA [69,96]. Interactions between hnRNP-C and the 3′ end of the EV negative-strand RNA could stabilize the partially duplexed RNA in the replication form by binding of uridylylated VPg, which is an essential step for prime elongation by the 3Dpol on the intermediate negative-strand RNA template for positive-strand RNA synthesis [69]. Consequently, lack of full-length ends of negative-strand RNA would induce a decrease in interactions between the 5′ and 3′ ends of negative-strand RNA and ribonucleoproteins resulting in a decreased efficiency of the polymerase to synthesize viral positive-strand RNA [96]. Loss of such negative-stranded viral RNA binding sites might contribute to the significant reduction of positive-strand RNA synthesis and aberrant low (+)/(−) viral RNA ratios as described in murine or human tissues chronically infected by persistent viral populations [16,97]. Moreover, low expression levels of replication complex cellular factors such as hnRNP-C could decrease viral genome replication activity levels of full-length ends of negative-strand viral RNA in various human target tissues and cells [18,98]. Overall, hnRNP-C or other unknown host proteins of RNP complexes formed with the 3′ ends of negative-strand could provide an early replicative advantage of full-length viral RNA forms in specific cells and drive the development of enteroviral human infections [17,18,96]. Identification of new cell host proteins of viral RNP complexes could help in designing new therapeutic strategies either targeting or inhibiting host protein activities involving EV-B RNA replication in human infections.

4. Role of Viral Ribonucleoprotein Complexes in the Regulation of Viral Translation and Genomic Replication Processes

Recruitment by RNA domain-I of various host cell proteins in RNP regulates viral translation and genomic replication processes [99]. The Cloverleaf binds PCBP1/2, which facilitates its interaction with the viral 3CDpro [50] and which, in the case of PCBP2, is required for both translations and viral RNA synthesis initiation in infected cells [51]. Moreover, cleavage of PCBP2 by the 3CDpro contributes to viral translation inhibition [100]. The 5′ NCR of EV-B interacts with hnRNP-K protein which is important for viral replication [64]. Moreover, the Cloverleaf-like structure composed of the initial 1–100 nt region in the 5′ UTR and the IRES of coxsackievirus B3 (CV-B3) are known to interact with Polypyrimidine tract-binding protein (PTB)-associated splicing factor (PSF) [101]. Additionally, the IRES structure of picornaviruses interacts with La, Sam68 (68 kDa Src-associated protein in mitosis), PTB, and SRp20 (Serine/arginine-rich splicing factor) host proteins improving translation [53,102,103,104,105]. On the other hand, interactions of ARE/poly(U)-binding/degradation factor 1 (AUF1), Gemin5, and far-upstream element-binding protein 2 (FBP-2) with the 5′ NCR of picornaviruses negatively regulate viral translation [66,106,107,108]. The functional significance of these various interactions for the progression of enteroviral infection in human cells remains to be investigated.

5. Natural 5′ Terminal Deletions in EV-B RNA Domain-I During Experimental or Human Infections

EV-B are well known for their high genomic plasticity mainly related to the lack of proofreading activity of the 3Dpol resulting in approximately one mutation per 103–105 copied nt [109] in every new genome generated during replication phase [110]. The acquisition of mutations (insertions, substitutions, or deletions) allows the generation of a set of viral sub-populations characterized in quasi-species [111,112]. Among these quasi-species, major EV-B populations were shown to contain 5′ terminal deletions ranging up to 49 nt, increasing in deletion size inside RNA domain-I over viral culture passage or time post-infection in experimental models [113]. Various characterized deletions at the end of the 5′ NCR of EV-B RNA affect the secondary structure, resulting in the loss of stem “a”, stem-loop “b”, “c”, and part of stem-loop “d” of the Cloverleaf structure. The related produced virions could still perform the critical functions of cell infection with low levels of genome replication and viral translation activities [75,114]. Interestingly, these 5′ TD viruses showed long-term viral persistence in vitro and in mice heart or pancreas tissues [113,114].
Similar 5′ NCR deleted viral RNA forms have been reported in EV-B induced cardiac human pathologies as myocarditis and DCM [113,114,115,116]. The first clinical characterization of nucleotide deletions in EV-B 5′ NCR was reported in 2008 in endomyocardial biopsy tissues of a CV-B2 induced fulminant viral myocarditis, demonstrating the presence of viral RNA forms deleted 22–36 nt in the 5′ NCR [115]. Recently using deep sequencing strategies, EV-B RNA forms harboring 5′ terminal deletions up to 50 nt were found either alone or present in majority (≈95%) with minor full-length viral forms (≈5%) in heart tissues of a cohort of adult patients with EV-B induced DCM. While 5′ TD viral forms are known to present major defects in viral RNA replication, in vitro experimental results showed that co-transfection of 95% 5′ TD forms with 5% of full-length genome forms into human cardiomyocytes produced higher viral RNA replication activities than transfection with full-length forms alone. In an interesting way, long-term viral replication after this transfection of the mixture led to a lower viral genome synthesis than with those produced by the full-length form alone [16], suggesting a helper function to the minor (full-length) population preserving a functional negative-strand replication but a defect of positive-strand synthesis without any impairment in viral structural and non-structural protein expression [16].
Described 5′ NCR deleted EV-B RNA forms would be defective viral genomes (DVGs) as previously defined with other RNA viruses such as influenza viruses, Sindbis virus, or Sendai virus [117,118]. DVGs are naturally produced viral RNA genomes [119] that displayed three described forms: deleted, copyback, and snapback [120,121]. These truncated forms are naturally produced in the infected host cell and the presence of a helper viral full-length genome (parental genome) is required to perform efficient viral replication activities. These truncated viral RNA forms will compete and decrease parental form replication levels [122]. Interestingly, these DVG forms can overcome the adaptive immune system that activation might clear parental full-length viral forms [123]. Overall, DVG forms could promote the development of high viral genetic diversity levels resulting in the generation of rapidly evolving major populations of viral quasi-species by comparison with the minor parental full-length viral RNA forms [124]. The dynamics of emergence and the impact of DVG or truncated forms onto the enteroviral replication mechanisms and the activation of the innate immune system during early phases of viral infection remains to be investigated.

6. Impact of Natural 5′ Terminal Deletions in Viral RNA Domain-I on EV-B Replication Activities

Partial deletions of the 5′ RNA domain-I sequences significantly disrupt the binding sites of proteins involved in the EV-B RNA synthesis possibly providing a molecular explanation for the low levels of 5′ TD RNA replication in target cells (Figure 4). Investigations by RNA mobility shift assays indicate that the binding of host protein PCBP2 and viral protein 3CDpro to the 5′ end of various deleted forms of positive-strand RNA was qualitatively conserved for the 5′ TD viruses independently of the size of the deletion. However, the stability or protein binding affinity of the RNP complexes decreased with the increase in size of the deletion [18] (Figure 4D). These recent in vitro findings obtained in cultured murine and human cardiac cells provide a possible explanation for the low levels of RNA replication of deleted strains as consequences of the deletions on the formation of RNP complexes. In addition, these data indicate that these low replicative 5′ TD RNA forms may be EV-B persistent populations similar to those observed in clinical studies on heart tissues of patients with unexplained DCM [18].
Interestingly, experimental genetic restoration of the partially deleted Cloverleaf-like secondary structure of described 5′ TD viruses allowed the re-establishment of the viral replication and cellular pathogenesis evidenced through the induction of cytopathic effect on cultured cells [115]. Disrupting PCBP2 binding to either stem-loop “b” or the C-rich sequence inhibits enteroviral RNA replication identifying PCBP2 cell host protein as one of the main cofactors during viral RNA replication [51,54,55] (Figure 4B,D). Within replication complexes, the presence of at least one PCBP binding site in the 5′ NCR (i.e., stem–loop “b” or C-rich sequence) appeared to be sufficient to support efficient negative-strand synthesis [83,85].
Despite different losses of secondary structures and binding sites of viral or host cell proteins in different known viral RNP complexes, natural 5′ TD forms up to 50 nt show a conservation of the stem-loop “d” structure allowing PCBP2 binding to C-rich sequence spacer 1, which is sufficient to support positive and negative strand synthesis at low levels [75] (Figure 4C,D). An unstable RNP complex requiring higher PCBP2 and 3CDpro concentrations sustains 5′ TD RNA low-level replication. These 5′ terminally deletions from 30 to 49 nt induce in silico the tripartite complex formation (PCBP2/3CDpro/Domain-I) at 50–60% efficiency compared to that observed with the full-length viral RNA [18]. Overall, partial deletions of this RNA domain up to the stem-loop “c” but preserving the stem-loop “d” resulted in significant decrease levels of viral infectivity by the production encapsidated negative-stranded RNA [18]. These deletions disrupt viral and host protein binding modulating the dynamics of enteroviral infection and potentially driving the development of acute or persistent infections in human target cells. In recent clinical investigations, these naturally generated low replicating deleted viral RNA forms were identified as major populations in human acute myocarditis and chronic DCM cases [16]. In cultured human cardiomyocytes, the deleted viral RNA forms can produce viral proteinases as 2Apro whose levels impact on cell host protein synthesis machinery by cleavage of major host cell proteins as eukaryotic translation initiation factor 4G (eIF4G) [16,18,125,126]. Overall, natural 5′ TD viral genomes resulting in the loss of different structures in the RNA domain-I could be major key players of host–cell interactions in EV-B human infections [17].

7. Impact of Natural 5′ Terminal Deletions in EV-B RNA Domain-I on Type I Interferon Signaling Pathway Activation

The innate immune system plays an essential role in pathogen recognition and initiation of the immune response through the recognition of pathogen-associated molecular profiles (Pathogen Associated Molecular Pattern, PAMPs) by its Pattern Recognition Receptors (PRRs). The viral nucleic acids, RNAs, or DNAs are recognized as PAMPs. PRRs are essential elements in the detection and induction of the innate response to pathogens including viruses. Recently various PRRs sensors of immunity were identified as Toll-Like Receptors (TLRs), RLRs including RIG-I, MDA5 and LGP2 (Laboratory of Genetics and Physiology 2), and NOD-Like Receptors (NLRs). RLRs have a conserved domain called CTD (C-Terminal Domain), which is responsible for attachment specificity to viral RNAs, whether they are double-stranded RNAs, single-stranded RNAs, or RNAs with a tri-phosphate group at the 5′ end. This C-terminal domain is a repressor domain (RD) which inactivates the caspase recruitment domains (CARD domains) and whose negative regulatory function has only been observed for the RIG-I and LGP2 receptors. The recognition specificity of virus families makes it possible to characterize RLRs; RIG-I is activated by numerous viruses such as Paramyxoviruses, Orthomyxoviruses, Rhabdoviruses, Flaviviruses, and Hepatitis C virus, while MDA5 mainly recognizes viruses of the Picornaviridae family and more specifically group-B Enteroviruses, including group-B Coxsackieviruses [127,128]. In order to limit the activation of the host innate immune response by genome RNA sequences, enteroviruses developed various strategies of direct interference on the sensing of cytosolic immune receptors as RLRs during the early phase of viral replication.
A reported strategy developed by EV-B to overcome antiviral immune response is to cleave or partially disrupt RLRs MDA5 receptor during EV-B infection by viral proteases activities, therefore impairing the innate immune sensing of viral RNA double-strand forms [127,129,130,131,132] (Figure 5). For poliovirus, CV-B3, and EV-71, the 2Apro activity results in the disruption of MDA5 structures, even though some authors discuss that such MDA5 or mitochondrial antiviral signaling protein (MAVS) cleavage may take place in a proteasome-dependent or activated caspase-3 pathway [129,133,134]. To transfer the signal downstream in the type I IFN signaling pathway, ligand-bound MDA5 (CARD) links to the adaptor molecule MAVS at the mitochondrial membrane (Figure 5). In a similar way, the 3Cpro of EV-B can cleave various intermediate protein factors of the type I IFN pathway [135] including RIG-I receptor (Figure 5) [133,136]. Indeed, 3Cpro activity can mediate the cleavage of key molecules such as transforming growth factor-β activated kinase 1 (TAK1)/TGF-β Activated Kinase 1 (TAB1)/TAB2/TAB3 complex which are known to activate the Nuclear factor kappa B pathway inducing a pro-inflammatory cytokine expression [137] (Figure 5). Moreover, EV-71 3Cpro activity was described as capable to cleave Interferon Regulatory Factor 7 (IRF-7) inhibiting its ability to transactivate IFN-β expression by its phosphorylated form during early phase of viral infection [133] (Figure 5).
Natural truncated forms of up to 50 nt in the 5′ NCR of the EV-B in experimental or clinical infections [16,116] would allow the maintenance of the viral genome in the infected host cell by a modulated activation of the innate immune system [17] (Figure 6). Recently, these 5′ TD EV-B RNA forms showed their ability, alone, to produce in silico, in vitro, and in vivo viral proteins such as the viral 2A proteinase resulting in 2Apro-specific eIF4G cleavage [16]. This suggests that natural 5′ TD forms helped by a full-length RNA virus are capable of replicating their genome and translating their proteins at low levels. These low levels of viral proteinases can lead to the potential cleavage and extinction of viral genome recognition by RLRs. Low replication of 5′ TD EV-B RNA forms associated with low synthesis and activity of 2Apro and 3Cpro inducing cleavage of the MDA5 or RIG-I receptors as well as MAVS intermediate remain to be fully elucidated in experimental in vitro and in vivo models.
EV-B RNA domain-I secondary structures in the 5′ UTR may be critical in the host recognition of the viral genome by RLRs, which is crucial for triggering an host innate immune response [20,133] (Figure 6A). Previous studies demonstrated the importance of some nucleotide variations in the 5′ NCR, showing that deletions or point mutations affect the secondary structures of EV-B RNA domain-I [99]. The loss of one or several secondary structures of RNA domain-I could impair the viral genomic RNA recognition by RLRs (RIG-I or MDA5) immune sensors during the early phase of antiviral innate immune response or could contribute to the virus ability to escape the innate immune response [99] (Figure 6B). This hypothesis is supported by a recent report demonstrating that two major 5′ TD viral RNA population groups were identified in peripheral blood or heart tissue samples of EV-B acute myocarditis cases and that their respective proportions appeared to be positively or negatively correlated with type I IFN levels [17]. To confirm immunomodulatory effects of these 5′ TD RNA forms on the type I IFN pathway induction, synthetic CV-B3/28 RNAs harboring various 5′ terminal full-length or deleted genomic sequences were produced by the authors and transfected into human primary cardiomyocytes cultures [17]. This experimental approach demonstrates that EV-B genomic RNA domain-I possesses essential immunomodulatory secondary-structural elements responsible for IFN-β pathway induction in human cardiomyocytes. Natural 5′ terminally RNA deletions disrupting secondary structures of RNA domain-I could result in a differential qualitative binding of EV-B 5′ TD RNA sequences to the RLRs (RIG-I or MDA5) acting as major innate immune sensors during EV-B acute infections (Figure 6). Modulation of RLRs activation by 5′ terminal RNA structures of EV-B 5′ TD viruses could impair type I IFN pathway activation, thus regulating IFN-β and interferon-stimulated gene (ISG) transcription and translation levels in target cells, a way to overcome antiviral innate immune response [17] (Figure 6).

8. Involvement of Natural 5′ Terminally Deleted EV-B RNA Forms in Persistent Infections

RNA viruses including picornaviruses have been shown to develop persistent infections in human target tissues [120,138,139,140]. Among human picornaviruses, polioviruses and group-B coxsackieviruses have been shown to establish persistent infections in target tissues resulting in chronic pathologies such as type 1 diabetes and DCM [120]. Cellular and molecular mechanisms underlying establishment and maintenance of persistent EV-B infections remain unknown. In vivo, specifically in chronic myocarditis or DCM, EV-B persistent infection has been defined by several associated viral criteria including: a low viral RNA load (<3 log10 genome copies per microgram of total nucleic acids extracted), an absence of detected infectious particle by classical culture assays, a plus- to minus-strand viral RNA ratio <5 and the detection of a viral protein synthesis activity [141]. Emergence and maintenance of persistent EV-B forms during acute infection depend on early host–virus interactions.
Emergence of 5′ TD viral populations has been proposed as a molecular mechanism for enterovirus persistence in vivo [116,140]. The association of minor full-length RNA and major 5′ TD viral populations has been found in mouse models and in human persistent cardiac infections [16,114,116]. Full-length EV-B RNA population could help 5′ TD RNA forms replication by genomic recombination processes, sustaining synthesis of negative-strand RNA but not positive-strand [18,142]. Such a scenario would lead to low viral loads, accumulation of double-strand and minus RNA complexes, and undetectable infectious particles which are hallmarks of viral persistence [140,143].
Maintenance of the viral genome is a key step to establish viral RNA persistence in target cells [19]. 5′ TD RNA forms associated with long-term viral persistence described above are known for low-levels of replication activity and progeny [16]. These truncated viral populations displayed 5′ NCR deletions in RNA domain-I disrupting secondary RNA structures, which could impair viral RNA sensing by RLRs and result in a modulation of type I IFN expression in target cells [17]. Given that 5′ TD viruses are low replicating forms, the generation or maintenance of such variants could be an efficient process for EV-B to develop a chronic persistent infection in human target cells long after the acute infection phase [16,18,144,145,146,147]. The combination of low-level replication and translation activities of naturally produced 5′ TD viruses, associated with the capacity of these RNA forms to modulate type-I IFN response activation pathway could explain long-lasting detection of EV-B genome in some human target tissues such as heart and pancreas.

9. Conclusions and Future Directions

The 5′ NCR of EV-B is involved in key steps of the viral life cycle. The recruitment by RNA domain-I of host cell and viral proteins in RNP complexes is essential for viral translation and genomic replication processes. Natural deletions within the 5′ NCR of EV-B forms affect functional secondary-structural elements of the RNA domain-I. The loss of these secondary structures of RNA Cloverleaf reduces the binding of viral and cellular proteins implicated in EV-B genome replication mechanisms, therefore modulating the viral fitness and infectivity of viral populations. Early emergence of 5′ TD viral populations and their maintenance in infected cells have been recently proposed as a potential molecular mechanism driving the development of EV-B human acute and persistent infections [16,17]. Further research programs will aim to identify unknown host cell proteins recruited in viral RNP complexes implicated in 5′ TD EV-B replication process or potentially acting as restriction factors providing an evolutionary advantage to the 5′ NCR truncated viral RNA forms. A genome wide CRISPR–Cas9 screen in near haploid human HAP1 cells approach could uncover differences between host proteins factors crucial for infection and replication of EV-B full-length or 5′ TD forms in human target cells [148]. Identification of new cell host proteins implicated in viral RNP complexes or unknown host cell restriction factors could be a promising new therapeutic strategy targeting or inhibiting host protein activities involved in either the replication or selection of viral populations with 5′ terminal deletions in viral RNA domain-I.
In addition, to disturb viral genome RNA replication processes, 5’NCR deletions could impair the viral RNA recognition by RLRs, inducing the modulation of type I IFN expression, specifically by regulating IFN-β and ISG transcription and translation levels in target cells [17]. Other unexplored mechanisms developed by EV-B could also overcome antiviral innate immune response. Indeed, post-transcriptional RNA modifications commonly called epitranscriptome, are observed at specific residues of cellular RNA and play a key role in their maturation and their functionality and can determine the sensing of viral RNA sequences by RLRs [149,150,151]. Among the numerous chemical modifications described, the methylation of nucleotides in various nucleotide positions is the most abundant in eukaryotic cells. To date, very few studies have so far been devoted to modifications of viral RNA and despite the identification of certain modifications of RNA present in viruses, little information is known about their roles in infection and spread of viral pathogens. Several modifications of viral RNA appear today fundamental for the regulation of arboviruses (ZIKA and Dengue) and other virus replication and spread [152,153]. The 2′O-Methylation of the viral mRNA cap by West Nile virus evades innate immune response by modulating type I IFN pathway activation [149,154]. However, such post-transcriptional modifications in EV-B RNA and their impact on the activation of type I IFN pathway and more specifically on the viral RNA sensing by RLRs are not yet investigated. Using new detection approaches based on genome-wide high-throughput sequencing, further scientific projects have to map the EV-B RNA post-transcriptional modifications and to investigate their impact on the regulation of innate immune response in the early phase of host cell infection.
Finally, published reports and ongoing scientific investigations highlight the importance of secondary structures of RNA domain-I and related functional RNP complexes in the EV-B infectious cycle and related host cell immune response. Overall, natural 5′ TD viral genomes resulting in the loss of different structures in the RNA domain-I could be major key players of host–cell interactions driving the development of acute or persistent EV-B infections in human target cells.

Author Contributions

Conceptualization, L.A., M.G. and L.H.; methodology, L.A., M.G. and L.H.; formal analysis, L.A., M.G., L.H., A.-L.L., D.C., P.-A.G. and F.B.; investigation, L.A., M.G., L.H., A.-L.L., Y.N., D.C., P.-A.G. and F.B.; writing—original draft preparation, L.A., M.G., L.H., A.-L.L., Y.N., D.C. and F.B.; writing—review and editing, L.A., M.G., L.H., A.-L.L., D.C. and F.B.; visualization, L.A., M.G., L.H., A-L.L., D.C. and F.B.; supervision, L.A.; project administration, L.A.; funding acquisition, L.A., F.B., Y.N. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by grant from the “Fédération Française de Cardiologie” (FFC grant 2019 “Etude CardioVir”) and by a local research grant for emerging research teams from the University of Reims Champagne-Ardenne (URCA), France.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to publish the results.

References

  1. Simonen-Tikka, M.-L.; Hiekka, A.-K.; Klemola, P.; Poussa, T.; Ludvigsson, J.; Korpela, R.; Vaarala, O.; Roivainen, M. Early human enterovirus infections in healthy Swedish children participating in the PRODIA pilot study. J. Med. Virol. 2012, 84, 923–930. [Google Scholar] [CrossRef] [PubMed]
  2. Witsø, E.; Palacios, G.; Cinek, O.; Stene, L.C.; Grinde, B.; Janowitz, D.; Lipkin, W.I.; Rønningen, K.S. High prevalence of human enterovirus a infections in natural circulation of human enteroviruses. J. Clin. Microbiol. 2006, 44, 4095–4100. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Oberste, M.S.; Maher, K.; Kilpatrick, D.R.; Pallansch, M.A. Molecular evolution of the human enteroviruses: Correlation of serotype with VP1 sequence and application to picornavirus classification. J. Virol. 1999, 73, 1941–1948. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Genoni, A.; Canducci, F.; Rossi, A.; Broccolo, F.; Chumakov, K.; Bono, G.; Salerno-Uriarte, J.; Salvatoni, A.; Pugliese, A.; Toniolo, A. Revealing enterovirus infection in chronic human disorders: An integrated diagnostic approach. Sci. Rep. 2017, 7, 5013. [Google Scholar] [CrossRef]
  5. Chia, J.K.S. The role of enterovirus in chronic fatigue syndrome. J. Clin. Pathol. 2005, 58, 1126–1132. [Google Scholar] [CrossRef]
  6. Huber, S.A. Viral Myocarditis and Dilated Cardiomyopathy: Etiology and Pathogenesis. Curr. Pharm. Des. 2016, 22, 408–426. [Google Scholar] [CrossRef]
  7. Nekoua, M.P.; Bertin, A.; Sane, F.; Alidjinou, E.K.; Lobert, D.; Trauet, J.; Hober, C.; Engelmann, I.; Moutairou, K.; Yessoufou, A.; et al. Pancreatic beta cells persistently infected with coxsackievirus B4 are targets of NK cell-mediated cytolytic activity. Cell. Mol. Life Sci. 2020, 77, 179–194. [Google Scholar] [CrossRef]
  8. King, A.M.Q.; Lefkowitz, E.J.; Mushegian, A.R.; Adams, M.J.; Dutilh, B.E.; Gorbalenya, A.E.; Harrach, B.; Harrison, R.L.; Junglen, S.; Knowles, N.J.; et al. Changes to taxonomy and the International Code of Virus Classification and Nomenclature ratified by the International Committee on Taxonomy of Viruses (2018). Arch. Virol. 2018, 163, 2601–2631. [Google Scholar] [CrossRef] [Green Version]
  9. Simmonds, P.; Gorbalenya, A.E.; Harvala, H.; Hovi, T.; Knowles, N.J.; Lindberg, A.M.; Oberste, M.S.; Palmenberg, A.C.; Reuter, G.; Skern, T.; et al. Recommendations for the nomenclature of enteroviruses and rhinoviruses. Arch. Virol 2020. [Google Scholar] [CrossRef] [Green Version]
  10. Zell, R.; Delwart, E.; Gorbalenya, A.E.; Hovi, T.; King, A.M.Q.; Knowles, N.J.; Lindberg, A.M.; Pallansch, M.A.; Palmenberg, A.C.; Reuter, G.; et al. ICTV Virus Taxonomy Profile: Picornaviridae. J. Gen. Virol. 2017, 98, 2421–2422. [Google Scholar] [CrossRef]
  11. Zell, R. Picornaviridae-the ever-growing virus family. Arch. Virol. 2018, 163, 299–317. [Google Scholar] [CrossRef] [PubMed]
  12. Walker, P.J.; Siddell, S.G.; Lefkowitz, E.J.; Mushegian, A.R.; Dempsey, D.M.; Dutilh, B.E.; Harrach, B.; Harrison, R.L.; Hendrickson, R.C.; Junglen, S.; et al. Changes to virus taxonomy and the International Code of Virus Classification and Nomenclature ratified by the International Committee on Taxonomy of Viruses (2019). Arch. Virol. 2019, 164, 2417–2429. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Lulla, V.; Dinan, A.M.; Hosmillo, M.; Chaudhry, Y.; Sherry, L.; Irigoyen, N.; Nayak, K.M.; Stonehouse, N.J.; Zilbauer, M.; Goodfellow, I.; et al. An upstream protein-coding region in enteroviruses modulates virus infection in gut epithelial cells. Nat. Microbiol 2019, 4, 280–292. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Fields, B.N.; Knipe, D.M.; Howley, P.M.; Griffin, D.E. Fields Virology, 4th ed.; Wolters Kluwer Health/Lippincott Williams & Wilkins: Philadelphia, PA, USA, 2001; Chapter 23. [Google Scholar]
  15. Badorff, C.; Lee, G.H.; Lamphear, B.J.; Martone, M.E.; Campbell, K.P.; Rhoads, R.E.; Knowlton, K.U. Enteroviral protease 2A cleaves dystrophin: Evidence of cytoskeletal disruption in an acquired cardiomyopathy. Nat. Med. 1999, 5, 320–326. [Google Scholar] [CrossRef]
  16. Bouin, A.; Gretteau, P.-A.; Wehbe, M.; Renois, F.; N’Guyen, Y.; Leveque, N.; Vu, M.N.; Tracy, S.; Chapman, N.M.; Bruneval, P.; et al. Enterovirus Persistence in Cardiac Cells of Patients Suffering from Idiopathic Dilated Cardiomyopathy is Linked to 5′ Terminal Genomic RNA-Deleted Viral Populations With Viral-Encoded Proteinase Activities. Circulation 2019. [Google Scholar] [CrossRef]
  17. Glenet, M.; N’Guyen, Y.; Mirand, A.; Henquell, C.; Lebreil, A.-L.; Berri, F.; Bani-Sadr, F.; Lina, B.; Schuffenecker, I.; Andreoletti, L.; et al. Major 5′terminally deleted enterovirus populations modulate type I IFN response in acute myocarditis patients and in human cultured cardiomyocytes. Sci. Rep. 2020, 10, 11947. [Google Scholar] [CrossRef]
  18. Lévêque, N.; Garcia, M.; Bouin, A.; Nguyen, J.H.C.; Tran, G.P.; Andreoletti, L.; Semler, B.L. Functional Consequences of RNA 5′-Terminal Deletions on Coxsackievirus B3 RNA Replication and Ribonucleoprotein Complex Formation. J. Virol. 2017, 91. [Google Scholar] [CrossRef] [Green Version]
  19. Ullmer, W.; Semler, B.L. Diverse Strategies Used by Picornaviruses to Escape Host RNA Decay Pathways. Viruses 2016, 8. [Google Scholar] [CrossRef] [Green Version]
  20. Feng, Q.; Langereis, M.A.; Olagnier, D.; Chiang, C.; van de Winkel, R.; van Essen, P.; Zoll, J.; Hiscott, J.; van Kuppeveld, F.J.M. Coxsackievirus cloverleaf RNA containing a 5′ triphosphate triggers an antiviral response via RIG-I activation. PLoS ONE 2014, 9, e95927. [Google Scholar] [CrossRef]
  21. Bailey, J.M.; Tapprich, W.E. Structure of the 5′ Nontranslated Region of the Coxsackievirus B3 Genome: Chemical Modification and Comparative Sequence Analysis. J. Virol. 2007, 81, 650–668. [Google Scholar] [CrossRef] [Green Version]
  22. Jang, S.K.; Kräusslich, H.G.; Nicklin, M.J.; Duke, G.M.; Palmenberg, A.C.; Wimmer, E. A segment of the 5′ nontranslated region of encephalomyocarditis virus RNA directs internal entry of ribosomes during in vitro translation. J. Virol. 1988, 62, 2636–2643. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Pelletier, J.; Sonenberg, N. Internal initiation of translation of eukaryotic mRNA directed by a sequence derived from poliovirus RNA. Nature 1988, 334, 320–325. [Google Scholar] [CrossRef] [PubMed]
  24. Trono, D.; Pelletier, J.; Sonenberg, N.; Baltimore, D. Translation in mammalian cells of a gene linked to the poliovirus 5′ noncoding region. Science 1988, 241, 445–448. [Google Scholar] [CrossRef] [PubMed]
  25. Shiroki, K.; Ishii, T.; Aoki, T.; Ota, Y.; Yang, W.X.; Komatsu, T.; Ami, Y.; Arita, M.; Abe, S.; Hashizume, S.; et al. Host range phenotype induced by mutations in the internal ribosomal entry site of poliovirus RNA. J. Virol. 1997, 71, 1–8. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Li, Z.; Liu, X.; Wang, S.; Li, J.; Hou, M.; Liu, G.; Zhang, W.; Yu, X.-F. Identification of a nucleotide in 5′ untranslated region contributing to virus replication and virulence of Coxsackievirus A16. Sci. Rep. 2016, 6, 20839. [Google Scholar] [CrossRef]
  27. Willian, S.; Tracy, S.; Chapman, N.; Leser, S.; Romero, J.; Shapiro, B.; Currey, K. Mutations in a conserved enteroviral RNA oligonucleotide sequence affect positive strand viral RNA synthesis. Arch. Virol. 2000, 145, 2061–2086. [Google Scholar] [CrossRef]
  28. Dunn, J.J.; Bradrick, S.S.; Chapman, N.M.; Tracy, S.M.; Romero, J.R. The stem loop II within the 5′ nontranslated region of clinical coxsackievirus B3 genomes determines cardiovirulence phenotype in a murine model. J. Infect. Dis. 2003, 187, 1552–1561. [Google Scholar] [CrossRef] [Green Version]
  29. Prusa, J.; Missak, J.; Kittrell, J.; Evans, J.J.; Tapprich, W.E. Major alteration in coxsackievirus B3 genomic RNA structure distinguishes a virulent strain from an avirulent strain. Nucleic Acids Res. 2014, 42, 10112–10121. [Google Scholar] [CrossRef] [Green Version]
  30. Knipe, D.M.; Howley, P.M. Fields Virology, 6th ed.; Wolters Kluwer Health/Lippincott Williams & Wilkins: Philadelphia, PA, USA, 2013; Chapter 16; pp. 453–489. [Google Scholar]
  31. Guo, H.; Li, Y.; Liu, G.; Jiang, Y.; Shen, S.; Bi, R.; Huang, H.; Cheng, T.; Wang, C.; Wei, W. A second open reading frame in human enterovirus determines viral replication in intestinal epithelial cells. Nat. Commun 2019, 10, 4066. [Google Scholar] [CrossRef] [Green Version]
  32. Andino, R.; Böddeker, N.; Silvera, D.; Gamarnik, A.V. Intracellular determinants of picornavirus replication. Trends Microbiol. 1999, 7, 76–82. [Google Scholar] [CrossRef]
  33. Cheung, P.K.; Yuan, J.; Zhang, H.M.; Chau, D.; Yanagawa, B.; Suarez, A.; McManus, B.; Yang, D. Specific interactions of mouse organ proteins with the 5′untranslated region of coxsackievirus B3: Potential determinants of viral tissue tropism. J. Med Virol. 2005, 77, 414–424. [Google Scholar] [CrossRef] [PubMed]
  34. Holmes, A.C.; Semler, B.L. Picornaviruses and RNA Metabolism: Local and Global Effects of Infection. J. Virol. 2019, 93. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Witwer, C.; Rauscher, S.; Hofacker, I.L.; Stadler, P.F. Conserved RNA secondary structures in Picornaviridae genomes. Nucleic Acids Res. 2001, 29, 5079–5089. [Google Scholar] [CrossRef] [PubMed]
  36. Andino, R.; Rieckhof, G.E.; Baltimore, D. A functional ribonucleoprotein complex forms around the 5′ end of poliovirus RNA. Cell 1990, 63, 369–380. [Google Scholar] [CrossRef]
  37. Rohll, J.B.; Percy, N.; Ley, R.; Evans, D.J.; Almond, J.W.; Barclay, W.S. The 5′-untranslated regions of picornavirus RNAs contain independent functional domains essential for RNA replication and translation. J. Virol. 1994, 68, 4384–4391. [Google Scholar] [CrossRef] [Green Version]
  38. Rieder, E.; Paul, A.V.; Kim, D.W.; van Boom, J.H.; Wimmer, E. Genetic and biochemical studies of poliovirus cis-acting replication element cre in relation to VPg uridylylation. J. Virol. 2000, 74, 10371–10380. [Google Scholar] [CrossRef] [Green Version]
  39. Barton, D.J.; O’Donnell, B.J.; Flanegan, J.B. 5′ cloverleaf in poliovirus RNA is a cis-acting replication element required for negative-strand synthesis. EMBO J. 2001, 20, 1439–1448. [Google Scholar] [CrossRef]
  40. Lyons, T.; Murray, K.E.; Roberts, A.W.; Barton, D.J. Poliovirus 5′-terminal cloverleaf RNA is required in cis for VPg uridylylation and the initiation of negative-strand RNA synthesis. J. Virol. 2001, 75, 10696–10708. [Google Scholar] [CrossRef] [Green Version]
  41. Vogt, D.A.; Andino, R. An RNA Element at the 5′-End of the Poliovirus Genome Functions as a General Promoter for RNA Synthesis. PLoS Pathog. 2010, 6. [Google Scholar] [CrossRef] [Green Version]
  42. Meredith, J.M.; Rohll, J.B.; Almond, J.W.; Evans, D.J. Similar interactions of the poliovirus and rhinovirus 3D polymerases with the 3′ untranslated region of rhinovirus 14. J. Virol. 1999, 73, 9952–9958. [Google Scholar] [CrossRef] [Green Version]
  43. Rohll, J.B.; Moon, D.H.; Evans, D.J.; Almond, J.W. The 3′ untranslated region of picornavirus RNA: Features required for efficient genome replication. J. Virol. 1995, 69, 7835–7844. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Herold, J.; Andino, R. Poliovirus RNA Replication Requires Genome Circularization through a Protein–Protein Bridge. Mol. Cell 2001, 7, 581–591. [Google Scholar] [CrossRef]
  45. van Ooij, M.J.M.; Vogt, D.A.; Paul, A.; Castro, C.; Kuijpers, J.; van Kuppeveld, F.J.M.; Cameron, C.E.; Wimmer, E.; Andino, R.; Melchers, W.J.G. Structural and functional characterization of the coxsackievirus B3 CRE(2C): Role of CRE(2C) in negative- and positive-strand RNA synthesis. J. Gen. Virol. 2006, 87, 103–113. [Google Scholar] [CrossRef] [PubMed]
  46. Murray, K.E.; Barton, D.J. Poliovirus CRE-dependent VPg uridylylation is required for positive-strand RNA synthesis but not for negative-strand RNA synthesis. J. Virol. 2003, 77, 4739–4750. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Goodfellow, I.; Chaudhry, Y.; Richardson, A.; Meredith, J.; Almond, J.W.; Barclay, W.; Evans, D.J. Identification of a cis-acting replication element within the poliovirus coding region. J. Virol. 2000, 74, 4590–4600. [Google Scholar] [CrossRef]
  48. Morasco, B.J.; Sharma, N.; Parilla, J.; Flanegan, J.B. Poliovirus cre(2C)-dependent synthesis of VPgpUpU is required for positive- but not negative-strand RNA synthesis. J. Virol. 2003, 77, 5136–5144. [Google Scholar] [CrossRef] [Green Version]
  49. Brown, D.M.; Cornell, C.T.; Tran, G.P.; Nguyen, J.H.C.; Semler, B.L. An authentic 3′ noncoding region is necessary for efficient poliovirus replication. J. Virol. 2005, 79, 11962–11973. [Google Scholar] [CrossRef] [Green Version]
  50. Gamarnik, A.V.; Andino, R. Two functional complexes formed by KH domain containing proteins with the 5′ noncoding region of poliovirus RNA. RNA 1997, 3, 882–892. [Google Scholar]
  51. Walter, B.L.; Parsley, T.B.; Ehrenfeld, E.; Semler, B.L. Distinct poly(rC) binding protein KH domain determinants for poliovirus translation initiation and viral RNA replication. J. Virol. 2002, 76, 12008–12022. [Google Scholar] [CrossRef] [Green Version]
  52. Murray, K.E.; Roberts, A.W.; Barton, D.J. Poly(rC) binding proteins mediate poliovirus mRNA stability. RNA 2001, 7, 1126–1141. [Google Scholar] [CrossRef] [Green Version]
  53. Bedard, K.M.; Daijogo, S.; Semler, B.L. A nucleo-cytoplasmic SR protein functions in viral IRES-mediated translation initiation. EMBO J. 2007, 26, 459–467. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Spear, A.; Sharma, N.; Flanegan, J.B. Protein-RNA tethering: The role of poly(C) binding protein 2 in poliovirus RNA replication. Virology 2008, 374, 280–291. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Toyoda, H.; Franco, D.; Fujita, K.; Paul, A.V.; Wimmer, E. Replication of poliovirus requires binding of the poly(rC) binding protein to the cloverleaf as well as to the adjacent C-rich spacer sequence between the cloverleaf and the internal ribosomal entry site. J. Virol. 2007, 81, 10017–10028. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Parsley, T.B.; Towner, J.S.; Blyn, L.B.; Ehrenfeld, E.; Semler, B.L. Poly (rC) binding protein 2 forms a ternary complex with the 5′-terminal sequences of poliovirus RNA and the viral 3CD proteinase. RNA 1997, 3, 1124–1134. [Google Scholar]
  57. Perera, R.; Daijogo, S.; Walter, B.L.; Nguyen, J.H.C.; Semler, B.L. Cellular Protein Modification by Poliovirus: The Two Faces of Poly(rC)-Binding Protein. J. Virol 2007, 81, 8919–8932. [Google Scholar] [CrossRef] [Green Version]
  58. Sean, P.; Nguyen, J.H.C.; Semler, B.L. The linker domain of poly(rC) binding protein 2 is a major determinant in poliovirus cap-independent translation. Virology 2008, 378, 243–253. [Google Scholar] [CrossRef] [Green Version]
  59. Zell, R.; Ihle, Y.; Seitz, S.; Gündel, U.; Wutzler, P.; Görlach, M. Poly(rC)-binding protein 2 interacts with the oligo(rC) tract of coxsackievirus B3. Biochem. Biophys. Res. Commun. 2008, 366, 917–921. [Google Scholar] [CrossRef]
  60. Zell, R.; Ihle, Y.; Effenberger, M.; Seitz, S.; Wutzler, P.; Görlach, M. Interaction of poly(rC)-binding protein 2 domains KH1 and KH3 with coxsackievirus RNA. Biochem. Biophys. Res. Commun. 2008, 377, 500–503. [Google Scholar] [CrossRef]
  61. Harris, K.S.; Xiang, W.; Alexander, L.; Lane, W.S.; Paul, A.V.; Wimmer, E. Interaction of poliovirus polypeptide 3CDpro with the 5′ and 3′ termini of the poliovirus genome. Identification of viral and cellular cofactors needed for efficient binding. J. Biol. Chem. 1994, 269, 27004–27014. [Google Scholar]
  62. Cheung, P.; Zhang, M.; Yuan, J.; Chau, D.; Yanagawa, B.; McManus, B.; Yang, D. Specific interactions of HeLa cell proteins with Coxsackievirus B3 RNA: La autoantigen binds differentially to multiple sites within the 5′ untranslated region. Virus Res. 2002, 90, 23–36. [Google Scholar] [CrossRef]
  63. Cheung, P.; Lim, T.; Yuan, J.; Zhang, M.; Chau, D.; McManus, B.; Yang, D. Specific interaction of HeLa cell proteins with coxsackievirus B3 3′UTR: La autoantigen binds the 3′ and 5′UTR independently of the poly(A) tail. Cell. Microbiol. 2007, 9, 1705–1715. [Google Scholar] [CrossRef] [PubMed]
  64. Lin, J.-Y.; Li, M.-L.; Huang, P.-N.; Chien, K.-Y.; Horng, J.-T.; Shih, S.-R. Heterogeneous nuclear ribonuclear protein K interacts with the enterovirus 71 5′ untranslated region and participates in virus replication. J. Gen. Virol. 2008, 89, 2540–2549. [Google Scholar] [CrossRef] [PubMed]
  65. Shih, S.-R.; Stollar, V.; Li, M.-L. Host Factors in Enterovirus 71 Replication. J. Virol. 2011, 85, 9658–9666. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Lin, J.-Y.; Shih, S.-R.; Pan, M.; Li, C.; Lue, C.-F.; Stollar, V.; Li, M.-L. hnRNP A1 interacts with the 5′ untranslated regions of enterovirus 71 and Sindbis virus RNA and is required for viral replication. J. Virol. 2009, 83, 6106–6114. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Levengood, J.D.; Tolbert, M.; Li, M.-L.; Tolbert, B.S. High-affinity interaction of hnRNP A1 with conserved RNA structural elements is required for translation and replication of enterovirus 71. RNA Biol 2013, 10, 1136–1145. [Google Scholar] [CrossRef] [PubMed]
  68. Li, Q.; Zheng, Z.; Liu, Y.; Zhang, Z.; Liu, Q.; Meng, J.; Ke, X.; Hu, Q.; Wang, H. 2C Proteins of Enteroviruses Suppress IKKβ Phosphorylation by Recruiting Protein Phosphatase 1. J. Virol. 2016, 90, 5141–5151. [Google Scholar] [CrossRef] [Green Version]
  69. Brunner, J.E.; Nguyen, J.H.C.; Roehl, H.H.; Ho, T.V.; Swiderek, K.M.; Semler, B.L. Functional interaction of heterogeneous nuclear ribonucleoprotein C with poliovirus RNA synthesis initiation complexes. J. Virol. 2005, 79, 3254–3266. [Google Scholar] [CrossRef] [Green Version]
  70. Dreyfuss, G.; Matunis, M.J.; Piñol-Roma, S.; Burd, C.G. hnRNP proteins and the biogenesis of mRNA. Annu. Rev. Biochem. 1993, 62, 289–321. [Google Scholar] [CrossRef]
  71. Roehl, H.H.; Semler, B.L. Poliovirus infection enhances the formation of two ribonucleoprotein complexes at the 3′ end of viral negative-strand RNA. J. Virol. 1995, 69, 2954–2961. [Google Scholar] [CrossRef] [Green Version]
  72. Svitkin, Y.V.; Costa-Mattioli, M.; Herdy, B.; Perreault, S.; Sonenberg, N. Stimulation of picornavirus replication by the poly(A) tail in a cell-free extract is largely independent of the poly(A) binding protein (PABP). RNA 2007, 13, 2330–2340. [Google Scholar] [CrossRef] [Green Version]
  73. Molla, A.; Harris, K.S.; Paul, A.V.; Shin, S.H.; Mugavero, J.; Wimmer, E. Stimulation of poliovirus proteinase 3Cpro-related proteolysis by the genome-linked protein VPg and its precursor 3AB. J. Biol. Chem. 1994, 269, 27015–27020. [Google Scholar] [PubMed]
  74. Zell, R.; Sidigi, K.; Bucci, E.; Stelzner, A.; Görlach, M. Determinants of the recognition of enteroviral cloverleaf RNA by coxsackievirus B3 proteinase 3C. RNA 2002, 8, 188–201. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Sharma, N.; Ogram, S.A.; Morasco, B.J.; Spear, A.; Chapman, N.M.; Flanegan, J.B. Functional role of the 5′terminal cloverleaf in Coxsackievirus RNA replication. Virology 2009, 393, 238–249. [Google Scholar] [CrossRef] [Green Version]
  76. Claridge, J.K.; Headey, S.J.; Chow, J.Y.H.; Schwalbe, M.; Edwards, P.J.; Jeffries, C.M.; Venugopal, H.; Trewhella, J.; Pascal, S.M. A picornaviral loop-to-loop replication complex. J. Struct. Biol. 2009, 166, 251–262. [Google Scholar] [CrossRef] [PubMed]
  77. Headey, S.J.; Huang, H.; Claridge, J.K.; Soares, G.A.; Dutta, K.; Schwalbe, M.; Yang, D.; Pascal, S.M. NMR structure of stem-loop D from human rhinovirus-14. RNA 2007, 13, 351–360. [Google Scholar] [CrossRef] [Green Version]
  78. Ohlenschläger, O.; Wöhnert, J.; Bucci, E.; Seitz, S.; Häfner, S.; Ramachandran, R.; Zell, R.; Görlach, M. The structure of the stemloop D subdomain of coxsackievirus B3 cloverleaf RNA and its interaction with the proteinase 3C. Structure 2004, 12, 237–248. [Google Scholar] [CrossRef]
  79. Pathak, H.B.; Arnold, J.J.; Wiegand, P.N.; Hargittai, M.R.S.; Cameron, C.E. Picornavirus genome replication: Assembly and organization of the VPg uridylylation ribonucleoprotein (initiation) complex. J. Biol. Chem. 2007, 282, 16202–16213. [Google Scholar] [CrossRef] [Green Version]
  80. Prostova, M.A.; Deviatkin, A.A.; Tcelykh, I.O.; Lukashev, A.N.; Gmyl, A.P. Independent evolution of tetraloop in enterovirus oriL replicative element and its putative binding partners in virus protein 3C. PeerJ 2017, 5, e3896. [Google Scholar] [CrossRef]
  81. Rieder, E.; Xiang, W.; Paul, A.; Wimmer, E. Analysis of the cloverleaf element in a human rhinovirus type 14/poliovirus chimera: Correlation of subdomain D structure, ternary protein complex formation and virus replication. J. Gen. Virol. 2003, 84, 2203–2216. [Google Scholar] [CrossRef]
  82. Sun, D.; Chen, S.; Cheng, A.; Wang, M. Roles of the Picornaviral 3C Proteinase in the Viral Life Cycle and Host Cells. Viruses 2016, 8, 82. [Google Scholar] [CrossRef] [Green Version]
  83. Gamarnik, A.V.; Andino, R. Switch from translation to RNA replication in a positive-stranded RNA virus. Genes Dev. 1998, 12, 2293–2304. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Cornell, C.T.; Semler, B.L. Subdomain specific functions of the RNA polymerase region of poliovirus 3CD polypeptide. Virology 2002, 298, 200–213. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Gamarnik, A.V.; Andino, R. Interactions of viral protein 3CD and poly(rC) binding protein with the 5′ untranslated region of the poliovirus genome. J. Virol. 2000, 74, 2219–2226. [Google Scholar] [CrossRef] [Green Version]
  86. Prostova, M.A.; Gmyl, A.P.; Bakhmutov, D.V.; Shishova, A.A.; Khitrina, E.V.; Kolesnikova, M.S.; Serebryakova, M.V.; Isaeva, O.V.; Agol, V.I. Mutational robustness and resilience of a replicative cis-element of RNA virus: Promiscuity, limitations, relevance. RNA Biol. 2015, 12, 1338–1354. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Xiang, W.; Harris, K.S.; Alexander, L.; Wimmer, E. Interaction between the 5′-terminal cloverleaf and 3AB/3CDpro of poliovirus is essential for RNA replication. J. Virol. 1995, 69, 3658–3667. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  88. Yang, Y.; Rijnbrand, R.; Watowich, S.; Lemon, S.M. Genetic evidence for an interaction between a picornaviral cis-acting RNA replication element and 3CD protein. J. Biol. Chem. 2004, 279, 12659–12667. [Google Scholar] [CrossRef] [Green Version]
  89. Lin, J.-Y.; Chen, T.-C.; Weng, K.-F.; Chang, S.-C.; Chen, L.-L.; Shih, S.-R. Viral and host proteins involved in picornavirus life cycle. J. Biomed. Sci. 2009, 16, 103. [Google Scholar] [CrossRef] [Green Version]
  90. Kräusslich, H.G.; Nicklin, M.J.; Toyoda, H.; Etchison, D.; Wimmer, E. Poliovirus proteinase 2A induces cleavage of eucaryotic initiation factor 4F polypeptide p220. J. Virol. 1987, 61, 2711–2718. [Google Scholar] [CrossRef] [Green Version]
  91. Joachims, M.; Van Breugel, P.C.; Lloyd, R.E. Cleavage of poly(A)-binding protein by enterovirus proteases concurrent with inhibition of translation in vitro. J. Virol. 1999, 73, 718–727. [Google Scholar] [CrossRef] [Green Version]
  92. Mu, Z.; Wang, B.; Zhang, X.; Gao, X.; Qin, B.; Zhao, Z.; Cui, S. Crystal structure of 2A proteinase from hand, foot and mouth disease virus. J. Mol. Biol. 2013, 425, 4530–4543. [Google Scholar] [CrossRef]
  93. Yang, C.-H.; Li, H.-C.; Jiang, J.-G.; Hsu, C.-F.; Wang, Y.-J.; Lai, M.-J.; Juang, Y.-L.; Lo, S.-Y. Enterovirus type 71 2A protease functions as a transcriptional activator in yeast. J. Biomed. Sci. 2010, 17, 65. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Banerjee, R.; Tsai, W.; Kim, W.; Dasgupta, A. Interaction of poliovirus-encoded 2C/2BC polypeptides with the 3′ terminus negative-strand cloverleaf requires an intact stem-loop b. Virology 2001, 280, 41–51. [Google Scholar] [CrossRef] [Green Version]
  95. Pettersson, R.F.; Ambros, V.; Baltimore, D. Identification of a protein linked to nascent poliovirus RNA and to the polyuridylic acid of negative-strand RNA. J. Virol. 1978, 27, 357–365. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Ertel, K.J.; Brunner, J.E.; Semler, B.L. Mechanistic consequences of hnRNP C binding to both RNA termini of poliovirus negative-strand RNA intermediates. J. Virol. 2010, 84, 4229–4242. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Andréoletti, L.; Hober, D.; Becquart, P.; Belaich, S.; Copin, M.C.; Lambert, V.; Wattré, P. Experimental CVB3-induced chronic myocarditis in two murine strains: Evidence of interrelationships between virus replication and myocardial damage in persistent cardiac infection. J. Med. Virol. 1997, 52, 206–214. [Google Scholar] [CrossRef]
  98. Brunner, J.E.; Ertel, K.J.; Rozovics, J.M.; Semler, B.L. Delayed kinetics of poliovirus RNA synthesis in a human cell line with reduced levels of hnRNP C proteins. Virology 2010, 400, 240–247. [Google Scholar] [CrossRef] [PubMed]
  99. Kloc, A.; Rai, D.K.; Rieder, E. The Roles of Picornavirus Untranslated Regions in Infection and Innate Immunity. Front. Microbiol. 2018, 9, 485. [Google Scholar] [CrossRef]
  100. Chase, A.J.; Daijogo, S.; Semler, B.L. Inhibition of Poliovirus-Induced Cleavage of Cellular Protein PCBP2 Reduces the Levels of Viral RNA Replication. J. Virol. 2014, 88, 3192–3201. [Google Scholar] [CrossRef] [Green Version]
  101. Dave, P.; George, B.; Sharma, D.K.; Das, S. Polypyrimidine tract-binding protein (PTB) and PTB-associated splicing factor in CVB3 infection: An ITAF for an ITAF. Nucleic Acids Res. 2017, 45, 9068–9084. [Google Scholar] [CrossRef] [Green Version]
  102. Hunt, S.L.; Hsuan, J.J.; Totty, N.; Jackson, R.J. unr, a cellular cytoplasmic RNA-binding protein with five cold-shock domains, is required for internal initiation of translation of human rhinovirus RNA. Genes Dev. 1999, 13, 437–448. [Google Scholar] [CrossRef] [Green Version]
  103. Ray, P.S.; Das, S. La autoantigen is required for the internal ribosome entry site-mediated translation of Coxsackievirus B3 RNA. Nucleic Acids Res. 2002, 30, 4500–4508. [Google Scholar] [CrossRef] [PubMed]
  104. Lawrence, P.; Schafer, E.A.; Rieder, E. The nuclear protein Sam68 is cleaved by the FMDV 3C protease redistributing Sam68 to the cytoplasm during FMDV infection of host cells. Virology 2012, 425, 40–52. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Rai, D.K.; Lawrence, P.; Kloc, A.; Schafer, E.; Rieder, E. Analysis of the interaction between host factor Sam68 and viral elements during foot-and-mouth disease virus infections. Virol. J. 2015, 12, 224. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Cathcart, A.L.; Rozovics, J.M.; Semler, B.L. Cellular mRNA decay protein AUF1 negatively regulates enterovirus and human rhinovirus infections. J. Virol. 2013, 87, 10423–10434. [Google Scholar] [CrossRef] [Green Version]
  107. Piñeiro, D.; Fernandez-Chamorro, J.; Francisco-Velilla, R.; Martinez-Salas, E. Gemin5: A Multitasking RNA-Binding Protein Involved in Translation Control. Biomolecules 2015, 5, 528–544. [Google Scholar] [CrossRef] [Green Version]
  108. Francisco-Velilla, R.; Fernandez-Chamorro, J.; Ramajo, J.; Martinez-Salas, E. The RNA-binding protein Gemin5 binds directly to the ribosome and regulates global translation. Nucleic Acids Res. 2016, 44, 8335–8351. [Google Scholar] [CrossRef]
  109. Domingo, E.; Holland, J.J. RNA virus mutations and fitness for survival. Annu. Rev. Microbiol. 1997, 51, 151–178. [Google Scholar] [CrossRef]
  110. Domingo, E.; Martin, V.; Perales, C.; Escarmis, C. Coxsackieviruses and quasispecies theory: Evolution of enteroviruses. Curr. Top. Microbiol. Immunol. 2008, 323, 3–32. [Google Scholar] [CrossRef]
  111. Mandary, M.B.; Masomian, M.; Poh, C.L. Impact of RNA Virus Evolution on Quasispecies Formation and Virulence. Int. J. Mol. Sci. 2019, 20. [Google Scholar] [CrossRef] [Green Version]
  112. Domingo, E.; Sheldon, J.; Perales, C. Viral Quasispecies Evolution. Microbiol. Mol. Biol. Rev. 2012, 76, 159–216. [Google Scholar] [CrossRef] [Green Version]
  113. Chapman, N.M.; Kim, K.S. Persistent coxsackievirus infection: Enterovirus persistence in chronic myocarditis and dilated cardiomyopathy. Curr. Top. Microbiol. Immunol. 2008, 323, 275–292. [Google Scholar] [PubMed]
  114. Kim, K.-S.; Tracy, S.; Tapprich, W.; Bailey, J.; Lee, C.-K.; Kim, K.; Barry, W.H.; Chapman, N.M. 5′-Terminal deletions occur in coxsackievirus B3 during replication in murine hearts and cardiac myocyte cultures and correlate with encapsidation of negative-strand viral RNA. J. Virol. 2005, 79, 7024–7041. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Chapman, N.M.; Kim, K.-S.; Drescher, K.M.; Oka, K.; Tracy, S. 5′ terminal deletions in the genome of a coxsackievirus B2 strain occurred naturally in human heart. Virology 2008, 375, 480–491. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Bouin, A.; Nguyen, Y.; Wehbe, M.; Renois, F.; Fornes, P.; Bani-Sadr, F.; Metz, D.; Andreoletti, L. Major Persistent 5′ Terminally Deleted Coxsackievirus B3 Populations in Human Endomyocardial Tissues. Emerg. Infect. Dis. 2016, 22, 1488–1490. [Google Scholar] [CrossRef] [PubMed]
  117. Xu, J.; Sun, Y.; Li, Y.; Ruthel, G.; Weiss, S.R.; Raj, A.; Beiting, D.; López, C.B. Replication defective viral genomes exploit a cellular pro-survival mechanism to establish paramyxovirus persistence. Nat. Commun. 2017, 8, 799. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  118. Sidhu, M.S.; Crowley, J.; Lowenthal, A.; Karcher, D.; Menonna, J.; Cook, S.; Udem, S.; Dowling, P. Defective measles virus in human subacute sclerosing panencephalitis brain. Virology 1994, 202, 631–641. [Google Scholar] [CrossRef]
  119. Dimmock, N.J.; Easton, A.J. Defective Interfering Influenza Virus RNAs: Time To Reevaluate Their Clinical Potential as Broad-Spectrum Antivirals? J. Virol. 2014, 88, 5217–5227. [Google Scholar] [CrossRef] [Green Version]
  120. Manzoni, T.B.; López, C.B. Defective (interfering) viral genomes re-explored: Impact on antiviral immunity and virus persistence. Future Virol. 2018, 13, 493–503. [Google Scholar] [CrossRef] [Green Version]
  121. Vignuzzi, M.; López, C.B. Defective viral genomes are key drivers of the virus-host interaction. Nat. Microbiol. 2019, 4, 1075–1087. [Google Scholar] [CrossRef]
  122. Poirier, E.Z.; Mounce, B.C.; Rozen-Gagnon, K.; Hooikaas, P.J.; Stapleford, K.A.; Moratorio, G.; Vignuzzi, M. Low-Fidelity Polymerases of Alphaviruses Recombine at Higher Rates To Overproduce Defective Interfering Particles. J. Virol. 2016, 90, 2446–2454. [Google Scholar] [CrossRef] [Green Version]
  123. Tapia, K.; Kim, W.; Sun, Y.; Mercado-López, X.; Dunay, E.; Wise, M.; Adu, M.; López, C.B. Defective Viral Genomes Arising In Vivo Provide Critical Danger Signals for the Triggering of Lung Antiviral Immunity. PLoS Pathog. 2013, 9. [Google Scholar] [CrossRef]
  124. Domingo, E.; Martin, V.; Perales, C.; Grande-Pérez, A.; García-Arriaza, J.; Arias, A. Viruses as quasispecies: Biological implications. Curr. Top. Microbiol. Immunol. 2006, 299, 51–82. [Google Scholar] [CrossRef] [PubMed]
  125. Liu, Z.; Carthy, C.M.; Cheung, P.; Bohunek, L.; Wilson, J.E.; McManus, B.M.; Yang, D. Structural and functional analysis of the 5′ untranslated region of coxsackievirus B3 RNA: In vivo translational and infectivity studies of full-length mutants. Virology 1999, 265, 206–217. [Google Scholar] [CrossRef] [PubMed]
  126. Roehl, H.H.; Parsley, T.B.; Ho, T.V.; Semler, B.L. Processing of a cellular polypeptide by 3CD proteinase is required for poliovirus ribonucleoprotein complex formation. J. Virol. 1997, 71, 578–585. [Google Scholar] [CrossRef] [Green Version]
  127. Kato, H.; Takeuchi, O.; Sato, S.; Yoneyama, M.; Yamamoto, M.; Matsui, K.; Uematsu, S.; Jung, A.; Kawai, T.; Ishii, K.J.; et al. Differential roles of MDA5 and RIG-I helicases in the recognition of RNA viruses. Nature 2006, 441, 101–105. [Google Scholar] [CrossRef] [PubMed]
  128. Loo, Y.-M.; Fornek, J.; Crochet, N.; Bajwa, G.; Perwitasari, O.; Martinez-Sobrido, L.; Akira, S.; Gill, M.A.; García-Sastre, A.; Katze, M.G.; et al. Distinct RIG-I and MDA5 signaling by RNA viruses in innate immunity. J. Virol. 2008, 82, 335–345. [Google Scholar] [CrossRef] [Green Version]
  129. Kuo, R.-L.; Kao, L.-T.; Lin, S.-J.; Wang, R.Y.-L.; Shih, S.-R. MDA5 plays a crucial role in enterovirus 71 RNA-mediated IRF3 activation. PLoS ONE 2013, 8, e63431. [Google Scholar] [CrossRef] [Green Version]
  130. Feng, Q.; Hato, S.V.; Langereis, M.A.; Zoll, J.; Virgen-Slane, R.; Peisley, A.; Hur, S.; Semler, B.L.; van Rij, R.P.; van Kuppeveld, F.J.M. MDA5 detects the double-stranded RNA replicative form in picornavirus-infected cells. Cell Rep. 2012, 2, 1187–1196. [Google Scholar] [CrossRef] [Green Version]
  131. Abe, Y.; Fujii, K.; Nagata, N.; Takeuchi, O.; Akira, S.; Oshiumi, H.; Matsumoto, M.; Seya, T.; Koike, S. The toll-like receptor 3-mediated antiviral response is important for protection against poliovirus infection in poliovirus receptor transgenic mice. J. Virol. 2012, 86, 185–194. [Google Scholar] [CrossRef] [Green Version]
  132. Wang, J.P.; Cerny, A.; Asher, D.R.; Kurt-Jones, E.A.; Bronson, R.T.; Finberg, R.W. MDA5 and MAVS mediate type I interferon responses to coxsackie B virus. J. Virol. 2010, 84, 254–260. [Google Scholar] [CrossRef] [Green Version]
  133. Feng, Q.; Langereis, M.A.; van Kuppeveld, F.J.M. Induction and suppression of innate antiviral responses by picornaviruses. Cytokine Growth Factor Rev. 2014, 25, 577–585. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Barral, P.M.; Morrison, J.M.; Drahos, J.; Gupta, P.; Sarkar, D.; Fisher, P.B.; Racaniello, V.R. MDA-5 is cleaved in poliovirus-infected cells. J. Virol. 2007, 81, 3677–3684. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Mukherjee, A.; Morosky, S.A.; Delorme-Axford, E.; Dybdahl-Sissoko, N.; Oberste, M.S.; Wang, T.; Coyne, C.B. The coxsackievirus B 3C protease cleaves MAVS and TRIF to attenuate host type I interferon and apoptotic signaling. PLoS Pathog. 2011, 7, e1001311. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  136. Barral, P.M.; Sarkar, D.; Fisher, P.B.; Racaniello, V.R. RIG-I is cleaved during picornavirus infection. Virology 2009, 391, 171–176. [Google Scholar] [CrossRef] [Green Version]
  137. Lei, X.; Han, N.; Xiao, X.; Jin, Q.; He, B.; Wang, J. Enterovirus 71 3C inhibits cytokine expression through cleavage of the TAK1/TAB1/TAB2/TAB3 complex. J. Virol. 2014, 88, 9830–9841. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  138. Maggioli, M.F.; Fernandes, M.H.V.; Joshi, L.R.; Sharma, B.; Tweet, M.M.; Noll, J.C.G.; Bauermann, F.V.; Diel, D.G. Persistent Infection and Transmission of Senecavirus A from Carrier Sows to Contact Piglets. J. Virol. 2019, 93. [Google Scholar] [CrossRef] [Green Version]
  139. Reetoo, K.N.; Osman, S.A.; Illavia, S.J.; Cameron-Wilson, C.L.; Banatvala, J.E.; Muir, P. Quantitative analysis of viral RNA kinetics in coxsackievirus B3-induced murine myocarditis: Biphasic pattern of clearance following acute infection, with persistence of residual viral RNA throughout and beyond the inflammatory phase of disease. J. Gen. Virol. 2000, 81, 2755–2762. [Google Scholar] [CrossRef] [Green Version]
  140. Tam, P.E.; Messner, R.P. Molecular mechanisms of coxsackievirus persistence in chronic inflammatory myopathy: Viral RNA persists through formation of a double-stranded complex without associated genomic mutations or evolution. J. Virol. 1999, 73, 10113–10121. [Google Scholar] [CrossRef] [Green Version]
  141. Andréoletti, L.; Bourlet, T.; Moukassa, D.; Rey, L.; Hot, D.; Li, Y.; Lambert, V.; Gosselin, B.; Mosnier, J.F.; Stankowiak, C.; et al. Enteroviruses can persist with or without active viral replication in cardiac tissue of patients with end-stage ischemic or dilated cardiomyopathy. J. Infect. Dis. 2000, 182, 1222–1227. [Google Scholar] [CrossRef] [Green Version]
  142. Lévêque, N.; Renois, F.; Talmud, D.; Nguyen, Y.; Lesaffre, F.; Boulagnon, C.; Bruneval, P.; Fornes, P.; Andréoletti, L. Quantitative genomic and anti-genomic human enterovirus RNA detection in explanted heart tissue samples of patients with end-stage idiopathic dilated cardiomyopathy. J. Clin. Microbiol. 2012. [Google Scholar] [CrossRef] [Green Version]
  143. Weber, F.; Wagner, V.; Rasmussen, S.B.; Hartmann, R.; Paludan, S.R. Double-stranded RNA is produced by positive-strand RNA viruses and DNA viruses but not in detectable amounts by negative-strand RNA viruses. J. Virol. 2006, 80, 5059–5064. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Tracy, S.; Smithee, S.; Alhazmi, A.; Chapman, N. Coxsackievirus can persist in murine pancreas by deletion of 5′ terminal genomic sequences. J. Med. Virol. 2015, 87, 240–247. [Google Scholar] [CrossRef] [PubMed]
  145. Alidjinou, E.K.; Engelmann, I.; Bossu, J.; Villenet, C.; Figeac, M.; Romond, M.-B.; Sané, F.; Hober, D. Persistence of Coxsackievirus B4 in pancreatic ductal-like cells results in cellular and viral changes. Virulence 2017, 8, 1229–1244. [Google Scholar] [CrossRef] [PubMed]
  146. Lietzén, N.; Hirvonen, K.; Honkimaa, A.; Buchacher, T.; Laiho, J.E.; Oikarinen, S.; Mazur, M.A.; Flodström-Tullberg, M.; Dufour, E.; Sioofy-Khojine, A.-B.; et al. Coxsackievirus B Persistence Modifies the Proteome and the Secretome of Pancreatic Ductal Cells. iScience 2019, 19, 340–357. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  147. Jaïdane, H.; Caloone, D.; Lobert, P.-E.; Sane, F.; Dardenne, O.; Naquet, P.; Gharbi, J.; Aouni, M.; Geenen, V.; Hober, D. Persistent infection of thymic epithelial cells with coxsackievirus B4 results in decreased expression of type 2 insulin-like growth factor. J. Virol. 2012, 86, 11151–11162. [Google Scholar] [CrossRef] [Green Version]
  148. Meertens, L.; Hafirassou, M.L.; Couderc, T.; Bonnet-Madin, L.; Kril, V.; Kümmerer, B.M.; Labeau, A.; Brugier, A.; Simon-Loriere, E.; Burlaud-Gaillard, J.; et al. FHL1 is a major host factor for chikungunya virus infection. Nature 2019. [Google Scholar] [CrossRef]
  149. Daffis, S.; Szretter, K.J.; Schriewer, J.; Li, J.; Youn, S.; Errett, J.; Lin, T.-Y.; Schneller, S.; Zust, R.; Dong, H.; et al. 2′-O methylation of the viral mRNA cap evades host restriction by IFIT family members. Nature 2010, 468, 452–456. [Google Scholar] [CrossRef]
  150. Dai, Q.; Moshitch-Moshkovitz, S.; Han, D.; Kol, N.; Amariglio, N.; Rechavi, G.; Dominissini, D.; He, C. Nm-seq maps 2′-O-methylation sites in human mRNA with base precision. Nat. Methods 2017, 14, 695–698. [Google Scholar] [CrossRef]
  151. Züst, R.; Cervantes-Barragan, L.; Habjan, M.; Maier, R.; Neuman, B.W.; Ziebuhr, J.; Szretter, K.J.; Baker, S.C.; Barchet, W.; Diamond, M.S.; et al. Ribose 2′- O -methylation provides a molecular signature for the distinction of self and non-self mRNA dependent on the RNA sensor Mda5. Nat. Immunol. 2011, 12, 137–143. [Google Scholar] [CrossRef] [Green Version]
  152. Gonzales-van Horn, S.R.; Sarnow, P. Making the Mark: The Role of Adenosine Modifications in the Life Cycle of RNA Viruses. Cell Host Microbe 2017, 21, 661–669. [Google Scholar] [CrossRef] [Green Version]
  153. Lichinchi, G.; Zhao, B.S.; Wu, Y.; Lu, Z.; Qin, Y.; He, C.; Rana, T.M. Dynamics of Human and Viral RNA Methylation during Zika Virus Infection. Cell Host Microbe 2016, 20, 666–673. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  154. Szretter, K.J.; Daniels, B.P.; Cho, H.; Gainey, M.D.; Yokoyama, W.M.; Gale, M.; Virgin, H.W.; Klein, R.S.; Sen, G.C.; Diamond, M.S. 2′-O methylation of the viral mRNA cap by West Nile virus evades ifit1-dependent and -independent mechanisms of host restriction in vivo. PLoS Pathog. 2012, 8, e1002698. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. Evolutionary relationship of human enterovirus species. Molecular phylogenetic analysis of human enteroviruses was inferred using the Neighbor-Joining method (Letunic and Bork, 2019). Phylogenetic tree was constructed using complete sequence of enteroviruses aligned by muscle method. The optimal tree with the sum of branch length = 18.13314604 is shown. The tree is drawn to scale, with branch lengths in the same units as those of the evolutionary distances used to infer the phylogenetic tree. The evolutionary distances were computed using the Kimura 2-parameter method and are in the units of the number of base substitutions per site. This analysis involved 128 nucleotide sequences. All ambiguous positions were removed for each sequence pair (pairwise deletion option). There was a total of 8522 positions in the final dataset. Evolutionary analyses were conducted in MEGA X (Kumar et al., 2018). EV-A: Enterovirus A; EV-B: Enterovirus B; EV-C: Enterovirus C; EV-D: Enterovirus D; RV-A: Rhinovirus A; RV-B: Rhinovirus B; RV-C: Rhinovirus C.
Figure 1. Evolutionary relationship of human enterovirus species. Molecular phylogenetic analysis of human enteroviruses was inferred using the Neighbor-Joining method (Letunic and Bork, 2019). Phylogenetic tree was constructed using complete sequence of enteroviruses aligned by muscle method. The optimal tree with the sum of branch length = 18.13314604 is shown. The tree is drawn to scale, with branch lengths in the same units as those of the evolutionary distances used to infer the phylogenetic tree. The evolutionary distances were computed using the Kimura 2-parameter method and are in the units of the number of base substitutions per site. This analysis involved 128 nucleotide sequences. All ambiguous positions were removed for each sequence pair (pairwise deletion option). There was a total of 8522 positions in the final dataset. Evolutionary analyses were conducted in MEGA X (Kumar et al., 2018). EV-A: Enterovirus A; EV-B: Enterovirus B; EV-C: Enterovirus C; EV-D: Enterovirus D; RV-A: Rhinovirus A; RV-B: Rhinovirus B; RV-C: Rhinovirus C.
Viruses 12 00919 g001
Figure 2. Group-B enterovirus genome organization. (A) Enterovirus-B viral genome is composed of two open reading frames (ORFs) which one is a polyprotein coding for capsidic (viral proteins 1 to 4) and non-structural viral proteins such as proteases 2A, 3C, and the 3D polymerase. This long ORF is flanked by two untranslated regions (UTRs): the 5′ UTR is composed of a Cloverleaf-like secondary structure (CL, stem-loop I) responsible for the viral genomic replication separated by a spacer sequence comprising two C-rich clusters from an Internal Ribosome Entry Site element type 1 (IRES, stem-loops II to VII) which has a role in the viral translation. Nucleotide numbers are indicated below the structures. (B,C) Schematic representation of the predicted 5′ end Cloverleaf secondary structures (CL, stem-loop I) of the enteroviral positive strand RNA. The CL is composed of a stem “a” (nucleotides 2–9 with nucleotides 80–87), stem-loop “b” (nucleotides 10–34), stem-loop “c” (nucleotides 35–45), and a stem-loop “d” (nucleotides 46–79).
Figure 2. Group-B enterovirus genome organization. (A) Enterovirus-B viral genome is composed of two open reading frames (ORFs) which one is a polyprotein coding for capsidic (viral proteins 1 to 4) and non-structural viral proteins such as proteases 2A, 3C, and the 3D polymerase. This long ORF is flanked by two untranslated regions (UTRs): the 5′ UTR is composed of a Cloverleaf-like secondary structure (CL, stem-loop I) responsible for the viral genomic replication separated by a spacer sequence comprising two C-rich clusters from an Internal Ribosome Entry Site element type 1 (IRES, stem-loops II to VII) which has a role in the viral translation. Nucleotide numbers are indicated below the structures. (B,C) Schematic representation of the predicted 5′ end Cloverleaf secondary structures (CL, stem-loop I) of the enteroviral positive strand RNA. The CL is composed of a stem “a” (nucleotides 2–9 with nucleotides 80–87), stem-loop “b” (nucleotides 10–34), stem-loop “c” (nucleotides 35–45), and a stem-loop “d” (nucleotides 46–79).
Viruses 12 00919 g002
Figure 3. Viral ribonucleoprotein complexes involved in group-B enterovirus genomic replication process. (A) The enterovirus genome is single-stranded positive RNA with a 5′ non-coding region (5′ NCR) including a Cloverleaf-like structure (RNA domain-I). The genomic RNA is linked to a viral protein genome-linked (VPg) at its 5′-terminus, which acts as a primer during viral RNA synthesis. (B) Host cellular protein Poly(A) binding protein (PABP) interacts with the poly(A) tail in the viral genome 3′ end; Poly(C) Binding Protein (PCBP) 1/2 binds to the stem-loop “b” of the Cloverleaf and PCBP 2 binds to the spacer 1. Viral precursor protein 3CD interacts with stem-loop “d” of the Cloverleaf. (C) Through protein–protein interactions between PCBP 1/2 and PABP, the 5′ ends interact to form a ribonucleoprotein complex (RNP). The circularization of the viral RNA initiates the negative-strand synthesis. (D) The viral 3D polymerase synthesizes the negative strand. (E) The product of the negative-strand synthesis is a double-strand RNA complex called replication form. (F) Viral double-strand RNA complex unwinds at both ends, which enables the two Cloverleaf-like structures to interact with cellular and viral proteins. Nuclear protein, heterogeneous nuclear RNP-C (hnRNP-C), can interact with both ends of the negative-strand. Interactions of viral double-strand RNA complex with cellular and viral proteins allow maintaining a single-stranded structure at both ends of the replication form. (G) hnRNP-C could promote negative-strand circularization by oligomerization. (H) The circularization allows the initiation of the positive-strand synthesis. The 3D polymerase recruitment allows the start of positive-strand synthesis using VPg as primer. This process results in the formation of neo-synthetized positive- strands RNA and viral double-strand RNA complexes.
Figure 3. Viral ribonucleoprotein complexes involved in group-B enterovirus genomic replication process. (A) The enterovirus genome is single-stranded positive RNA with a 5′ non-coding region (5′ NCR) including a Cloverleaf-like structure (RNA domain-I). The genomic RNA is linked to a viral protein genome-linked (VPg) at its 5′-terminus, which acts as a primer during viral RNA synthesis. (B) Host cellular protein Poly(A) binding protein (PABP) interacts with the poly(A) tail in the viral genome 3′ end; Poly(C) Binding Protein (PCBP) 1/2 binds to the stem-loop “b” of the Cloverleaf and PCBP 2 binds to the spacer 1. Viral precursor protein 3CD interacts with stem-loop “d” of the Cloverleaf. (C) Through protein–protein interactions between PCBP 1/2 and PABP, the 5′ ends interact to form a ribonucleoprotein complex (RNP). The circularization of the viral RNA initiates the negative-strand synthesis. (D) The viral 3D polymerase synthesizes the negative strand. (E) The product of the negative-strand synthesis is a double-strand RNA complex called replication form. (F) Viral double-strand RNA complex unwinds at both ends, which enables the two Cloverleaf-like structures to interact with cellular and viral proteins. Nuclear protein, heterogeneous nuclear RNP-C (hnRNP-C), can interact with both ends of the negative-strand. Interactions of viral double-strand RNA complex with cellular and viral proteins allow maintaining a single-stranded structure at both ends of the replication form. (G) hnRNP-C could promote negative-strand circularization by oligomerization. (H) The circularization allows the initiation of the positive-strand synthesis. The 3D polymerase recruitment allows the start of positive-strand synthesis using VPg as primer. This process results in the formation of neo-synthetized positive- strands RNA and viral double-strand RNA complexes.
Viruses 12 00919 g003
Figure 4. Natural 5′ terminal deletions in RNA domain-I disrupt viral ribonucleoprotein complexes involved in group-B enterovirus replication activities. (A) Schematic representation of the complete predicted 5′ end Cloverleaf secondary structures (CL, stem-loop I) of the enteroviral positive-strand RNA. The CL is composed of a stem “a” (nucleotides 2–9 with nucleotides 80–87), stem-loop “b” (nucleotides 10–34), stem-loop “c” (nucleotides 35–45), and a stem-loop “d” (nucleotides 46–79). (B) Host cellular protein PABP interacts with the poly(A) tail in the viral genome 3′ end, PCBP 1/2 bind to the stem-loop “b” of the Cloverleaf and in addition, PCBP 2 binds to the spacer 1 at the end of the Cloverleaf. Viral precursor protein 3CD interacts with stem-loop “d” of the Cloverleaf. (C) Schematic representation of the 5′ terminally deleted Enterovirus RNA with 50 nucleotides deletions in Cloverleaf secondary structures. (D) 5′ terminal deletions in RNA domain-I disrupts the formation of viral ribonucleoprotein complexes: PABP interacts with the 3′ end poly(A) tail, PCBP 2 binds to the spacer 1 at the end of the Cloverleaf, and the precursor 3CD interacts with stem-loop “d” of the Cloverleaf. This loss of interactions with viral protease and cellular factor in stem-loop “b” alters EV-B replication activities.
Figure 4. Natural 5′ terminal deletions in RNA domain-I disrupt viral ribonucleoprotein complexes involved in group-B enterovirus replication activities. (A) Schematic representation of the complete predicted 5′ end Cloverleaf secondary structures (CL, stem-loop I) of the enteroviral positive-strand RNA. The CL is composed of a stem “a” (nucleotides 2–9 with nucleotides 80–87), stem-loop “b” (nucleotides 10–34), stem-loop “c” (nucleotides 35–45), and a stem-loop “d” (nucleotides 46–79). (B) Host cellular protein PABP interacts with the poly(A) tail in the viral genome 3′ end, PCBP 1/2 bind to the stem-loop “b” of the Cloverleaf and in addition, PCBP 2 binds to the spacer 1 at the end of the Cloverleaf. Viral precursor protein 3CD interacts with stem-loop “d” of the Cloverleaf. (C) Schematic representation of the 5′ terminally deleted Enterovirus RNA with 50 nucleotides deletions in Cloverleaf secondary structures. (D) 5′ terminal deletions in RNA domain-I disrupts the formation of viral ribonucleoprotein complexes: PABP interacts with the 3′ end poly(A) tail, PCBP 2 binds to the spacer 1 at the end of the Cloverleaf, and the precursor 3CD interacts with stem-loop “d” of the Cloverleaf. This loss of interactions with viral protease and cellular factor in stem-loop “b” alters EV-B replication activities.
Viruses 12 00919 g004
Figure 5. Group-B enterovirus proteinase activities impair type 1 signaling pathway activation in infected cells. Enterovirus proteinase 3C (3Cpro) and 2Apro are mainly involved in downregulation of type I IFN, pro-inflammatory cytokines at different stages. The interacting cellular signaling molecules with different viral proteins are indicated at each level. (MDA5: melanoma-differentiation-associated protein 5, RIG-I: retinoic acid-inducible gene 1, MAVS: mitochondrial antiviral-signaling protein, TBK1: TANK-binding kinase 1, IRF3/7: Interferon Regulatory Factor 3/7, TRIF: TIR-domain-containing adapter-inducing interferon-β, TLR: toll-like receptors, TRAF6: TNF receptor-associated factor 6, TAK1: transforming growth factor-β activated kinase 1, TAB2/3: TGF-β Activated Kinase 2/3, IFNs: interferons).
Figure 5. Group-B enterovirus proteinase activities impair type 1 signaling pathway activation in infected cells. Enterovirus proteinase 3C (3Cpro) and 2Apro are mainly involved in downregulation of type I IFN, pro-inflammatory cytokines at different stages. The interacting cellular signaling molecules with different viral proteins are indicated at each level. (MDA5: melanoma-differentiation-associated protein 5, RIG-I: retinoic acid-inducible gene 1, MAVS: mitochondrial antiviral-signaling protein, TBK1: TANK-binding kinase 1, IRF3/7: Interferon Regulatory Factor 3/7, TRIF: TIR-domain-containing adapter-inducing interferon-β, TLR: toll-like receptors, TRAF6: TNF receptor-associated factor 6, TAK1: transforming growth factor-β activated kinase 1, TAB2/3: TGF-β Activated Kinase 2/3, IFNs: interferons).
Viruses 12 00919 g005
Figure 6. Natural 5′ terminally deleted group-B enterovirus RNA forms can modulate type I interferon signaling pathway activation. (A) Full-length viral RNA is recognized by cytoplasmic sensors RIG-I or MDA5 in EV-B infections. Signaling through the adaptor protein MAVS leads to IRF3 activation and translocation to the nucleus. These molecules stimulate high level of IFN-α/β and ISG56 production for the development of effective antiviral responses to EV-B infections. (B) According to recent reports, deleted viral genomes were associated with parental complete virus in early phase group-B enterovirus infection. Host cell proteins recruited in viral RNP complexes or potentially acting as a restriction factors could provide an evolutionary advantage to the 5′ NCR truncated viral RNA forms. Deletion in RNA domain-I (5′ terminally deleted viral genomes) could impair the viral genomic RNA recognition by RLRs (RIG-I or MDA5) immune sensors during the early phase of antiviral innate immune response resulting in low level of IFN-α/β and ISG56 production. IFN: interferon; RIG-I: retinoic acid-inducible gene-I; MDA5: melanoma differentiation-associated protein 5; MAVS: mitochondria antiviral-signaling protein; IRF: Interferon Regulatory Factor 3; ISG56: interferon stimulated gene 56; 5′ NCR: 5′ non-coding region.
Figure 6. Natural 5′ terminally deleted group-B enterovirus RNA forms can modulate type I interferon signaling pathway activation. (A) Full-length viral RNA is recognized by cytoplasmic sensors RIG-I or MDA5 in EV-B infections. Signaling through the adaptor protein MAVS leads to IRF3 activation and translocation to the nucleus. These molecules stimulate high level of IFN-α/β and ISG56 production for the development of effective antiviral responses to EV-B infections. (B) According to recent reports, deleted viral genomes were associated with parental complete virus in early phase group-B enterovirus infection. Host cell proteins recruited in viral RNP complexes or potentially acting as a restriction factors could provide an evolutionary advantage to the 5′ NCR truncated viral RNA forms. Deletion in RNA domain-I (5′ terminally deleted viral genomes) could impair the viral genomic RNA recognition by RLRs (RIG-I or MDA5) immune sensors during the early phase of antiviral innate immune response resulting in low level of IFN-α/β and ISG56 production. IFN: interferon; RIG-I: retinoic acid-inducible gene-I; MDA5: melanoma differentiation-associated protein 5; MAVS: mitochondria antiviral-signaling protein; IRF: Interferon Regulatory Factor 3; ISG56: interferon stimulated gene 56; 5′ NCR: 5′ non-coding region.
Viruses 12 00919 g006
Table 1. Cellular and viral factors involved in the formation of ribonucleoprotein complexes in group-B enterovirus genomic replication or translation process.
Table 1. Cellular and viral factors involved in the formation of ribonucleoprotein complexes in group-B enterovirus genomic replication or translation process.
Cellular/Viral FactorsBinding Site in RNA StructureInteraction PartnerRole in EV-B Replication ActivitiesReferences
PCBP 1Poly(C) Binding Protein 1SL-I “b” (Cloverleaf) and reduced affinity to SL-IV (IRES)Viral 3C proteases (3Cpro) and 3CDproInitiation of the viral translation and RNA synthesis[50,51,52]
PCBP 2Poly(C) Binding Protein 2SL-I “b”, C-rich spacer) and SL-IV (IRES)SRp20 (Serine/arginine-rich splicing factor), PCBP2 dimerization, viral 2Apro, 3Cpro, and 3CDproInitiation of the viral translation and RNA synthesis. Switch from viral translation to viral genome replication[50,52,53,54,55,56,57,58,59,60]
EF-1αEukaryotic elongation factor 1αSL-IViral 3CDproCofactor candidate of the enteroviral genome replication[61]
LaLupus autoantigenReduced affinity to SL-I, IRES, and 3′ UTRLa dimerization5′ UTR-La/La-3′ UTR complex may form a replication loop and enhance viral replication initiation. Cofactor of the pre-translation initiation protein-RNA complex[62,63]
hnRNP KHeterogeneous nuclear ribonucleoprotein KSL-I to II and SL-IV (IRES) and biotinylated EV71 5′ UTRK Homology (KH) 2 domain, the proline-rich domain, and one neighboring KH domain (KH1 or KH3)Inhibition of viral RNA synthesis[64,65]
hnRNP A1 & A2Heterogeneous nuclear ribonucleoprotein A1 and A2SL-II to II and SL-VI (IRES)-Activation of IRES activity/regulation of alternative splicing, and it antagonizes the activity of serine-arginine rich (SR) family proteins[66,67,68]
hnRNP CHeterogeneous nuclear ribonucleoprotein CPositive-strand SL-V, negative-strand 5′ and SL-I “a” of 3′ endsViral P2 and P3 and 3Dpol, 3CDpro Enhance viral RNA synthesis. Stabilize and promote efficient positive-strand RNA synthesis[69,70,71]
PABPPoly(A) binding proteinSL-I “b” (Cloverleaf) and Poly(A) tailPCBP2, 3Cpro, and 3CDproEnhance IRES-mediated translation and RNA synthesis. Genome circularization for the initiation of negative strand RNA synthesis[39,44,72]
3ABProtein 3AB, precursor of viral proteins 3A and VPg (viral protein genome-linked)SL-I “b” (Cloverleaf) complexed with 3CDproViral protease 3C and 3CD in absence of RNAEnhance IRES-mediated translation and RNA synthesis. Destabilize the secondary structures of RNA, enhance its hybridization in viral replication and 3Dpol stimulation[73]
3CDproProteinase 3CD, precursor of viral 3Cpro and 3DpolSL-I “d” (Cloverleaf) and 3′ UTR and Cis acting replication element (cre)3AB, PABP1/2, EF-1α, hnRNP C, PABPShut off of the cellular transcription, protease activity, circularization of the viral genome and complex formation for RNA replication[36,74,75,76,77,78,79,80,81,82]
3CproProteinase 3CSL-I “b” or “d” (Cloverleaf)PABP, PCBP2Cleavage of cellular proteins like translation initiation factor eIF4G and PABP leading to shut off of host translation + cleavage of hnRNP A1 leading to reduce inhibitor effect of hnRNP A1 on apoptosis + initiations of the viral translation and RNA synthesis[36,56,80,82,83,84,85,86,87,88]
FBP1Far-upstream element-binding protein 1Only linker region (nt 637 to 745) of the EV71 5′ UTRFUSE upstream of the c-myc gene and KH3 and KH4 motifsPositive regulatory factor for IRES activity and enhance viral protein synthesis[65,89]
FBP2Far-upstream element-binding protein 2SL I, II region, the SL II, III region, and SL V, VI and linker regions of EV71 5′ UTRFUSE upstream of the c-myc gene and KH2 and KH4 motifsNegative regulatory factor for IRES activity and inhibition of viral protein synthesis[65,89]
2AproProteinase 2A-Cleave 3CD to produce 3C′ and 3D′ + VP1 and P2Cleavage of cellular proteins like translation initiation factor eIF4G and PABP leading to shutoff of host translation[90,91,92,93]
2C & 2BC2CATPase and precursor 2BCSL I “b” (Cloverleaf)-Only negative-strand synthesis initiation[36,94]
3BVPg (Viral protein genome-linked)5′ ends3DpolPrimer in both positive- and negative-strand RNA synthesis[95]
3DpolViral 3D RNA-dependent RNA polymerase3′ UTR3B, 3AB, and 3DpolElongation activity[61]
EV-B: group-B enterovirus, SL: stem-loop, UTR: untranslated region, IRES: Internal Ribosome Entry Site, FUSE: far-upstream element.

Share and Cite

MDPI and ACS Style

Glenet, M.; Heng, L.; Callon, D.; Lebreil, A.-L.; Gretteau, P.-A.; Nguyen, Y.; Berri, F.; Andreoletti, L. Structures and Functions of Viral 5′ Non-Coding Genomic RNA Domain-I in Group-B Enterovirus Infections. Viruses 2020, 12, 919. https://doi.org/10.3390/v12090919

AMA Style

Glenet M, Heng L, Callon D, Lebreil A-L, Gretteau P-A, Nguyen Y, Berri F, Andreoletti L. Structures and Functions of Viral 5′ Non-Coding Genomic RNA Domain-I in Group-B Enterovirus Infections. Viruses. 2020; 12(9):919. https://doi.org/10.3390/v12090919

Chicago/Turabian Style

Glenet, Marie, Laetitia Heng, Domitille Callon, Anne-Laure Lebreil, Paul-Antoine Gretteau, Yohan Nguyen, Fatma Berri, and Laurent Andreoletti. 2020. "Structures and Functions of Viral 5′ Non-Coding Genomic RNA Domain-I in Group-B Enterovirus Infections" Viruses 12, no. 9: 919. https://doi.org/10.3390/v12090919

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