Next Article in Journal
Mechanisms of Cell–Cell Fusion in SARS-CoV-2: An Evolving Strategy for Transmission and Immune Evasion
Previous Article in Journal
Clinical Characteristics of Paediatric RSV, Influenza, and SARS-CoV-2 Infections: Insights from Three Consecutive Seasons
Previous Article in Special Issue
Codon Usage Bias in Human RNA Viruses and Its Impact on Viral Translation, Fitness, and Evolution
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Roles of RNA Structures in the Genome Translation of (+) Sense RNA Viruses

Department of Chemistry and Biochemistry, University of Maryland, Baltimore County, Baltimore, MD 21250, USA
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Viruses 2025, 17(11), 1404; https://doi.org/10.3390/v17111404
Submission received: 4 September 2025 / Revised: 17 October 2025 / Accepted: 18 October 2025 / Published: 22 October 2025

Abstract

Positive (+) sense RNA viruses include many important pathogens that exploit noncanonical translation mechanisms to express their genomes within the host cells. Unlike DNA or negative (−) sense RNA viruses, (+) sense RNA viruses can directly function as mRNAs, even though they lack typical features of host mRNAs, such as the 5′ cap structure required for canonical translation initiation. Instead, they exploit structured RNA elements to recruit host translational machinery without the 5′ cap, bypassing the canonical translation initiation mechanism. Prominent examples include internal ribosome entry sites (IRESs) and 3′ cap-independent translation enhancers (3′ CITEs). These RNA modules facilitate translation initiation by recruiting the ribosomal subunits, either directly or through initiation factors, and mediating long-range RNA-RNA interactions. Other regulatory motifs, such as frameshifting signals, allow the ribosome to shift reading frames to regulate protein output. All these RNA elements function through RNA-protein interactions and often utilize host and virus-encoded proteins to hijack the host’s translational apparatus. Over the past several years, various structural biology approaches, including biochemical and enzymatic probing, X-ray crystallography, nuclear magnetic resonance (NMR) spectroscopy, and cryogenic electron microscopy (cryo-EM), have revealed the unique structural roles of these viral RNA elements and their protein complexes. Although a few structures of IRES and CITE domains have been solved through these methods, the structures of these RNA elements and their structure-function relationship have remained largely unknown. This review discusses the current understanding of translation-related RNA structures in (+) sense RNA viruses, the critical RNA-protein interactions they mediate, and various structural biology approaches used to study them. Since the genome of these viruses serves as a template for two mutually exclusive virological processes, namely genome translation and replication, the review also discusses how viruses can utilize RNA structure-based strategies to regulate the switch between genome translation and replication, highlighting future directions for exploring these fundamental virological processes to develop antiviral therapeutics able to combat diseases caused by these pathogens.

1. Introduction

Viruses are classified into several groups based on their genome type and replication strategy. The Baltimore classification divides viruses into five broad groups: double-stranded DNA (dsDNA), single-stranded DNA (ssDNA), double-stranded RNA (dsRNA), positive (+) sense single-stranded (ss) RNA, negative (−) sense ssRNA, and reverse-transcribing viruses [1,2]. Among these groups, (+) sense RNA viruses include many medically important families, including zoonotic viruses such as those belonging to the Picornaviridae, Flaviviridae, and Coronaviridae families and plant viruses such as those belonging to the Potyviridae and Tombusviridae families [3,4,5,6,7]. In contrast to DNA viruses or (−) sense RNA viruses, (+) sense RNA viral genomes can directly serve as an mRNA upon entry to the host cell, allowing immediate translation of viral proteins by the host translation machinery [8,9,10].
Among several families of (+) sense RNA viruses, the Picornaviridae family with 68 genera and 159 species includes the most significant number of human and animal pathogens (Figure 1a), such as Enteroviruses, Hepatoviruses, Cardioviruses and Aphthoviruses [10,11,12]. Unlike host mRNAs (Figure 1b), the genomes of (+) sense RNA viruses typically do not contain a 5′ cap (7-methylguanosine, m7G) but often carry a viral protein VPg (viral protein genome-linked) covalently attached to the 5′ end (Figure 1c) [13,14,15]. These viruses must therefore evolve mechanisms for ribosome recruitment that bypass the canonical mRNA translation initiation, which relies on the 5′ cap [16,17]. Thus, they frequently harbor structured RNA elements that mimic or replace the host mRNA translation process, especially the most critical initiation steps. Notably, (+) sense RNA picornaviruses tightly pack multiple functions into their limited-sized genome (typically less than 10 kb), using RNA structures that regulate multiple virological processes such as translation, new RNA synthesis, and genome packaging [10,18]. For example, cis-acting replication elements (cre) in the RNA genome of enteroviruses can mediate and regulate both replication and translation by bringing the genome’s 5′ and 3′ ends together and recruiting several viral and host protein factors for each mechanism [19,20,21].
The lifecycle of a typical (+) sense RNA virus (Figure 2) begins with the virus particle attaching to the host cell surface and entering the cell through membrane fusion or endocytosis [22]. Once in the host cytoplasm, the virus releases its positive-sense genome, which the host’s ribosomes directly translate to produce genome-encoded proteins. Viral translation usually produces a polyprotein (in picornaviruses and flaviviruses) or sets of structural and nonstructural proteins (in coronaviruses) [23]. The nonstructural proteins include an RNA-dependent RNA polymerase (RdRp) that catalyzes the genomic RNA replication [24]. This often involves synthesizing a complementary (−) sense strand, which then serves as a template for synthesizing the new (+) sense genomic RNA. Because the same RNA sense is a template for two mutually exclusive processes, replication and translation, these two processes must be coordinated, and (+) sense RNA viruses must temporally regulate the use of their genomes for translation and replication. Typically, the incoming genome is first translated to produce nonstructural proteins, including replication factors, before switching to serve as a template for RNA synthesis [25,26]. Finally, the newly synthesized genomic RNAs are packaged with capsid proteins to form new virions, which exit the cell by lysis or vesicle formation [22]. While both translation and replication of viral genomes depend on various host factors and involve multiple steps, this review focuses on the fundamentals of (+) sense RNA genome translation mechanisms in eukaryotic host cells, with particular attention to recent advancements in understanding the structural biology of these unique RNA elements.
In uninfected eukaryotic cells, most mRNAs have an m7G cap at their 5′ ends, which is recognized by the eukaryotic initiation factor 4E (eIF4E), a subunit of the eIF4F complex (composed of eIF4E, eIF4G, and eIF4A) [17,27]. Upon binding with the cap via eIF4E, the eIF4F complex recruits the 43S pre-initiation complex (40S ribosomal subunit bound to eIF1, eIF1A, eIF3, eIF5, and the ternary complex eIF2–Met-tRNA-iGTP) [17]. The 43S complex then scans along the 5′ untranslated region (UTR) from the cap to find an authentic AUG start codon in a favorable context [28,29]. Upon recognizing the start codon by the 43S complex, GTP hydrolysis and the release of initiation factors promote the joining of the 60S subunit to form an 80S ribosome initiation complex ready for elongation [30]. This cap-dependent scanning mechanism (Figure 3a) is highly efficient for typical cellular mRNAs, but can be inhibited or suppressed during stress conditions or viral infections by disrupting the normal function of one or more components of the translation machinery [31].
Many (+) sense RNA viruses bypass the cap-dependent initiation mechanism by using cis-acting RNA elements in their genomes, such as internal ribosome entry sites (IRESs) and 3′ cap-independent translation enhancers (3′ CITEs) (Figure 3b,c). These elements usually contain structured RNA domains that can recruit ribosomes directly or via translation initiation factors. For example, IRES elements within the 5′ UTR of enteroviral genomes recruit the 40S subunit through complex interactions with multiple initiation factors and IRES trans-acting factors (ITAFs), forming initiation complexes without requiring a 5′ cap [16,32,33,34,35,36,37]. Similarly, 3′ CITEs within the 3′ UTR of several plant viruses bind initiation factors (often the eIF4F complex) or ribosomal subunits and interact with the 5′ end through long-range base-pairing, facilitating translation initiation [38,39,40]. Such cap-independent mechanisms allow viral protein synthesis even when host cap-dependent translation is downregulated during infection or stress. Besides IRESs and 3′ CITEs that directly regulate the translation initiation of viral genomes, some RNA structures, such as programmed ribosomal frame-shifting elements, are involved during the translation elongation to control the abundance of specific viral proteins at different stages of the viral life cycle [41,42,43] (Figure 3d). Furthermore, some viruses use small proteins, such as VPg, to covalently attach to their genomic RNAs to mimic the 5′ cap structure, which can interact with eIF4E or other host factors to initiate translation [44]. Therefore, by exploiting these specialized RNA structures in their genomes, (+) sense RNA viruses can effectively hijack the host translational apparatus to ensure efficient viral protein production in host cells.

2. RNA Structures Associated with Viral Genome Translation

2.1. Internal Ribosome Entry Site (IRES)

IRESs are cis-acting RNA elements found in the genomes of many (+) sense RNA viruses and some cellular RNAs that promote cap-independent translation initiation by directly recruiting the ribosome to internal sites within viral genomes or cellular mRNAs [45]. Such a mechanism bypasses the need for a 5′ cap structure essential in the canonical translation mechanism of cellular mRNAs. Therefore, IRESs are crucial for efficient translation when cap-dependent initiation is compromised, like during viral infection or cellular stress [36,46,47,48]. IRES-mediated translation is often regulated by IRES ITAFs, which can either enhance or suppress IRES activity depending on the cellular environment [49,50]. While IRESs are common among (+) sense RNA viruses, including members of the Picornaviridae, Flaviviridae, and Dicistroviridae families, they show considerable sequence and structural diversity despite their similar role in viral genome translation. Including those found in the Picornaviridae family, which is the largest group of (+) sense RNA viruses and includes many viruses that infect humans, IRESs can be categorized into five main types based on their predicted RNA secondary structures, requirements for initiation factors, and differences in how ribosomes are recruited for translation initiation (Table 1).
Type I: These are found in enteroviral genomes and are among the first IRESs identified in Enterovirus coxsackiepol (formerly termed as poliovirus, PV) and enterovirus 71 (EV71), belonging to the Enterovirus alphacoxsackie species [16,51]. The IRES segment typically spans approximately 450 nucleotides and has five structural domains (II to VI, Figure 4a). Translation initiation by type I IRESs requires multiple initiation factors, including eIF1A, eIF2, eIF3, eIF4A, eIF4B, eIF4G and eIF5B [52,53,54] as well as ITAFs such as PCBP2 [53,54,55], PTB [54,56], La [57], and Unr [58]. Each IRES domain tends to bind to these specific factors that promote the recruitment of the 40S ribosome. Initiation involves ribosomal scanning by the 43S preinitiation complex to find the authentic AUG start codon [53]. Unlike the canonical mechanism, they do not require eIF4E and utilize the cleaved form of eIF4G (residues 737-1116) [51,59]. The domain V of type I IRES has been shown to interact with the central domain of eIF4G, which recruits eIF4A and causes downstream conformational changes that may enable the 43S preinitiation complex to bind [59].
Type II: These are typically found in members of the Cardiovirus and Aphthovirus genera, such as Cardiovirus rueckerti (formerly named encephalomyocarditis virus, EMCV) and Aphthovirus vesiculae (formerly named foot-and-mouth disease virus, FMDV) [60,61]. Type II IRES is also found in calicivirus genomes [62]. Similar to Type I IRESs, Type II IRESs are approximately 450 nucleotides long [63] and contain modular elements whose domains are designated H to L (Figure 4b) [64]. These elements directly recruit eIF4G and eIF4A via a conserved J-K domain, promoting ribosome loading without eIF4E or scanning factors such as eIF1 and eIF1A [65]. However, a study has shown that EMCV IRES can directly bind to the 40S subunit without the assistance of eIF4G/4A, but its start codon positioning and assembly efficiency of the 48S pre-initiation complex are higher with the help of factors [66]. Moreover, studies on FMDV and Avian Calicivirus IRESs demonstrate that initiation efficiency and site preference are dynamically regulated by canonical initiation factors and ITAFs [54,67,68,69]. In particular, PTB, Gemin5, Sam68, or ITAF45 can either stimulate or repress initiation at specific sites, often through binding to distinct structural domains within the IRES and inducing conformational remodeling that alters factor recruitment [70,71,72]. Thus, the interplay between initiation factors and ITAFs, together with structural plasticity of the type II IRES, can fine-tune ribosomal recruitment and start codon selection during cap-independent translation.
Type III: Only the Hepatovirus A (formerly named hepatitis A virus, HAV) has been considered in this category, so some studies and previous reviews have highlighted only four types of viral IRESs [29,73,74]. The HAV IRES spans approximately 735 nts in the 5′ UTR and comprises six structured domains (designated I to VI) [75,76]. Although HAV IRES requires all six domains for optimal function, domains III–VI (Figure 4c) have been shown to effectively drive internal initiation, nearly reaching the level of the IRES containing all six domains [76,77]. The HAV IRES requires the complete eIF4F complex, including eIF4E [78], distinguishing it from other picornaviral IRES types. Moreover, unlike type II IRESs, which function with a cleaved form of eIF4G, HAV IRES requires a full-length eIF4G [73,78]. Mechanistically, the proposed initiation proceeds as follows: the IRES structure is recognized by eIF4E binding, which recruits eIF4G; eIF4A-mediated helicase activity helps to resolve local secondary structure near the initiation codon; this allows for recruitment of the 43S pre-initiation complex [79]. Subsequently, the 60S subunit joins to form the 80S ribosome and begins elongation. HAV IRES activity is further modulated by liver cell-specific enhancers or auxiliary factors (including recently found PDAP1), which increase translation efficiency in hepatocyte extracts relative to non-hepatic cells [80].
Table 1. Classification of six types of IRESs based on their predicted secondary structures and their requirements for eukaryotic initiation factors (eIFs) for translation.
Table 1. Classification of six types of IRESs based on their predicted secondary structures and their requirements for eukaryotic initiation factors (eIFs) for translation.
IRES TypesLengthVirusSecondary StructureeIFsITAFs
Type I~450 ntsEnterovirus coxsackiepol, Enterovirus alpharhino, Enterovirus betarhino, Enterovirus cerhino, CalicivirusesFive modular
domains
(II to VI) [81]
eIF1A, eIF2, eIF3, eIF4A, eIF4G, eIF4B, eIF5B [52,53,54]PCBP2 [53,54,55], PTB [54,56], La [57], Unr [58], Sam68 [82]
Type II~450 ntsCardiovirus rueckerti, Aphthovirus vesiculae, CaliciviridaeFive modular domains
(H to L) [64,83,84]
eIF1, eIF1A, eIF2, eIF3, eIF4G, eIF4A [67,68,85,86]PTB [70], PTB + ITAF45/EBP1 [68], Gemin5 [87], Sam68 [88]
Type III~700 ntsHepatovirus AFive modular domains
(II to VI) [76,77]
eIF2, eIF3, eIF4A, eIF4G, eIF4B, eIF4E [73,78,79]PDAP1 [80], PTB [89], PCBP2 [90,91], La [92]
Type IV~300 ntsHepatitis C virus, Classical swine fever virus, Penguin megirivirus, Ruddy turnstone calicivirusCompact structure with pseudoknots
(II to III) [93,94]
eIF1A, eIF2, eIF3, eIF5B [95,96,97,98,99,100,101]La [57,102], hnRNP L [103], hnRNP D [104]
Type V~450 ntsKobuvirus aichi, CalicivirusEight modular domains
(E to L) [105]
eIF2, eIF3, eIF4A, eIF4G [105]DHX29 [105], PTB [105]
IGR~200 ntsCricket paralysis
Virus
Three nested pseudoknots (PKI to PKIII) [106]NANA
Type IV: Although these IRESs have been historically discussed in the context of viruses like the hepatitis C virus (HCV) and the classical swine fever virus (CSFV), which are members of the Flaviviridae [107,108,109], members of the Picornaviridae family have also been shown to contain type IV IRESs. Some recently discovered examples include the penguin megirivirus [110] and Seneca virus Cetus [111], demonstrating that these IRES types are widespread in many viruses affecting even marine animals such as penguins and porpoises. The type IV IRES has an intricate and well-folded RNA structure comprising multiple stem loops and pseudoknots (Figure 4d) [94,112,113]. The HCV IRES can directly bind the 40S ribosomal subunit and position the start codon in the P site without requiring eIF4F, eIF4A, or eIF4B, a process reminiscent of prokaryotic translation initiation [95]. Moreover, studies have shown that IRES domain II facilitates 80S ribosome formation by promoting eIF5-induced GTP hydrolysis and eIF2–GDP release, thereby driving subunit joining [96]. Remarkably, under stress conditions when eIF2 is phosphorylated and inactivated, the HCV IRES can bypass the canonical eIF2-dependent initiation and instead use an alternative pathway involving only eIF3 and eIF5B [98]. However, its function is regulated by multiple specific ITAFs, such as La autoantigen, hnRNP L and hnRNP D [57,102,103,104]. Upon binding with the 40S ribosomal subunit and eIF3, type IV IRES aligns the AUG start codon into the P site through base pairing with 18S rRNA [114]. Subsequently, the recruitment of eIF2-GTP-Met-tRNAᵢ and the joining of the 60S subunit, mediated by eIF5B, complete the translation initiation process [115,116,117].
Type V: The IRESs in this group are believed to be of chimeric evolutionary origin, exhibiting the structural features of both type I and II IRESs. Examples of type V IRESs include those found in Kobuvirus aichi (formerly named Aichivirus, AV) [105]. These IRESs possess eight structural domains, designated E to L (Figure 4e) and require eIF2, eIF3, eIF4A, and the middle domain of the eIF4G (residues 736-1115). Notably, the critical dependency on the DEAH-box RNA helicase DHX29 distinguishes Type V IRESs from other picornavirus IRES classes. The dependency on DHX29 likely indicates the need to unwind stable RNA structures, such as hairpins, which can likely sequester the start codon [105].
IGR: Another distinct class of IRESs is intergenic region (IGR) IRESs (Figure 4f), which represent a unique class of viral IRESs found in the Dicistroviridae family [118], such as Cricket paralysis virus (CrPV) [119,120]. Unlike most viral IRESs in the 5′ UTR, IGR IRESs reside between two open reading frames. They directly assemble elongation-competent 80S ribosomes without canonical initiation factors or Met-tRNA [119,121,122,123]. Structural and biochemical studies have revealed that the IGR IRES adopts a tRNA-mimicking pseudoknot architecture, comprising three nested pseudoknots (PKI to PKIII), with PKI mimicking an anticodon stem–loop base-paired to mRNA and occupying the ribosomal P site [106,122,123,124]. Mechanistically, the IGR IRES first binds to the 40S subunit without eIFs and subsequently recruits the 60S subunit to form an 80S complex. While the PKI domain occupies the ribosomal P site, parts of PKII and PKIII occupy regions normally occupied by E-site and A-site tRNAs, thus positioning the start codon for translation elongation. Although eIF1, eIF1A, and eIF3 can inhibit subunit joining in vitro, the IGR IRES can bypass these inhibitory effects via direct 80S ribosome binding [119,125]. Notably, in Halastavi arva virus (HalV), which lacks domain 2, the IGR IRES employs an even simpler mechanism—directly binding 80S ribosomes with PKI pre-positioned in the P site—allowing the A site to immediately accept the first sense codon without the need for eEF2-mediated translocation [126]. Recent in-cell studies using infectious CrPV clones confirm the physiological relevance of specific IRES structural elements, showing that mutations disrupting pseudoknot formation can impair viral translation and prevent virus production in Drosophila S2 cells. These studies highlight that IGR IRES function is not only a product of its unique structure but also of dynamic interactions with the ribosome in a cellular context [122]. Recent bioinformatics and functional screens revealed that dicistro-like IRES elements are widespread, with about 32% of over 4700 genomes containing putative IRESs, including multiple IRESs or IRESs embedded within ORFs [127]. These elements bind ribosomes directly and support internal translation in vitro and in vivo. Similar factor-less IRES-like structures were also identified in non-dicistrovirus genomes, including Tombusviridae and Narnaviridae families, highlighting the diversity of this mechanism in many viruses [127]. Taken together, these findings reveal the mechanistic strategy by which IGR IRESs recruit ribosomes and initiate translation independently of canonical eukaryotic translation initiation factors, providing insights into their molecular architecture and physiological function.

2.2. 3′ Cap-Independent Translation Enhancer (3′ CITE)

The 3′ cap-independent translation enhancer (3′ CITEs) are structured RNA elements typically found in the 3′ UTRs of many (+) sense RNA viruses. To date, they have only been identified in plant viruses, especially those in the Tombusviridae and Luteoviridae families [128,129,130]. These elements play key roles in promoting efficient translation of viral RNAs that lack typical eukaryotic mRNA features, such as the 5′ cap and poly(A) tail [131,132,133]. The discovery of 3′ CITEs demonstrates an alternative and convergently evolved strategy for cap-independent translation, highlighting the adaptability of viral genomes to exploit host translation machinery, regardless of the viral sources, infected hosts or location of these RNA domains within their genomes (5′ or 3′ UTRs).
Based on their predicted RNA secondary structures and how initiation factors or ribosome recruitment occur, at least seven distinct classes of 3′ CITEs have been identified (Figure 5 and Table 2). Each class presents unique secondary structures and the mechanisms of action [40]. These elements boost viral translation mainly by recruiting host translation initiation factors (eIFs) to the 3′ end of the viral RNA and then helping deliver them to the 5′ end. The mechanistic model of CITE-mediated translation often involves long-distance RNA-RNA or RNA-protein interactions to circularize the viral genome (Figure 3c). Such a circularization mimics the closed-loop structure in capped and polyadenylated eukaryotic mRNAs, enhancing ribosome recruitment and initiation efficiency [134].
Different 3′ CITE classes utilize specific molecular interactions to enhance translation. For example, the barley yellow dwarf virus-like translation element (BTE) adopts a relatively simple branched hairpin structure that directly interacts with the eIF4G subunit of the eIF4F complex. Structural and biochemical studies have confirmed that the BTE-eIF4G interaction is vital for its enhancer activity [135,142,143,144,145]. Conversely, other 3′ CITEs, such as the panicum mosaic virus-like translational enhancer (PTE) [146], the translation enhancer domain (TED) [128,147], the Y-shaped structure (YSS) [131,138,148], and the I-shaped structure (ISS) [40,134,137,149], mainly interact with the eIF4E subunit. CXTE (X representing the Xiajiang isolate) represents a more complex system; it was first identified in cucurbit aphid-borne yellows virus (CABYV) and later in Melon necrotic spot virus (MNSV). It is a modular and transferable architecture with an intricate stem loop structure that binds eIF4G [141,150]. Some classes of 3′ CITEs display additional functional features. For instance, T-shaped structure (TSS) elements have been shown to bind directly to ribosomal subunits, suggesting a more immediate role in translation initiation [139,151]. Additionally, many 3′ CITEs participate in long-range interactions with complementary sequences in the 5′ UTR, forming kissing-loop structures that facilitate the recruitment and transfer of initiation factors from the 3′ to the 5′ end. A well-known example is the PTE of panicum mosaic virus (PMV), which forms a kissing-loop interaction with the 5′ UTR to promote genome circularization and efficient ribosome assembly [40,136].
Table 2. Classification of seven types of 3′ CITEs based on their predicted secondary structures and their interactions with eukaryotic initiation factors (eIFs) for translation. These RNA elements have so far been identified in the plant viruses belonging to the Tombusviridae and Luteoviridae families.
Table 2. Classification of seven types of 3′ CITEs based on their predicted secondary structures and their interactions with eukaryotic initiation factors (eIFs) for translation. These RNA elements have so far been identified in the plant viruses belonging to the Tombusviridae and Luteoviridae families.
3′ CITEs LengthVirusSecondary StructureRoles
BTE~100 ntsBarley Yellow Dwarf Virus [135]Multiple stem loops [142]Interacts with eIF4G [144,152]
TED~120 to150 ntsSatellite Tobacco Necrosis Virus [128]Complex structure, multiple stem loops [128]Interacts with eIF4E [38]
PTE~100 ntsPanicum Mosaic Virus [136], Pea enation mosaic virus [153]Pseudoknot containing structure [153,154]Interacts with eIF4E, promotes RNA circularization [153]
ISS~60 ntsTombusviruses [134], Maize necrotic streak virus [137]Complex stem loop structure with multiple bulges [134,137]Interacts with eIF4E [134,155]
YSS~130 to 150 ntsTomato bushy stunt virus [131], Tombusvirus [138]Y-shaped three-way junction architecture [131]Interacts with eIF4E [153,156]
TSS~100 ntsTurnip crinkle
Virus [139]
T-shaped structure, tRNA-like [139]Interacts directly to ribosomal subunits, promotes RNA circularization [140]
CXTE~55 ntsCucurbit aphid-borne yellows virus [141], Melon Necrotic Spot Virus [150]Complex structure with multiple stem loops [150]Interacts with eIF4G [141]
Notably, 3′ CITEs are often functionally modular and can be transferred between viruses through recombination events [129]. Such mobility emphasizes their evolutionary importance and indicates a role in viral adaptation, such as evading host defenses or expanding host range [157]. Despite several differences in structural and mechanistic strategies to ensure efficient translation, 3′ CITEs act as functional equivalents of the 5′ cap and poly(A) tail in plant viruses. Such sequence diversity and evolutionary flexibility of 3′ CITES also highlight the advanced molecular strategies used by the plant viruses to hijack host translational machinery and maximize viral protein production.

2.3. Other Regulatory RNA Elements in Viral Translation

IRESs and 3′ CITEs are typically involved in translation initiation; however, many (+) sense RNA viruses utilize structured RNA elements, such as programmed ribosomal frameshifting or readthrough signals, during translation elongation to produce multiple proteins from a single RNA [29]. A well-known example is the programmed -1 ribosomal frameshifting element (FSE) in coronaviruses and retroviruses, which is mediated by a heptanucleotide “slippery” sequence and a downstream pseudoknot structure that causes the ribosome to slip back by one nucleotide [158,159]. Such ribosomal frameshifts help balance the production and abundance of specific viral proteins by altering the reading frame during the translation elongation step. For instance, maintaining the ratio of RdRp to structural proteins is essential for balancing the viral life cycle′s translation, replication, and packaging steps [160,161,162]. While there is no consensus structure of the FSEs, they are typically compact pseudoknot or hairpin structures with high mechanical resistance to ribosome unwinding, which causes a pause to trigger the frame change [163]. Additionally, some viruses employ internal termination-reinitiation signals or programmed +1 frameshifts to regulate gene expression [161]. While detailed structures of these frameshift stimulators are emerging, the core principle remains that these RNA structures are very dynamic, can adopt different conformations, and act as checkpoints to switch translational reading frames, thereby controlling the viral protein stoichiometry in the host cells. Classic studies have shown that viral frameshift pseudoknots and stem–loop structures can adopt multiple competing conformations, and this structural plasticity is directly linked to −1 programmed ribosomal frameshifting (PRF) efficiency [42,163,164,165].
Several other viral or host-derived RNA elements also play crucial roles in translation regulation. For example, the termination–reinitiation system in caliciviruses depends on a short RNA motif called the termination upstream ribosome binding site (TURBS), which base-pairs with 18S rRNA and allows the post-termination 40S subunit to stay engaged and reinitiate translation at a downstream AUG [166]. Another example is the 3′ tRNA-like structure (TLS) found in plants and some viral RNAs, which mimics authentic tRNAs, can be aminoacylated by host tRNA synthetases, and thereby contributes to translation efficiency, RNA stability, and replication control [29]. Similarly, the Gamma-interferon-Activated Inhibitor of Translation (GAIT)and GAIT-like elements, initially identified in cellular mRNAs, form structured motifs in the 3′ UTR that recruit host inhibitory complexes to repress translation initiation. Viruses such as SARS-CoV-2 contain GAIT-like motifs, indicating they can hijack this host regulatory mechanism to adjust their gene expression [167].

3. Protein-RNA Interactions in Translation-Associated Complexes

RNA viruses hijack host cellular machinery through specific interactions between viral RNA structures and the viral or host proteins. IRES elements frequently engage a subset of canonical eukaryotic initiation factors (eIFs) and various ITAFs, predominantly a group of host RNA-binding proteins (Table 3). For instance, the poliovirus IRES activity depends on polypyrimidine tract-binding protein (PTB) [47,56], poly(C)-binding protein 2 (PCBP2) [47,168,169,170], and poly(A)-binding protein (PABP), in addition to eIF4G [59] and eIF4A [171]. For picornaviral IRESs, PTB interacts with polypyrimidine tracts within the IRESs [172] and likely induces RNA structural remodeling to facilitate eIF4G positioning [56]. PCBP2 binds to the central domain of type I and III IRESs, enhancing translation activity perhaps by constraining the dynamic regions of the IRES and facilitating the recruitment of other ITAFs and eIFs [168,169,170]. Although the IRES body directly contacts the 40S and positions the AUG into the P site for pairing with the eIF2·GTP·Met-tRNAiMet complex, HCV-type IRESs bind eIF3 through junction IIIabc. Notably, structural data show that the IRES-eIF3 interaction can displace eIF3 from its ribosomal binding site, thus reconfiguring the initiation complex [99,117,173,174,175]. Conversely, Dicistroviridae IRESs are distinctive because they bypass the requirement for eIFs; their RNA pseudoknot directly binds to the ribosomal decoding sites to recruit the ribosome [123,176].
Similarly, 3′ CITE functions depend on their interactions with viral and host proteins, though different CITE types exhibit distinct binding mechanisms with translation initiation factors. For example, the TED element of STNV directly interacts with eIF4F to promote translation initiation, while the YSS of tombusvirus cooperatively engages eIF4E and eIF4G to enhance the ribosome recruitment [38,156]. These interactions mimic cellular mRNAs’ canonical cap-dependent translation strategy, through mechanisms that bypass the 5′ cap [144,152,184]. Additionally, PABP often plays a role in several of these non-canonical translation systems. For example, plant viral systems use CITEs that frequently involve PABP-mediated mechanisms to compensate for the absence of a 5′ cap structure, which otherwise interacts with the 3′ end of mRNAs via poly(A) tail and PABP complex [141,152,185].
Many viral RNAs also interact with virus-encoded proteins, playing vital roles in the viral life cycle. A notable example is the VPg protein in picornaviruses and caliciviruses. VPg is covalently attached to the 5′ end of the viral RNA genome via a phosphodiester bond and modulates translation through interactions with host initiation factors such as eIF4E [14,15,186]. In a separate context, it also acts as a primer for viral RNA replication [187,188]. In caliciviruses, VPg binds to eIF4E, effectively replacing the function of the usual mRNA cap structure [189]. In enteroviruses, VPg is crucial for RNA replication: cre forms a stem-loop in the viral genome that guides the uridylation of VPg, initiating (−) sense RNA synthesis, which is the first step in (+) sense viral genome replication [21,190,191]. RNA-protein interactions can also serve as molecular switches that coordinate viral genome translation and replication [66,192]. Many viruses contain specialized RNA sequences, called cis-acting replication elements, that can recruit both host cellular proteins and viral RNA polymerase [19]. These elements create a regulatory network: when bound by replication machinery, they can suppress translation through the nearby IRES domains, allowing the virus to shift from protein synthesis to genome replication [193]. This temporal regulation ensures efficient viral production by preventing simultaneous translation and replication on the same RNA template [194]. Table 3 above summarizes key host factors (ITAFs) facilitating this coordinated process.

4. Structural Studies of Translation-Related RNAs

The diversities among the IRES and 3′ CITE-mediated mechanisms reflect various evolutionary strategies viruses have developed to simplify cap-independent translation of their (+) sense RNA genomes. These elements allow viral genomes to bypass host translational controls and maintain protein synthesis even during a global shutdown of cap-dependent translation, supporting viral propagation, host adaptation, and pathogenicity. To perform their functions, IRES and 3′ CITE elements exhibit complex folds that enable the precise spatial arrangement of functional RNA motifs, which are essential for interactions with ribosomes and initiation factors. For example, the HAV IRES contains a three-way junction stabilized by an adenine-rich loop [195], similar to the J-K domains in EMCV IRES [65], indicating a conserved mechanism for recruiting eIF4G/eIF4A [195]. Recent cryo-EM analyses of the CrPV IRES have shown that it forms a compact fold with multiple pseudoknot structures that directly interact with the 40S ribosomal head and the P site. Instead of using host initiation factors or an upstream AUG to form a 48S preinitiation complex, the CrPV IRES positions a pseudoknot into the P site and employs built-in mimicry to initiate translation, with eEF2-mediated translocation steps required to transition into an elongation-competent 80S ribosome [196,197]. Similarly, the crystal structures of Saguaro cactus virus (SCV) and Pea enation mosaic virus 2 (PEMV2) 3′ CITE domains illustrated how an uncapped viral RNA domain mimics the 5′ cap to recruit the host cap-binding protein, eIF4E [198,199]. While gaining insights into the three-dimensional structure of viral RNA elements and their complexes is crucial to advancing our understanding of viral translation processes [41,151,164,170,196,200], only a handful of high-resolution structures have been determined so far, and the structural biology of the majority of these RNA elements remains elusive. Over the past several years, various experimental techniques have been used to study the structures of IRESs, 3′ CITEs, ribosomal frame-shifting, and other translation-related RNAs and their protein complexes, which are summarized in Table 4 along with their advantages and limitations.

4.1. Biochemical Probing Methods

Historically, the characterization of RNA secondary structures has depended on enzymatic and chemical probing techniques, such as RNase digestion, DMS-probing, and SHAPE chemistry. Combined with covariation studies in natural isolates, these methods effectively identify base-pairing interactions and flexible regions within RNA molecules, identifying potential stems and loops in IRESs and CITEs and predicting long-range interactions [31,114]. An example of successful DMS-probing can be seen with earlier work on HCV IRES, where 18s rRNA was determined to bind in the unprotected, apical loop of HCV IRES IIId [114]. This was achieved by modifying the triple CCC region in 18s rRNA and using reverse transcription to show which positions are protected when bound to HCV IRES. DMS probing has also been used to predict secondary structure of the SARS-CoV-2 FSE [200], which is then further used for determining how antisense oligonucleotides change the conformation of FSE, and its effect on -1 frame shifts [200]. After the discovery of the SHAPE (selective 2′-hydroxyl acylation analyzed by primer extension) [220] method, secondary structures of many RNAs, including IRESs, 3′ CITEs, FSEs, and even whole genomes of select viruses, have been probed by this strategy [41,212,221,222,223]. For example, SHAPE reactivity has been used on various types of IRESs to determine flexible regions and their dynamic behavior [81,174,215,224,225]. For FMDV IRES, the changes in flexibility, measured by SHAPE reactivity, were combined with computer modeling software to show long-range interactions between domains IV and V of the 3′ UTR and the multiple stem-loop structures within the 5′ IRES [224]. However, probing data alone cannot reveal tertiary contacts or resolve complex folding patterns. Moreover, some RNA conformations are inherently dynamic or may only take certain structures in the presence of specific proteins; such transient or alternative conformations are challenging to characterize with static probing techniques that provide average signals of several conformations that can coexist in solution.

4.2. X-Ray Crystallography

While only a few studies report X-ray crystal structures of translation-related viral RNA domains, these high-resolution crystal structures of viral RNA domains have provided detailed insights into potential mechanisms of viral translation. Notably, the 2.84 Å resolution crystal structure of HAV IRES domain V, complexed with a synthetic antibody fragment, revealed a three-way junction scaffold held by an adenine-rich motif [195]. This structure uncovered a conserved architectural theme later identified in other picornavirus IRESs, such as EMCV J-K domain [65]. X-ray crystallography has also been used to solve structures of smaller motifs, such as the IIId domain of HCV IRES or junctions of dicistrovirus IRESs [174,175,204]. Other translation-related structures derived by X-ray crystallography were PTEs from SCV and PEMV2 RNAs, both in complex with a Fab chaperone [198,199]. These structures were topologically homologous, consistent with their similar binding to eIF4E. Subsequent electrophoretic mobility shift assays showed that a particular flipped-out guanine in both structures was critical for binding eIF4E and how these RNAs effectively mimic the mRNA 5′ cap to hijack eIF4E for viral translation [198,199]. Although a few are determined so far, the atomic models of translation-related RNAs show how particular loops and bulges interact with metal ions or fold into pseudoknots, pre-organizing the RNA for ribosome binding. An example of this is seen in the X-ray crystal structure of the frameshifting element found in SARS-CoV-2, where P1 and P2 Helices connect to loops 1–3 to form an H-type pseudoknot stabilized by cobalt(III)-hexamine, Mg2+, and K+ ions [164]. Nevertheless, studying full-length IRESs and 3′ CITEs with X-ray crystallography is difficult because they are often too large and flexible to form crystals. Even obtaining domain-level crystals requires extensive engineering, such as using antibody chaperones (as seen in the HAV IRES case) or removing flexible regions. Crystal structures show only one static conformation; many viral RNAs can adopt multiple structures depending on the context.

4.3. NMR Spectroscopy

NMR offers insights into the structure and dynamics of RNA elements, especially those under around 100 nucleotides. It has been used to detect local and global changes in IRES elements when they bind to ligands or proteins. For example, Davila-Calderón et al. [206] used NMR to determine the structure of the stem-loop II domain of the EV71 IRES both when free and when bound to a small-molecule ligand, DMA-135 [206]. They found that the binding of DMA-135 allosterically stabilized a conformation that recruits the host protein AUF1, thereby repressing viral translation. NMR has also helped model protein-RNA complexes. For instance, Dorn et al. [208] combined NMR, mass spectrometry, and small-angle X-ray scattering (SAXS) to generate an ensemble structure of the EMCV IRES domains D-F, which are located upstream of the core IRES bound to four RRMs of the ITAF PTBP1, and demonstrated that PTBP1 acts as an RNA chaperone, compacting and organizing the IRES. Structures of EMCV Domains J and K were also solved using NMR and shown to interact with the HEAT-1 domain of eIF4G through multiple interactions between the stem loop domains and bulges orchestrated by an A-rich pentaloop [65]. The secondary structure for the ribosome-binding structural element (RBSE) found in turnip crinkle virus (TCV) 3′ CITE was also determined by NMR; however, SAXS data were needed to generate its 3D molecular envelope and resolve the angles within the hairpin structures of RBSE [151]. The primary limitation of NMR is its size restriction: full IRES RNAs are often too large to analyze, so NMR typically focuses on individual domains or segments. It also requires isotope labeling and high concentrations, which can be challenging. The rapidly exchanging or flexible regions may not be well-resolved.

4.4. Cryo-Electron Microscopy (Cryo-EM)

Cryo-EM is an emerging technique for RNA structural determination, and it has been crucial for elucidating the structures of many proteins and protein-RNA complexes. It has significantly advanced the internal ribosome entry site (IRES) structural biology field by enabling detailed visualization of IRES-ribosome assemblies. For instance, Neupane et al. determined the structure of a CrPV IRES bound to the 40S ribosomal subunit at a 3.3 Å map resolution, revealing its intricate architecture that encircles the ribosomal head [211]. Recent cryo-EM investigations have captured interactions such as HCV IRES binding to eIF3-40S complexes and EMCV IRES engaging with initiation factors [101]. The 2023 study by Imai et al. employed cryo-EM to elucidate a dynamically regulated interaction between viral RNA and two distinct sites on eIF4G, providing insights into the mechanism by which EMCV IRESs recruit host factors [205]. The cryo-EM structure has also been solved for the complex between PCBP2 and domain IV of type I IRES found in poliovirus RNA [170]. Although determined at a resolution of 6.1 Å, the cryo-EM structure indicated that the GNRA tetraloop in domain IV does not directly interact with PCBP2, but the GNRA motif itself is essential for poliovirus translation, possibly through allosteric effects in modulating domain IV structure [170]. While cryo-EM excels in resolving large, asymmetric complexes, it necessitates homogeneous and stable samples. IRES-ribosome complexes often exhibit multiple conformational states, requiring sophisticated image classification techniques. Additionally, due to size limitations, the plant virus 3′ CITEs and smaller IRES elements have yet to be resolved using cryo-EM. One successful attempt at determining smaller structures is the cryo-EM structure of the SARS-CoV-2 FSE [41]. Although this cryo-EM structure is only at 6.9 Å resolution, it reveals the detailed ring formed by the three-stem pseudoknot of the FSE. Notably, cryoEM provides information from reconstituted complexes in vitro, which could differ from the RNA or RNA-protein complexes in infected cells. However, ongoing improvements in resolution and technology, such as cryo-tomography, are expected to enable atomic-level modeling of viral translation initiation complexes on the ribosome inside the cells.

5. Regulation of the Replication-Translation Switching

Because (+) sense RNA genomes serve as templates for two mutually exclusive processes, translation and replication, these viruses must regulate when and how their genomes are used for each process. Early in infection, translation dominates the production of viral proteins. Later, once sufficient viral polymerases build up, the viral RNA is redirected to replication [8,226]. Several mechanisms control this switch. For example, viral proteases cleave host initiation factors such as eIF4G during picornavirus infection. This cleavage suppresses host cap-dependent translation and enhances the relative efficiency of viral IRES-mediated initiation, reprogramming the host translation machinery for viral protein synthesis [227,228]. In poliovirus and enterovirus, the RNA elements play critical roles in regulating the translation-replication switching. For instance, although both domain I, which contributes to structural stability and replication initiation, and domain IV, which is crucial for translation initiation, can bind the same host factor PCBP2, the viral protease 3C or 3CD cleaves PCBP2 bound to domain IV, decreasing the translation and promoting replication [59,170,229,230]. Notably, domain I serve as a binding site for both viral proteins, like 3CD, and host proteins like PCBP2, facilitating the switch to replication by stabilizing RNA structures needed for VPg uridylation [231,232]. Conversely, viral replication complexes can physically sequester viral RNAs at membranes, blocking ribosome access. Some cis-acting elements have dual roles: a stem-loop may initially recruit the ribosome, but it may also help VPg uridylation and replication priming after polyprotein synthesis [13,190,233].
Differential expression of viral proteins can also regulate the balance between replication and translation processes [226]. Internal transcription of sub-genomic RNAs (in coronaviruses, alphaviruses) or frameshifting (in HIV, coronaviruses) modulates the ratio of structural to nonstructural proteins [234,235]. In flaviviruses, including HCV and related genera, translation and replication are tightly interconnected through long-range RNA rearrangements involving structured elements in the 5′ and 3′ UTRs. These rearrangements facilitate genome cyclization via complementary interactions between the 5′ and 3′ ends, supporting the assembly of the replication complex while also regulating translation by influencing ribosome access or initiation factor binding. Disruption of these UTR interactions can shift the balance between translation and replication, emphasizing their dual regulatory roles [175,236]. Additionally, the binding of viral or host proteins to IRES or CITE elements can be regulated by concentration or post-translational modifications [29,72,237]. Some viral ITAFs are produced only after infection, enhancing IRES activity [48]. For example, the stress protein DDX3 binds certain IRESs only under stress conditions, linking environmental cues to viral translation [238,239,240]. While this review focused on RNA components of the genome translation, viruses use a variety of structural RNA switches and factor engagements to coordinate the transition from translation to genome replication. A better understanding of the underpinnings of structural and mechanistic aspects of this crucial topic requires further investigation.

6. Perspectives and Future Directions

Despite significant progress in understanding the roles of RNA structures in translating (+) sense RNA viruses, several unanswered questions and ongoing challenges persist. Obtaining high-resolution structural information for large, dynamic viral RNA-protein complexes remains a significant challenge. While advances in cryo-EM offer new opportunities for such studies, further methodological improvements are necessary to fully clarify the intricate architectures of these assemblies. Regarding IRES-mediated translation, the specific roles of all the various ITAFs identified for different IRES types are not yet fully understood and need further research. In plant virology, dissecting the complex infection cycle within plant hosts and developing more robust experimental systems for biochemistry, genetics, and cellular biology are essential for a more comprehensive understanding of plant virus translation mechanisms. Moreover, many (+) sense RNA viruses contain multiple cis-acting elements within a single RNA molecule, and the functional interactions among these elements and the reasons for their multiplicity warrant further investigation.
The unique RNA structures and their interactions within positive-sense RNA viruses offer promising targets for developing new antiviral treatments. For example, the FSE in coronaviruses, vital for producing the viral RdRp, has attracted interest as a potential drug target. Blocking this frameshifting process has been shown to hinder viral replication, as seen in studies with SARS-CoV-2. This can be seen in previously mentioned cryo-EM structures, which have been applied in developing antisense oligonucleotides made of locked nucleic acids (LNAs) [41]. Likewise, targeting interactions between viral RNA structures, such as IRESs, and key host or viral proteins presents another effective strategy for stopping viral translation without significantly disrupting the host cell’s normal functions. An example of this can be seen in the previously mentioned NMR structures demonstrating conformational changes by DMA-135 that stabilize the AUF1-RNA complex and inhibit translation [206].
Future research in this field will likely focus on several key areas. Continued advancements in structural biology techniques, especially cryo-EM, will be vital for gaining more detailed structural insights into large and dynamic viral RNA-protein complexes. A deeper understanding of the complex host-virus interactions, including the roles of various host RNA-binding proteins (RBPs) and translation factors in modulating viral translation, will be crucial for developing targeted antiviral strategies. Exploring the similarities between viral and host mRNA translation mechanisms, particularly under conditions of cellular stress, could provide new insights into fundamental aspects of translation regulation and uncover novel antiviral targets. RNA-based therapeutics, which include strategies like antisense oligonucleotides and siRNAs targeting viral RNA structures, are also rapidly growing and have significant potential for fighting (+) sense RNA virus infections. Understanding the various RNA processing events in viruses and how they influence translation could also guide future therapeutic approaches. The idea of personalized medicine, where treatment is tailored to an individual’s genetic profile, might become increasingly relevant in RNA virus infections, leading to more effective and targeted therapies. Finally, the remarkable success of mRNA vaccine technology, especially in response to recent viral outbreaks, highlights the critical importance of a solid understanding of RNA virus translation and regulation for developing effective preventive measures.

Author Contributions

Conceptualization, G.L., B.G.B., J.M.C. and D.K.; writing—original draft preparation, G.L. and B.G.B.; writing—review and editing, J.M.C.; visualization, B.G.B.; supervision, D.K.; project administration, D.K.; funding acquisition, D.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Institutes of Health (NIH)—National Institute of General Medical Sciences (NIGMS), MIRA R35GM150869.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

No new data were created in this study. Data sharing is not applicable to this article.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Baltimore, D. Expression of animal virus genomes. Bacteriol. Rev. 1971, 35, 235–241. [Google Scholar] [CrossRef]
  2. Knipe, D.M.; Howley, P. Fields Virology; Wolters Kluwer Health: Philadelphia, PA, USA, 2013. [Google Scholar]
  3. Fehr, A.R.; Perlman, S. Coronaviruses: An Overview of Their Replication and Pathogenesis. In Coronaviruses: Methods and Protocols; Maier, H.J., Bickerton, E., Britton, P., Eds.; Springer: New York, NY, USA, 2015; pp. 1–23. [Google Scholar]
  4. Flather, D.; Semler, B.L. Picornaviruses and nuclear functions: Targeting a cellular compartment distinct from the replication site of a positive-strand RNA virus. Front. Microbiol. 2015, 6, 594. [Google Scholar] [CrossRef]
  5. Pyle, J.D.; Mandadi, K.K.; Scholthof, K.B.G. Panicum Mosaic Virus and Its Satellites Acquire RNA Modifications Associated with Host-Mediated Antiviral Degradation. mBio 2019, 10, e01900-19. [Google Scholar] [CrossRef]
  6. Qin, L.; Shen, W.; Tang, Z.; Hu, W.; Shangguan, L.; Wang, Y.; Tuo, D.; Li, Z.; Miao, W.; Valli, A.A.; et al. A Newly Identified Virus in the Family Potyviridae Encodes Two Leader Cysteine Proteases in Tandem That Evolved Contrasting RNA Silencing Suppression Functions. J. Virol. 2020, 95, e01414-20. [Google Scholar] [CrossRef] [PubMed]
  7. Abram, Q.H.; Landry, B.N.; Wang, A.B.; Kothe, R.F.; Hauch, H.C.H.; Sagan, S.M. The myriad roles of RNA structure in the flavivirus life cycle. RNA Biol. 2024, 21, 14–30. [Google Scholar] [CrossRef]
  8. Ahlquist, P.; Noueiry, A.O.; Lee, W.M.; Kushner, D.B.; Dye, B.T. Host factors in positive-strand RNA virus genome replication. J. Virol. 2003, 77, 8181–8186. [Google Scholar] [CrossRef]
  9. Nagy, P.D.; Pogany, J. The dependence of viral RNA replication on co-opted host factors. Nat. Rev. Microbiol. 2011, 10, 137–149. [Google Scholar] [CrossRef] [PubMed]
  10. Modrow, S.; Falke, D.; Truyen, U.; Schätzl, H. Viruses with Single-Stranded, Positive-Sense RNA Genomes. In Molecular Virology; Springer: Berlin/Heidelberg, Germany, 2013; pp. 185–349. [Google Scholar]
  11. Cifuente, J.O.; Moratorio, G. Evolutionary and Structural Overview of Human Picornavirus Capsid Antibody Evasion. Front. Cell. Infect. Microbiol. 2019, 9, 283. [Google Scholar] [CrossRef]
  12. Andino, R.; Kirkegaard, K.; Macadam, A.; Racaniello, V.R.; Rosenfeld, A.B. The Picornaviridae Family: Knowledge Gaps, Animal Models, Countermeasures, and Prototype Pathogens. J. Infect. Dis. 2023, 228 (Suppl. 6), S427–S445. [Google Scholar] [CrossRef] [PubMed]
  13. Goodfellow, I.G.; Kerrigan, D.; Evans, D.J. Structure and function analysis of the poliovirus cis-acting replication element (CRE). RNA 2003, 9, 124–137. [Google Scholar] [CrossRef]
  14. Goodfellow, I.; Chaudhry, Y.; Gioldasi, I.; Gerondopoulos, A.; Natoni, A.; Labrie, L.; Laliberte, J.F.; Roberts, L. Calicivirus translation initiation requires an interaction between VPg and eIF 4 E. EMBO Rep. 2005, 6, 968–972. [Google Scholar] [CrossRef]
  15. Warsaba, R.; Stoynov, N.; Moon, K.M.; Flibotte, S.; Foster, L.; Jan, E. Multiple Viral Protein Genome-Linked Proteins Compensate for Viral Translation in a Positive-Sense Single-Stranded RNA Virus Infection. J. Virol. 2022, 96, e0069922. [Google Scholar] [CrossRef]
  16. 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]
  17. Sonenberg, N.; Hinnebusch, A.G. Regulation of translation initiation in eukaryotes: Mechanisms and biological targets. Cell 2009, 136, 731–745. [Google Scholar] [CrossRef]
  18. 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]
  19. Herold, J.; Andino, R. Poliovirus RNA replication requires genome circularization through a protein-protein bridge. Mol. Cell 2001, 7, 581–591. [Google Scholar] [CrossRef]
  20. 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]
  21. Cordey, S.; Gerlach, D.; Junier, T.; Zdobnov, E.M.; Kaiser, L.; Tapparel, C. The cis-acting replication elements define human enterovirus and rhinovirus species. RNA 2008, 14, 1568–1578. [Google Scholar] [CrossRef]
  22. den Boon, J.A.; Nishikiori, M.; Zhan, H.; Ahlquist, P. Positive-strand RNA virus genome replication organelles: Structure, assembly, control. Trends Genet. 2024, 40, 681–693. [Google Scholar] [CrossRef]
  23. Bosch, B.J.; van der Zee, R.; de Haan, C.A.; Rottier, P.J. The coronavirus spike protein is a class I virus fusion protein: Structural and functional characterization of the fusion core complex. J. Virol. 2003, 77, 8801–8811. [Google Scholar] [CrossRef]
  24. Machitani, M.; Yasukawa, M.; Nakashima, J.; Furuichi, Y.; Masutomi, K. RNA-dependent RNA polymerase, RdRP, a promising therapeutic target for cancer and potentially COVID-19. Cancer Sci. 2020, 111, 3976–3984. [Google Scholar] [CrossRef]
  25. Lindenbach, B.D.; Rice, C.M. Unravelling hepatitis C virus replication from genome to function. Nature 2005, 436, 933–938. [Google Scholar] [CrossRef]
  26. Boersma, S.; Rabouw, H.H.; Bruurs, L.J.M.; Pavlovic, T.; van Vliet, A.L.W.; Beumer, J.; Clevers, H.; van Kuppeveld, F.J.M.; Tanenbaum, M.E. Translation and Replication Dynamics of Single RNA Viruses. Cell 2020, 183, 1930–1945.e23. [Google Scholar] [CrossRef]
  27. Jackson, R.J.; Hellen, C.U.; Pestova, T.V. The mechanism of eukaryotic translation initiation and principles of its regulation. Nat. Rev. Mol. Cell Biol. 2010, 11, 113–127. [Google Scholar] [CrossRef]
  28. Pestova, T.V.; Kolupaeva, V.G. The roles of individual eukaryotic translation initiation factors in ribosomal scanning and initiation codon selection. Genes Dev. 2002, 16, 2906–2922. [Google Scholar] [CrossRef]
  29. Jaafar, Z.A.; Kieft, J.S. Viral RNA structure-based strategies to manipulate translation. Nat. Rev. Microbiol. 2019, 17, 110–123. [Google Scholar] [CrossRef]
  30. Pestova, T.; Lorsch, J.; Hellen, C. The mechanism of translation initiation in eukaryotes. In Translational Control in Biology and Medicine; Mathews, M.B., Sonenberg, N., Hershey, J.W.B., Eds.; Cold Spring Harbor Laboratory Press: Cold Spring Harbor, NY, USA, 2007; pp. 87–128. [Google Scholar]
  31. Walsh, D.; Mohr, I. Viral subversion of the host protein synthesis machinery. Nat. Rev. Microbiol. 2011, 9, 860–875. [Google Scholar] [CrossRef]
  32. Tsukiyama-Kohara, K.; Iizuka, N.; Kohara, M.; Nomoto, A. Internal ribosome entry site within hepatitis C virus RNA. J. Virol. 1992, 66, 1476–1483. [Google Scholar] [CrossRef]
  33. Jang, S.K.; Krausslich, 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]
  34. Jackson, R.J. Alternative mechanisms of initiating translation of mammalian mRNAs. Biochem. Soc. Trans. 2005, 33 Pt 6, 1231–1241. [Google Scholar] [CrossRef]
  35. Martinez-Salas, E.; Pacheco, A.; Serrano, P.; Fernandez, N. New insights into internal ribosome entry site elements relevant for viral gene expression. J. Gen. Virol. 2008, 89 Pt 3, 611–626. [Google Scholar] [CrossRef]
  36. Niepmann, M. Internal translation initiation of picornaviruses and hepatitis C virus. Biochim. Biophys. Acta 2009, 1789, 529–541. [Google Scholar]
  37. Sorokin, I.I.; Vassilenko, K.S.; Terenin, I.M.; Kalinina, N.O.; Agol, V.I.; Dmitriev, S.E. Non-Canonical Translation Initiation Mechanisms Employed by Eukaryotic Viral mRNAs. Biochemistry 2021, 86, 1060–1094. [Google Scholar] [CrossRef]
  38. Gazo, B.M.; Murphy, P.; Gatchel, J.R.; Browning, K.S. A novel interaction of Cap-binding protein complexes eukaryotic initiation factor (eIF) 4F and eIF(iso)4F with a region in the 3′-untranslated region of satellite tobacco necrosis virus. J. Biol. Chem. 2004, 279, 13584–13592. [Google Scholar] [CrossRef]
  39. Gao, F.; Kasprzak, W.; Stupina, V.A.; Shapiro, B.A.; Simon, A.E. A ribosome-binding, 3′ translational enhancer has a T-shaped structure and engages in a long-distance RNA-RNA interaction. J. Virol. 2012, 86, 9828–9842. [Google Scholar] [CrossRef]
  40. Simon, A.E.; Miller, W.A. 3′ cap-independent translation enhancers of plant viruses. Annu. Rev. Microbiol. 2013, 67, 21–42. [Google Scholar] [CrossRef]
  41. Zhang, K.; Zheludev, I.N.; Hagey, R.J.; Haslecker, R.; Hou, Y.J.; Kretsch, R.; Pintilie, G.D.; Rangan, R.; Kladwang, W.; Li, S.; et al. Cryo-EM and antisense targeting of the 28-kDa frameshift stimulation element from the SARS-CoV-2 RNA genome. Nat. Struct. Mol. Biol. 2021, 28, 747–754. [Google Scholar]
  42. Roman, C.; Lewicka, A.; Koirala, D.; Li, N.S.; Piccirilli, J.A. The SARS-CoV-2 Programmed −1 Ribosomal Frameshifting Element Crystal Structure Solved to 2.09 Å Using Chaperone-Assisted RNA Crystallography. ACS Chem. Biol. 2021, 16, 1469–1481. [Google Scholar] [CrossRef]
  43. Bhatt, P.R.; Scaiola, A.; Loughran, G.; Leibundgut, M.; Kratzel, A.; Meurs, R.; Dreos, R.; O’Connor, K.M.; McMillan, A.; Bode, J.W.; et al. Structural basis of ribosomal frameshifting during translation of the SARS-CoV-2 RNA genome. Science 2021, 372, 1306–1313. [Google Scholar] [CrossRef]
  44. Kneller, E.L.; Rakotondrafara, A.M.; Miller, W.A. Cap-independent translation of plant viral RNAs. Virus Res. 2006, 119, 63–75. [Google Scholar] [CrossRef]
  45. Martinez-Salas, E.; Francisco-Velilla, R.; Fernandez-Chamorro, J.; Embarek, A.M. Insights into Structural and Mechanistic Features of Viral IRES Elements. Front. Microbiol. 2017, 8, 2629. [Google Scholar]
  46. Hanson, P.J.; Zhang, H.M.; Hemida, M.G.; Ye, X.; Qiu, Y.; Yang, D. IRES-Dependent Translational Control during Virus-Induced Endoplasmic Reticulum Stress and Apoptosis. Front. Microbiol. 2012, 3, 92. [Google Scholar]
  47. Godet, A.C.; David, F.; Hantelys, F.; Tatin, F.; Lacazette, E.; Garmy-Susini, B.; Prats, A.C. IRES Trans-Acting Factors, Key Actors of the Stress Response. Int. J. Mol. Sci. 2019, 20, 924. [Google Scholar] [CrossRef]
  48. Marques, R.; Lacerda, R.; Romao, L. Internal Ribosome Entry Site (IRES)-Mediated Translation and Its Potential for Novel mRNA-Based Therapy Development. Biomedicines 2022, 10, 1865. [Google Scholar] [CrossRef]
  49. Chen, L.L.; Kung, Y.A.; Weng, K.F.; Lin, J.Y.; Horng, J.T.; Shih, S.R. Enterovirus 71 infection cleaves a negative regulator for viral internal ribosomal entry site-driven translation. J. Virol. 2013, 87, 3828–3838. [Google Scholar] [CrossRef]
  50. 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]
  51. Thompson, S.R.; Sarnow, P. Enterovirus 71 contains a type I IRES element that functions when eukaryotic initiation factor eIF4G is cleaved. Virology 2003, 315, 259–266. [Google Scholar] [CrossRef]
  52. White, J.P.; Reineke, L.C.; Lloyd, R.E. Poliovirus switches to an eIF2-independent mode of translation during infection. J. Virol. 2011, 85, 8884–8893. [Google Scholar] [CrossRef]
  53. Sweeney, T.R.; Abaeva, I.S.; Pestova, T.V.; Hellen, C.U. The mechanism of translation initiation on Type 1 picornavirus IRESs. EMBO J. 2014, 33, 76–92. [Google Scholar] [CrossRef]
  54. Arhab, Y.; Pestova, T.V.; Hellen, C.U.T. Translation of Overlapping Open Reading Frames Promoted by Type 2 IRESs in Avian Calicivirus Genomes. Viruses 2024, 16, 1413. [Google Scholar] [CrossRef]
  55. Blyn, L.B.; Towner, J.S.; Semler, B.L.; Ehrenfeld, E. Requirement of poly(rC) binding protein 2 for translation of poliovirus RNA. J. Virol. 1997, 71, 6243–6246. [Google Scholar] [CrossRef]
  56. Kafasla, P.; Morgner, N.; Robinson, C.V.; Jackson, R.J. Polypyrimidine tract-binding protein stimulates the poliovirus IRES by modulating eIF4G binding. EMBO J. 2010, 29, 3710–3722. [Google Scholar] [CrossRef]
  57. Costa-Mattioli, M.; Svitkin, Y.; Sonenberg, N. La autoantigen is necessary for optimal function of the poliovirus and hepatitis C virus internal ribosome entry site in vivo and in vitro. Mol. Cell. Biol. 2004, 24, 6861–6870. [Google Scholar] [CrossRef]
  58. 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]
  59. de Breyne, S.; Yu, Y.; Unbehaun, A.; Pestova, T.V.; Hellen, C.U. Direct functional interaction of initiation factor eIF4G with type 1 internal ribosomal entry sites. Proc. Natl. Acad. Sci. USA 2009, 106, 9197–9202. [Google Scholar] [CrossRef] [PubMed]
  60. Luz, N.; Beck, E. Interaction of a cellular 57-kilodalton protein with the internal translation initiation site of foot-and-mouth disease virus. J. Virol. 1991, 65, 6486–6494. [Google Scholar] [CrossRef] [PubMed]
  61. Gao, Y.; Sun, S.Q.; Guo, H.C. Biological function of Foot-and-mouth disease virus non-structural proteins and non-coding elements. Virol. J. 2016, 13, 107. [Google Scholar] [CrossRef] [PubMed]
  62. Arhab, Y.; Miścicka, A.; Pestova, T.V.; Hellen, C.U.T. Horizontal gene transfer as a mechanism for the promiscuous acquisition of distinct classes of IRES by avian caliciviruses. Nucleic Acids Res. 2022, 50, 1052–1068. [Google Scholar] [CrossRef]
  63. Asnani, M.; Pestova, T.V.; Hellen, C.U. Initiation on the divergent Type I cadicivirus IRES: Factor requirements and interactions with the translation apparatus. Nucleic Acids Res. 2016, 44, 3390–3407. [Google Scholar] [CrossRef]
  64. Maloney, A.; Joseph, S. Validating the EMCV IRES Secondary Structure with Structure-Function Analysis. Biochemistry 2024, 63, 107–115. [Google Scholar] [CrossRef]
  65. Imai, S.; Kumar, P.; Hellen, C.U.; D’Souza, V.M.; Wagner, G. An accurately preorganized IRES RNA structure enables eIF4G capture for initiation of viral translation. Nat. Struct. Mol. Biol. 2016, 23, 859–864. [Google Scholar] [CrossRef] [PubMed]
  66. Chamond, N.; Deforges, J.; Ulryck, N.; Sargueil, B. 40S recruitment in the absence of eIF4G/4A by EMCV IRES refines the model for translation initiation on the archetype of Type II IRESs. Nucleic Acids Res. 2014, 42, 10373–10384. [Google Scholar] [CrossRef]
  67. Pestova, T.V.; Hellen, C.U.; Shatsky, I.N. Canonical eukaryotic initiation factors determine initiation of translation by internal ribosomal entry. Mol. Cell. Biol. 1996, 16, 6859–6869. [Google Scholar] [CrossRef] [PubMed]
  68. Pilipenko, E.V.; Pestova, T.V.; Kolupaeva, V.G.; Khitrina, E.V.; Poperechnaya, A.N.; Agol, V.I.; Hellen, C.U. A cell cycle-dependent protein serves as a template-specific translation initiation factor. Genes Dev. 2000, 14, 2028–2045. [Google Scholar] [CrossRef] [PubMed]
  69. Andreev, D.E.; Fernandez-Miragall, O.; Ramajo, J.; Dmitriev, S.E.; Terenin, I.M.; Martinez-Salas, E.; Shatsky, I.N. Differential factor requirement to assemble translation initiation complexes at the alternative start codons of foot-and-mouth disease virus RNA. RNA 2007, 13, 1366–1374. [Google Scholar] [CrossRef]
  70. Kaminski, A.; Jackson, R.J. The polypyrimidine tract binding protein (PTB) requirement for internal initiation of translation of cardiovirus RNAs is conditional rather than absolute. RNA 1998, 4, 626–638. [Google Scholar] [CrossRef]
  71. Yu, Y.; Abaeva, I.S.; Marintchev, A.; Pestova, T.V.; Hellen, C.U. Common conformational changes induced in type 2 picornavirus IRESs by cognate trans-acting factors. Nucleic Acids Res. 2011, 39, 4851–4865. [Google Scholar] [CrossRef]
  72. López-Ulloa, B.; Fuentes, Y.; Pizarro-Ortega, M.S.; López-Lastra, M. RNA-Binding Proteins as Regulators of Internal Initiation of Viral mRNA Translation. Viruses 2022, 14, 188. [Google Scholar] [CrossRef]
  73. Borman, A.M.; Kean, K.M. Intact eukaryotic initiation factor 4G is required for hepatitis A virus internal initiation of translation. Virology 1997, 237, 129–136. [Google Scholar] [CrossRef]
  74. Yang, Y.; Wang, Z. IRES-mediated cap-independent translation, a path leading to hidden proteome. J. Mol. Cell Biol. 2019, 11, 911–919. [Google Scholar] [CrossRef]
  75. Cohen, J.I.; Ticehurst, J.R.; Purcell, R.H.; Buckler-White, A.; Baroudy, B.M. Complete nucleotide sequence of wild-type hepatitis A virus: Comparison with different strains of hepatitis A virus and other picornaviruses. J. Virol. 1987, 61, 50–59. [Google Scholar] [CrossRef]
  76. Brown, E.A.; Day, S.P.; Jansen, R.W.; Lemon, S.M. The 5′ nontranslated region of hepatitis A virus RNA: Secondary structure and elements required for translation in vitro. J. Virol. 1991, 65, 5828–5838. [Google Scholar] [CrossRef] [PubMed]
  77. Brown, E.A.; Zajac, A.J.; Lemon, S.M. In vitro characterization of an internal ribosomal entry site (IRES) present within the 5′ nontranslated region of hepatitis A virus RNA: Comparison with the IRES of encephalomyocarditis virus. J. Virol. 1994, 68, 1066–1074. [Google Scholar] [CrossRef]
  78. Ali, I.K.; McKendrick, L.; Morley, S.J.; Jackson, R.J. Activity of the hepatitis A virus IRES requires association between the cap-binding translation initiation factor (eIF4E) and eIF4G. J. Virol. 2001, 75, 7854–7863. [Google Scholar] [PubMed]
  79. Borman, A.M.; Michel, Y.M.; Kean, K.M. Detailed analysis of the requirements of hepatitis A virus internal ribosome entry segment for the eukaryotic initiation factor complex eIF4F. J. Virol. 2001, 75, 7864–7871. [Google Scholar] [CrossRef]
  80. Shirasaki, T.; Lenarcic, E.; Misumi, I.; Xie, L.; Fusco, W.G.; Yonish, B.; Das, A.; Kim, H.; Cameron, C.E.; Leger-Abraham, M.; et al. Hepatovirus translation requires PDGFA-associated protein 1, an eIF4E-binding protein regulating endoplasmic reticulum stress responses. Sci. Adv. 2024, 10, eadq6342. [Google Scholar] [CrossRef]
  81. Burrill, C.P.; Westesson, O.; Schulte, M.B.; Strings, V.R.; Segal, M.; Andino, R. Global RNA structure analysis of poliovirus identifies a conserved RNA structure involved in viral replication and infectivity. J. Virol. 2013, 87, 11670–11683. [Google Scholar] [CrossRef]
  82. Zhang, H.; Song, L.; Cong, H.; Tien, P. Nuclear Protein Sam68 Interacts with the Enterovirus 71 Internal Ribosome Entry Site and Positively Regulates Viral Protein Translation. J. Virol. 2015, 89, 10031–10043. [Google Scholar] [CrossRef]
  83. Duke, G.M.; Hoffman, M.A.; Palmenberg, A.C. Sequence and structural elements that contribute to efficient encephalomyocarditis virus RNA translation. J. Virol. 1992, 66, 1602–1609. [Google Scholar] [CrossRef]
  84. Bochkov, Y.A.; Palmenberg, A.C. Translational efficiency of EMCV IRES in bicistronic vectors is dependent upon IRES sequence and gene location. Biotechniques 2006, 41, 283–292. [Google Scholar] [CrossRef] [PubMed]
  85. Kolupaeva, V.G.; Pestova, T.V.; Hellen, C.U.; Shatsky, I.N. Translation eukaryotic initiation factor 4G recognizes a specific structural element within the internal ribosome entry site of encephalomyocarditis virus RNA. J. Biol. Chem. 1998, 273, 18599–18604. [Google Scholar] [CrossRef]
  86. Pestova, T.V.; Borukhov, S.I.; Hellen, C.U. Eukaryotic ribosomes require initiation factors 1 and 1A to locate initiation codons. Nature 1998, 394, 854–859. [Google Scholar] [CrossRef]
  87. Pacheco, A.; Lopez de Quinto, S.; Ramajo, J.; Fernandez, N.; Martinez-Salas, E. A novel role for Gemin5 in mRNA translation. Nucleic Acids Res. 2009, 37, 582–590. [Google Scholar] [CrossRef]
  88. 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]
  89. Venkatramana, M.; Ray, P.S.; Chadda, A.; Das, S. A 25 kDa cleavage product of polypyrimidine tract binding protein (PTB) present in mouse tissues prevents PTB binding to the 5′ untranslated region and inhibits translation of hepatitis A virus RNA. Virus Res. 2003, 98, 141–149. [Google Scholar] [CrossRef] [PubMed]
  90. Graff, J.; Cha, J.; Blyn, L.B.; Ehrenfeld, E. Interaction of poly(rC) binding protein 2 with the 5′ noncoding region of hepatitis A virus RNA and its effects on translation. J. Virol. 1998, 72, 9668–9675. [Google Scholar] [CrossRef] [PubMed]
  91. Zhang, B.; Seitz, S.; Kusov, Y.; Zell, R.; Gauss-Muller, V. RNA interaction and cleavage of poly(C)-binding protein 2 by hepatitis A virus protease. Biochem. Biophys. Res. Commun. 2007, 364, 725–730. [Google Scholar] [CrossRef]
  92. Jiang, X.; Kanda, T.; Wu, S.; Nakamoto, S.; Saito, K.; Shirasawa, H.; Kiyohara, T.; Ishii, K.; Wakita, T.; Okamoto, H.; et al. Suppression of La antigen exerts potential antiviral effects against hepatitis A virus. PLoS ONE 2014, 9, e101993. [Google Scholar] [CrossRef]
  93. Rijnbrand, R.; van der Straaten, T.; van Rijn, P.A.; Spaan, W.J.; Bredenbeek, P.J. Internal entry of ribosomes is directed by the 5′ noncoding region of classical swine fever virus and is dependent on the presence of an RNA pseudoknot upstream of the initiation codon. J. Virol. 1997, 71, 451–457. [Google Scholar] [CrossRef]
  94. Lukavsky, P.J.; Otto, G.A.; Lancaster, A.M.; Sarnow, P.; Puglisi, J.D. Structures of two RNA domains essential for hepatitis C virus internal ribosome entry site function. Nat. Struct. Biol. 2000, 7, 1105–1110. [Google Scholar] [PubMed]
  95. Pestova, T.V.; Shatsky, I.N.; Fletcher, S.P.; Jackson, R.J.; Hellen, C.U. A prokaryotic-like mode of cytoplasmic eukaryotic ribosome binding to the initiation codon during internal translation initiation of hepatitis C and classical swine fever virus RNAs. Genes Dev. 1998, 12, 67–83. [Google Scholar]
  96. Locker, N.; Easton, L.E.; Lukavsky, P.J. HCV and CSFV IRES domain II mediate eIF2 release during 80S ribosome assembly. EMBO J. 2007, 26, 795–805. [Google Scholar] [CrossRef] [PubMed]
  97. Pestova, T.V.; de Breyne, S.; Pisarev, A.V.; Abaeva, I.S.; Hellen, C.U. eIF2-dependent and eIF2-independent modes of initiation on the CSFV IRES: A common role of domain II. EMBO J. 2008, 27, 1060–1072. [Google Scholar] [CrossRef] [PubMed]
  98. Terenin, I.M.; Dmitriev, S.E.; Andreev, D.E.; Shatsky, I.N. Eukaryotic translation initiation machinery can operate in a bacterial-like mode without eIF2. Nat. Struct. Mol. Biol. 2008, 15, 836–841. [Google Scholar] [CrossRef] [PubMed]
  99. Hashem, Y.; des Georges, A.; Dhote, V.; Langlois, R.; Liao, H.Y.; Grassucci, R.A.; Pestova, T.V.; Hellen, C.U.; Frank, J. Hepatitis-C-virus-like internal ribosome entry sites displace eIF3 to gain access to the 40S subunit. Nature 2013, 503, 539–543. [Google Scholar] [CrossRef]
  100. Jaafar, Z.A.; Oguro, A.; Nakamura, Y.; Kieft, J.S. Translation initiation by the hepatitis C virus IRES requires eIF1A and ribosomal complex remodeling. eLife 2016, 5, e21198. [Google Scholar] [CrossRef]
  101. Brown, Z.P.; Abaeva, I.S.; De, S.; Hellen, C.U.T.; Pestova, T.V.; Frank, J. Molecular architecture of 40S translation initiation complexes on the hepatitis C virus IRES. EMBO J. 2022, 41, e110581. [Google Scholar] [CrossRef]
  102. Ali, N.; Pruijn, G.J.; Kenan, D.J.; Keene, J.D.; Siddiqui, A. Human La antigen is required for the hepatitis C virus internal ribosome entry site-mediated translation. J. Biol. Chem. 2000, 275, 27531–27540. [Google Scholar] [CrossRef]
  103. Hahm, B.; Kim, Y.K.; Kim, J.H.; Kim, T.Y.; Jang, S.K. Heterogeneous nuclear ribonucleoprotein L interacts with the 3′ border of the internal ribosomal entry site of hepatitis C virus. J. Virol. 1998, 72, 8782–8788. [Google Scholar] [CrossRef]
  104. Paek, K.Y.; Kim, C.S.; Park, S.M.; Kim, J.H.; Jang, S.K. RNA-binding protein hnRNP D modulates internal ribosome entry site-dependent translation of hepatitis C virus RNA. J. Virol. 2008, 82, 12082–12093. [Google Scholar] [CrossRef]
  105. Yu, Y.; Sweeney, T.R.; Kafasla, P.; Jackson, R.J.; Pestova, T.V.; Hellen, C.U. The mechanism of translation initiation on Aichivirus RNA mediated by a novel type of picornavirus IRES. EMBO J. 2011, 30, 4423–4436. [Google Scholar] [CrossRef] [PubMed]
  106. Schuler, M.; Connell, S.R.; Lescoute, A.; Giesebrecht, J.; Dabrowski, M.; Schroeer, B.; Mielke, T.; Penczek, P.A.; Westhof, E.; Spahn, C.M. Structure of the ribosome-bound cricket paralysis virus IRES RNA. Nat. Struct. Mol. Biol. 2006, 13, 1092–1096. [Google Scholar] [CrossRef] [PubMed]
  107. Willcocks, M.M.; Locker, N.; Gomwalk, Z.; Royall, E.; Bakhshesh, M.; Belsham, G.J.; Idamakanti, N.; Burroughs, K.D.; Reddy, P.S.; Hallenbeck, P.L.; et al. Structural features of the Seneca Valley virus internal ribosome entry site (IRES) element: A picornavirus with a pestivirus-like IRES. J. Virol. 2011, 85, 4452–4461. [Google Scholar] [CrossRef]
  108. Asnani, M.; Kumar, P.; Hellen, C.U. Widespread distribution and structural diversity of Type IV IRESs in members of Picornaviridae. Virology 2015, 478, 61–74. [Google Scholar] [CrossRef]
  109. Arhab, Y.; Bulakhov, A.G.; Pestova, T.V.; Hellen, C.U.T. Dissemination of Internal Ribosomal Entry Sites (IRES) Between Viruses by Horizontal Gene Transfer. Viruses 2020, 12, 612. [Google Scholar] [CrossRef]
  110. Yinda, C.K.; Esefeld, J.; Peter, H.U.; Matthijnssens, J.; Zell, R. Penguin megrivirus, a novel picornavirus from an Adelie penguin (Pygoscelis adeliae). Arch. Virol. 2019, 164, 2887–2890. [Google Scholar] [CrossRef] [PubMed]
  111. Vernygora, O.; Sullivan, D.; Nielsen, O.; Huntington, K.B.; Rouse, N.; Popov, V.L.; Lung, O. Senecavirus cetus a novel picornavirus isolated from cetaceans represents a major host switching to the marine environment. npj Viruses 2024, 2, 33. [Google Scholar] [CrossRef]
  112. Liu, F.; Wang, N.; Wang, Q.; Shan, H. Motif mutations in pseudoknot stem I upstream of start codon in Senecavirus A genome: Impacts on activity of viral IRES and on rescue of recombinant virus. Vet. Microbiol. 2021, 262, 109223. [Google Scholar] [CrossRef]
  113. Easton, L.E.; Locker, N.; Lukavsky, P.J. Conserved functional domains and a novel tertiary interaction near the pseudoknot drive translational activity of hepatitis C virus and hepatitis C virus-like internal ribosome entry sites. Nucleic Acids Res. 2009, 37, 5537–5549. [Google Scholar] [CrossRef]
  114. Malygin, A.A.; Kossinova, O.A.; Shatsky, I.N.; Karpova, G.G. HCV IRES interacts with the 18S rRNA to activate the 40S ribosome for subsequent steps of translation initiation. Nucleic Acids Res. 2013, 41, 8706–8714. [Google Scholar] [CrossRef]
  115. Pisarev, A.V.; Chard, L.S.; Kaku, Y.; Johns, H.L.; Shatsky, I.N.; Belsham, G.J. Functional and structural similarities between the internal ribosome entry sites of hepatitis C virus and porcine teschovirus, a picornavirus. J. Virol. 2004, 78, 4487–4497. [Google Scholar] [CrossRef]
  116. Willcocks, M.M.; Zaini, S.; Chamond, N.; Ulryck, N.; Allouche, D.; Rajagopalan, N.; Davids, N.A.; Fahnoe, U.; Hadsbjerg, J.; Rasmussen, T.B.; et al. Distinct roles for the IIId2 sub-domain in pestivirus and picornavirus internal ribosome entry sites. Nucleic Acids Res. 2017, 45, 13016–13028. [Google Scholar] [CrossRef] [PubMed]
  117. Yokoyama, T.; Machida, K.; Iwasaki, W.; Shigeta, T.; Nishimoto, M.; Takahashi, M.; Sakamoto, A.; Yonemochi, M.; Harada, Y.; Shigematsu, H.; et al. HCV IRES Captures an Actively Translating 80S Ribosome. Mol. Cell 2019, 74, 1205–1214.e8. [Google Scholar] [CrossRef] [PubMed]
  118. Abaeva, I.S.; Pestova, T.V.; Hellen, C.U.T. Genetic mechanisms underlying the structural elaboration and dissemination of viral internal ribosomal entry sites. bioRxiv 2024, 590008. [Google Scholar] [CrossRef] [PubMed]
  119. Pestova, T.V.; Lomakin, I.B.; Hellen, C.U. Position of the CrPV IRES on the 40S subunit and factor dependence of IRES/80S ribosome assembly. EMBO Rep. 2004, 5, 906–913. [Google Scholar] [CrossRef]
  120. Miścicka, A.; Lu, K.; Abaeva, I.S.; Pestova, T.V.; Hellen, C.U.T. Initiation of translation on nedicistrovirus and related intergenic region IRESs by their factor-independent binding to the P site of 80S ribosomes. RNA 2023, 29, 1051–1068. [Google Scholar] [CrossRef]
  121. Wilson, J.E.; Pestova, T.V.; Hellen, C.U.; Sarnow, P. Initiation of protein synthesis from the A site of the ribosome. Cell 2000, 102, 511–520. [Google Scholar] [CrossRef]
  122. Kerr, C.H.; Ma, Z.W.; Jang, C.J.; Thompson, S.R.; Jan, E. Molecular analysis of the factorless internal ribosome entry site in Cricket Paralysis virus infection. Sci. Rep. 2016, 6, 37319. [Google Scholar] [CrossRef]
  123. Abaeva, I.S.; Young, C.; Warsaba, R.; Khan, N.; Tran, L.V.; Jan, E.; Pestova, T.V.; Hellen, C.U.T. The structure and mechanism of action of a distinct class of dicistrovirus intergenic region IRESs. Nucleic Acids Res. 2023, 51, 9294–9313. [Google Scholar] [CrossRef]
  124. Fernandez, I.S.; Bai, X.C.; Murshudov, G.; Scheres, S.H.; Ramakrishnan, V. Initiation of translation by cricket paralysis virus IRES requires its translocation in the ribosome. Cell 2014, 157, 823–831. [Google Scholar] [CrossRef]
  125. Pisareva, V.P.; Pisarev, A.V.; Fernandez, I.S. Dual tRNA mimicry in the Cricket Paralysis Virus IRES uncovers an unexpected similarity with the Hepatitis C Virus IRES. eLife 2018, 7, e34062. [Google Scholar] [CrossRef]
  126. Abaeva, I.S.; Vicens, Q.; Bochler, A.; Soufari, H.; Simonetti, A.; Pestova, T.V.; Hashem, Y.; Hellen, C.U.T. The Halastavi árva Virus Intergenic Region IRES Promotes Translation by the Simplest Possible Initiation Mechanism. Cell Rep. 2020, 33, 108476. [Google Scholar] [CrossRef]
  127. Chapagain, S.; Salcedo-Porras, N.; Abdolahzadeh, A.; Zhang, Y.; Pereira, H.S.; Flibotte, S.; Low, K.; Young, C.; Wu, Y.; Wang, S.; et al. Discovery of functional factorless internal ribosome entry site-like structures through virome mining. PLoS Pathog. 2025, 21, e1013255. [Google Scholar] [CrossRef]
  128. Danthinne, X.; Seurinck, J.; Meulewaeter, F.; Van Montagu, M.; Cornelissen, M. The 3′ untranslated region of satellite tobacco necrosis virus RNA stimulates translation in vitro. Mol. Cell. Biol. 1993, 13, 3340–3349. [Google Scholar]
  129. Truniger, V.; Pechar, G.S.; Aranda, M.A. Advances in Understanding the Mechanism of Cap-Independent Cucurbit Aphid-Borne Yellows Virus Protein Synthesis. Int. J. Mol. Sci. 2023, 24, 17598. [Google Scholar] [CrossRef]
  130. Paudel, D.B.; Sanfacon, H. Mapping of sequences in the 5′ region and 3′ UTR of tomato ringspot virus RNA2 that facilitate cap-independent translation of reporter transcripts in vitro. PLoS ONE 2021, 16, e0249928. [Google Scholar] [CrossRef] [PubMed]
  131. Fabian, M.R.; White, K.A. 5′-3′ RNA-RNA interaction facilitates cap- and poly(A) tail-independent translation of tomato bushy stunt virus mrna: A potential common mechanism for tombusviridae. J. Biol. Chem. 2004, 279, 28862–28872. [Google Scholar] [CrossRef] [PubMed]
  132. Rakotondrafara, A.M.; Polacek, C.; Harris, E.; Miller, W.A. Oscillating kissing stem-loop interactions mediate 5′ scanning-dependent translation by a viral 3′-cap-independent translation element. RNA 2006, 12, 1893–1906. [Google Scholar] [CrossRef]
  133. Nicholson, B.L.; White, K.A. 3′ Cap-independent translation enhancers of positive-strand RNA plant viruses. Curr. Opin. Virol. 2011, 1, 373–380. [Google Scholar] [CrossRef] [PubMed]
  134. Nicholson, B.L.; Wu, B.; Chevtchenko, I.; White, K.A. Tombusvirus recruitment of host translational machinery via the 3′ UTR. RNA 2010, 16, 1402–1419. [Google Scholar] [CrossRef]
  135. Wang, S.; Browning, K.S.; Miller, W.A. A viral sequence in the 3′-untranslated region mimics a 5′ cap in facilitating translation of uncapped mRNA. EMBO J. 1997, 16, 4107–4116. [Google Scholar] [CrossRef]
  136. Batten, J.S.; Desvoyes, B.; Yamamura, Y.; Scholthof, K.B. A translational enhancer element on the 3′-proximal end of the Panicum mosaic virus genome. FEBS Lett. 2006, 580, 2591–2597. [Google Scholar] [CrossRef]
  137. Scheets, K.; Redinbaugh, M.G. Infectious cDNA transcripts of Maize necrotic streak virus: Infectivity and translational characteristics. Virology 2006, 350, 171–183. [Google Scholar] [CrossRef] [PubMed]
  138. Fabian, M.R.; White, K.A. Analysis of a 3′-translation enhancer in a tombusvirus: A dynamic model for RNA-RNA interactions of mRNA termini. RNA 2006, 12, 1304–1314. [Google Scholar] [CrossRef]
  139. McCormack, J.C.; Yuan, X.; Yingling, Y.G.; Kasprzak, W.; Zamora, R.E.; Shapiro, B.A.; Simon, A.E. Structural domains within the 3′ untranslated region of Turnip crinkle virus. J. Virol. 2008, 82, 8706–8720. [Google Scholar] [CrossRef] [PubMed]
  140. Stupina, V.A.; Meskauskas, A.; McCormack, J.C.; Yingling, Y.G.; Shapiro, B.A.; Dinman, J.D.; Simon, A.E. The 3′ proximal translational enhancer of Turnip crinkle virus binds to 60S ribosomal subunits. RNA 2008, 14, 2379–2393. [Google Scholar] [CrossRef]
  141. Truniger, V.; Miras, M.; Aranda, M.A. Structural and Functional Diversity of Plant Virus 3′-Cap-Independent Translation Enhancers (3′-CITEs). Front. Plant Sci. 2017, 8, 2047. [Google Scholar] [CrossRef]
  142. Guo, L.; Allen, E.; Miller, W.A. Structure and function of a cap-independent translation element that functions in either the 3′ or the 5′ untranslated region. RNA 2000, 6, 1808–1820. [Google Scholar] [CrossRef] [PubMed]
  143. Wang, Z.; Kraft, J.J.; Hui, A.Y.; Miller, W.A. Structural plasticity of Barley yellow dwarf virus-like cap-independent translation elements in four genera of plant viral RNAs. Virology 2010, 402, 177–186. [Google Scholar]
  144. Zhao, P.; Liu, Q.; Miller, W.A.; Goss, D.J. Eukaryotic translation initiation factor 4G (eIF4G) coordinates interactions with eIF4A, eIF4B, and eIF4E in binding and translation of the barley yellow dwarf virus 3′ cap-independent translation element (BTE). J. Biol. Chem. 2017, 292, 5921–5931. [Google Scholar] [CrossRef]
  145. Ilyas, M.; Du, Z.; Simon, A.E. Opium Poppy Mosaic Virus Has an Xrn-Resistant, Translated Subgenomic RNA and a BTE 3′ CITE. J. Virol. 2021, 95, e02109-20. [Google Scholar] [CrossRef]
  146. Chattopadhyay, M.; Kuhlmann, M.M.; Kumar, K.; Simon, A.E. Position of the kissing-loop interaction associated with PTE-type 3′CITEs can affect enhancement of cap-independent translation. Virology 2014, 458–459, 43–52. [Google Scholar] [CrossRef]
  147. van Lipzig, R.; Van Montagu, M.; Cornelissen, M.; Meulewaeter, F. Functionality of the STNV translational enhancer domain correlates with affinity for two wheat germ factors. Nucleic Acids Res. 2001, 29, 1080–1086. [Google Scholar] [CrossRef]
  148. Chattopadhyay, M.; Shi, K.; Yuan, X.; Simon, A.E. Long-distance kissing loop interactions between a 3′ proximal Y-shaped structure and apical loops of 5′ hairpins enhance translation of Saguaro cactus virus. Virology 2011, 417, 113–125. [Google Scholar] [PubMed]
  149. Miras, M.; Truniger, V.; Querol-Audi, J.; Aranda, M.A. Analysis of the interacting partners eIF4F and 3′-CITE required for Melon necrotic spot virus cap-independent translation. Mol. Plant Pathol. 2017, 18, 635–648. [Google Scholar] [PubMed]
  150. Miras, M.; Rodríguez-Hernández, A.M.; Romero-López, C.; Berzal-Herranz, A.; Colchero, J.; Aranda, M.A.; Truniger, V. A Dual Interaction Between the 5′- and 3′-Ends of the Melon Necrotic Spot Virus (MNSV) RNA Genome Is Required for Efficient Cap-Independent Translation. Front. Plant Sci. 2018, 9, 625. [Google Scholar] [CrossRef]
  151. Zuo, X.; Wang, J.; Yu, P.; Eyler, D.; Xu, H.; Starich, M.R.; Tiede, D.M.; Simon, A.E.; Kasprzak, W.; Schwieters, C.D.; et al. Solution structure of the cap-independent translational enhancer and ribosome-binding element in the 3′ UTR of turnip crinkle virus. Proc. Natl. Acad. Sci. USA 2010, 107, 1385–1390. [Google Scholar] [CrossRef]
  152. Treder, K.; Kneller, E.L.; Allen, E.M.; Wang, Z.; Browning, K.S.; Miller, W.A. The 3′ cap-independent translation element of Barley yellow dwarf virus binds eIF4F via the eIF4G subunit to initiate translation. RNA 2008, 14, 134–147. [Google Scholar] [PubMed]
  153. Wang, Z.; Treder, K.; Miller, W.A. Structure of a viral cap-independent translation element that functions via high affinity binding to the eIF4E subunit of eIF4F. J. Biol. Chem. 2009, 284, 14189–14202. [Google Scholar] [CrossRef]
  154. Du, Z.; Alekhina, O.M.; Vassilenko, K.S.; Simon, A.E. Concerted action of two 3′ cap-independent translation enhancers increases the competitive strength of translated viral genomes. Nucleic Acids Res. 2017, 45, 9558–9572. [Google Scholar] [CrossRef]
  155. Liu, Q.; Goss, D.J. The 3′ mRNA I-shaped structure of maize necrotic streak virus binds to eukaryotic translation factors for eIF4F-mediated translation initiation. J. Biol. Chem. 2018, 293, 9486–9495. [Google Scholar] [CrossRef]
  156. Nicholson, B.L.; Zaslaver, O.; Mayberry, L.K.; Browning, K.S.; White, K.A. Tombusvirus Y-shaped translational enhancer forms a complex with eIF4F and can be functionally replaced by heterologous translational enhancers. J. Virol. 2013, 87, 1872–1883. [Google Scholar] [CrossRef]
  157. Miras, M.; Sempere, R.N.; Kraft, J.J.; Miller, W.A.; Aranda, M.A.; Truniger, V. Interfamilial recombination between viruses led to acquisition of a novel translation-enhancing RNA element that allows resistance breaking. New Phytol. 2014, 202, 233–246. [Google Scholar]
  158. Harger, J.W.; Meskauskas, A.; Dinman, J.D. An “integrated model” of programmed ribosomal frameshifting. Trends Biochem. Sci. 2002, 27, 448–454. [Google Scholar] [CrossRef]
  159. Plant, E.P. Ribosomal Frameshift Signals in Viral Genomes. In Viral Genomes—Molecular Structure, Diversity, Gene Expression Mechanisms and Host-Virus Interactions; Garcia, M.L., Romanowski, V., Eds.; IntechOpen: Rijeka, Croatia, 2012. [Google Scholar]
  160. Dinman, J.D.; Wickner, R.B. Ribosomal frameshifting efficiency and gag/gag-pol ratio are critical for yeast M1 double-stranded RNA virus propagation. J. Virol. 1992, 66, 3669–3676. [Google Scholar]
  161. Karasev, A.V.; Boyko, V.P.; Gowda, S.; Nikolaeva, O.V.; Hilf, M.E.; Koonin, E.V.; Niblett, C.L.; Cline, K.; Gumpf, D.J.; Lee, R.F.; et al. Complete sequence of the citrus tristeza virus RNA genome. Virology 1995, 208, 511–520. [Google Scholar] [CrossRef] [PubMed]
  162. Plant, E.P.; Rakauskaite, R.; Taylor, D.R.; Dinman, J.D. Achieving a golden mean: Mechanisms by which coronaviruses ensure synthesis of the correct stoichiometric ratios of viral proteins. J. Virol. 2010, 84, 4330–4340. [Google Scholar] [CrossRef] [PubMed]
  163. Giedroc, D.P.; Cornish, P.V. Frameshifting RNA pseudoknots: Structure and mechanism. Virus Res. 2009, 139, 193–208. [Google Scholar] [CrossRef]
  164. Jones, C.P.; Ferre-D’Amare, A.R. Crystal structure of the severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2) frameshifting pseudoknot. RNA 2022, 28, 239–249. [Google Scholar] [PubMed]
  165. Peterson, J.M.; Becker, S.T.; O’Leary, C.A.; Juneja, P.; Yang, Y.; Moss, W.N. Structure of the SARS-CoV-2 Frameshift Stimulatory Element with an Upstream Multibranch Loop. Biochemistry 2024, 63, 1287–1296. [Google Scholar] [CrossRef]
  166. Poyry, T.A.; Kaminski, A.; Connell, E.J.; Fraser, C.S.; Jackson, R.J. The mechanism of an exceptional case of reinitiation after translation of a long ORF reveals why such events do not generally occur in mammalian mRNA translation. Genes Dev. 2007, 21, 3149–3162. [Google Scholar] [CrossRef]
  167. Khan, D.; Fox, P.L. Host-like RNA Elements Regulate Virus Translation. Viruses 2024, 16, 468. [Google Scholar] [CrossRef] [PubMed]
  168. Walter, B.L.; Nguyen, J.H.; Ehrenfeld, E.; Semler, B.L. Differential utilization of poly(rC) binding protein 2 in translation directed by picornavirus IRES elements. RNA 1999, 5, 1570–1585. [Google Scholar] [CrossRef]
  169. 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]
  170. Beckham, S.A.; Matak, M.Y.; Belousoff, M.J.; Venugopal, H.; Shah, N.; Vankadari, N.; Elmlund, H.; Nguyen, J.H.C.; Semler, B.L.; Wilce, M.C.J.; et al. Structure of the PCBP2/stem-loop IV complex underlying translation initiation mediated by the poliovirus type I IRES. Nucleic Acids Res. 2020, 48, 8006–8021. [Google Scholar] [CrossRef] [PubMed]
  171. 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]
  172. Hellen, C.U.; Witherell, G.W.; Schmid, M.; Shin, S.H.; Pestova, T.V.; Gil, A.; Wimmer, E. A cytoplasmic 57-kDa protein that is required for translation of picornavirus RNA by internal ribosomal entry is identical to the nuclear pyrimidine tract-binding protein. Proc. Natl. Acad. Sci. USA 1993, 90, 7642–7646. [Google Scholar] [CrossRef] [PubMed]
  173. Berry, K.E.; Waghray, S.; Doudna, J.A. The HCV IRES pseudoknot positions the initiation codon on the 40S ribosomal subunit. RNA 2010, 16, 1559–1569. [Google Scholar] [CrossRef]
  174. Berry, K.E.; Waghray, S.; Mortimer, S.A.; Bai, Y.; Doudna, J.A. Crystal structure of the HCV IRES central domain reveals strategy for start-codon positioning. Structure 2011, 19, 1456–1466. [Google Scholar] [CrossRef]
  175. Koirala, D.; Lewicka, A.; Koldobskaya, Y.; Huang, H.; Piccirilli, J.A. Synthetic Antibody Binding to a Preorganized RNA Domain of Hepatitis C Virus Internal Ribosome Entry Site Inhibits Translation. ACS Chem. Biol. 2020, 15, 205–216. [Google Scholar] [CrossRef]
  176. Hertz, M.I.; Thompson, S.R. Mechanism of translation initiation by Dicistroviridae IGR IRESs. Virology 2011, 411, 355–361. [Google Scholar] [CrossRef]
  177. Cammas, A.; Pileur, F.; Bonnal, S.; Lewis, S.M.; Leveque, N.; Holcik, M.; Vagner, S. Cytoplasmic relocalization of heterogeneous nuclear ribonucleoprotein A1 controls translation initiation of specific mRNAs. Mol. Biol. Cell 2007, 18, 5048–5059. [Google Scholar] [CrossRef]
  178. Meerovitch, K.; Svitkin, Y.V.; Lee, H.S.; Lejbkowicz, F.; Kenan, D.J.; Chan, E.K.; Agol, V.I.; Keene, J.D.; Sonenberg, N. La autoantigen enhances and corrects aberrant translation of poliovirus RNA in reticulocyte lysate. J. Virol. 1993, 67, 3798–3807. [Google Scholar] [CrossRef] [PubMed]
  179. Ali, N.; Siddiqui, A. The La antigen binds 5′ noncoding region of the hepatitis C virus RNA in the context of the initiator AUG codon and stimulates internal ribosome entry site-mediated translation. Proc. Natl. Acad. Sci. USA 1997, 94, 2249–2254. [Google Scholar] [PubMed]
  180. Yu, Y.; Ji, H.; Doudna, J.A.; Leary, J.A. Mass spectrometric analysis of the human 40S ribosomal subunit: Native and HCV IRES-bound complexes. Protein Sci. 2005, 14, 1438–1446. [Google Scholar] [CrossRef] [PubMed]
  181. Majzoub, K.; Hafirassou, M.L.; Meignin, C.; Goto, A.; Marzi, S.; Fedorova, A.; Verdier, Y.; Vinh, J.; Hoffmann, J.A.; Martin, F.; et al. RACK1 controls IRES-mediated translation of viruses. Cell 2014, 159, 1086–1095. [Google Scholar] [CrossRef]
  182. Pineiro, D.; Fernandez, N.; Ramajo, J.; Martinez-Salas, E. Gemin5 promotes IRES interaction and translation control through its C-terminal region. Nucleic Acids Res. 2013, 41, 1017–1028. [Google Scholar]
  183. 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]
  184. Komar, A.A.; Hatzoglou, M. Cellular IRES-mediated translation: The war of ITAFs in pathophysiological states. Cell Cycle 2011, 10, 229–240. [Google Scholar]
  185. Iwakawa, H.O.; Tajima, Y.; Taniguchi, T.; Kaido, M.; Mise, K.; Tomari, Y.; Taniguchi, H.; Okuno, T. Poly(A)-binding protein facilitates translation of an uncapped/nonpolyadenylated viral RNA by binding to the 3′ untranslated region. J. Virol. 2012, 86, 7836–7849. [Google Scholar]
  186. Michon, T.; Estevez, Y.; Walter, J.; German-Retana, S.; Le Gall, O. The potyviral virus genome-linked protein VPg forms a ternary complex with the eukaryotic initiation factors eIF4E and eIF4G and reduces eIF4E affinity for a mRNA cap analogue. FEBS J. 2006, 273, 1312–1322. [Google Scholar] [CrossRef] [PubMed]
  187. Pathak, H.B.; Oh, H.S.; Goodfellow, I.G.; Arnold, J.J.; Cameron, C.E. Picornavirus genome replication: Roles of precursor proteins and rate-limiting steps in oriI-dependent VPg uridylylation. J. Biol. Chem. 2008, 283, 30677–30688. [Google Scholar] [CrossRef] [PubMed]
  188. Paul, A.V.; Wimmer, E. Initiation of protein-primed picornavirus RNA synthesis. Virus Res. 2015, 206, 12–26. [Google Scholar] [CrossRef]
  189. Chaudhry, Y.; Nayak, A.; Bordeleau, M.E.; Tanaka, J.; Pelletier, J.; Belsham, G.J.; Roberts, L.O.; Goodfellow, I.G. Caliciviruses differ in their functional requirements for eIF4F components. J. Biol. Chem. 2006, 281, 25315–25325. [Google Scholar] [CrossRef]
  190. 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] [PubMed]
  191. Sun, Y.; Wang, Y.; Shan, C.; Chen, C.; Xu, P.; Song, M.; Zhou, H.; Yang, C.; Xu, W.; Shi, P.Y.; et al. Enterovirus 71 VPg uridylation uses a two-molecular mechanism of 3D polymerase. J. Virol. 2012, 86, 13662–13671. [Google Scholar] [CrossRef]
  192. Hanson, W.A.; Romero Agosto, G.A.; Rouskin, S. Viral RNA Interactome: The Ultimate Researcher’s Guide to RNA-Protein Interactions. Viruses 2024, 16, 1702. [Google Scholar]
  193. Perera, R.; Daijogo, S.; Walter, B.L.; Nguyen, J.H.; Semler, B.L. Cellular protein modification by poliovirus: The two faces of poly(rC)-binding protein. J. Virol. 2007, 81, 8919–8932. [Google Scholar]
  194. Steil, B.P.; Barton, D.J. Cis-active RNA elements (CREs) and picornavirus RNA replication. Virus Res. 2009, 139, 240–252. [Google Scholar] [CrossRef]
  195. Koirala, D.; Shao, Y.; Koldobskaya, Y.; Fuller, J.R.; Watkins, A.M.; Shelke, S.A.; Pilipenko, E.V.; Das, R.; Rice, P.A.; Piccirilli, J.A. A conserved RNA structural motif for organizing topology within picornaviral internal ribosome entry sites. Nat. Commun. 2019, 10, 3629. [Google Scholar] [CrossRef]
  196. Muhs, M.; Hilal, T.; Mielke, T.; Skabkin, M.A.; Sanbonmatsu, K.Y.; Pestova, T.V.; Spahn, C.M. Cryo-EM of ribosomal 80S complexes with termination factors reveals the translocated cricket paralysis virus IRES. Mol. Cell 2015, 57, 422–432. [Google Scholar] [CrossRef]
  197. Murray, J.; Savva, C.G.; Shin, B.-S.; Dever, T.E.; Ramakrishnan, V.; Fernández, I.S. Structural characterization of ribosome recruitment and translocation by type IV IRES. eLife 2016, 5, e13567. [Google Scholar] [CrossRef]
  198. Ojha, M.; Vogt, J.; Das, N.K.; Redmond, E.; Singh, K.; Banna, H.A.; Sadat, T.; Koirala, D. Structure of saguaro cactus virus 3′ translational enhancer mimics 5′ cap for eIF4E binding. Proc. Natl. Acad. Sci. USA 2024, 121, e2313677121. [Google Scholar] [CrossRef]
  199. Lewicka, A.; Roman, C.; Jones, S.; Disare, M.; Rice, P.A.; Piccirilli, J.A. Crystal structure of a cap-independent translation enhancer RNA. Nucleic Acids Res. 2023, 51, 8891–8907. [Google Scholar] [CrossRef]
  200. Pekarek, L.; Zimmer, M.M.; Gribling-Burrer, A.S.; Buck, S.; Smyth, R.; Caliskan, N. Cis-mediated interactions of the SARS-CoV-2 frameshift RNA alter its conformations and affect function. Nucleic Acids Res. 2023, 51, 728–743. [Google Scholar] [CrossRef]
  201. Akiyama, B.M.; Graham, M.E.; Z, O.D.; Beckham, J.D.; Kieft, J.S. Three-dimensional structure of a flavivirus dumbbell RNA reveals molecular details of an RNA regulator of replication. Nucleic Acids Res. 2021, 49, 7122–7138. [Google Scholar] [CrossRef] [PubMed]
  202. Dibrov, S.M.; Ding, K.; Brunn, N.D.; Parker, M.A.; Bergdahl, B.M.; Wyles, D.L.; Hermann, T. Structure of a hepatitis C virus RNA domain in complex with a translation inhibitor reveals a binding mode reminiscent of riboswitches. Proc. Natl. Acad. Sci. USA 2012, 109, 5223–5228. [Google Scholar] [CrossRef] [PubMed]
  203. Banna, H.A.; Berg, K.; Sadat, T.; Das, N.K.; Paudel, R.; D’Souza, V.; Koirala, D. Synthetic anti-RNA antibody derivatives for RNA visualization in mammalian cells. Nucleic Acids Res. 2025, 53, gkae1275. [Google Scholar] [CrossRef] [PubMed]
  204. Costantino, D.A.; Pfingsten, J.S.; Rambo, R.P.; Kieft, J.S. tRNA-mRNA mimicry drives translation initiation from a viral IRES. Nat. Struct. Mol. Biol. 2008, 15, 57–64. [Google Scholar] [CrossRef]
  205. Imai, S.; Suzuki, H.; Fujiyoshi, Y.; Shimada, I. Dynamically regulated two-site interaction of viral RNA to capture host translation initiation factor. Nat. Commun. 2023, 14, 4977. [Google Scholar] [CrossRef]
  206. Davila-Calderon, J.; Patwardhan, N.N.; Chiu, L.Y.; Sugarman, A.; Cai, Z.; Penutmutchu, S.R.; Li, M.L.; Brewer, G.; Hargrove, A.E.; Tolbert, B.S. IRES-targeting small molecule inhibits enterovirus 71 replication via allosteric stabilization of a ternary complex. Nat. Commun. 2020, 11, 4775. [Google Scholar] [CrossRef] [PubMed]
  207. Mohammed, S.; Phelan, M.M.; Rasul, U.; Ramesh, V. NMR elucidation of the role of Mg2+ in the structure and stability of the conserved RNA motifs of the EMCV IRES element. Org. Biomol. Chem. 2014, 12, 1495–1509. [Google Scholar] [PubMed]
  208. Dorn, G.; Gmeiner, C.; de Vries, T.; Dedic, E.; Novakovic, M.; Damberger, F.F.; Maris, C.; Finol, E.; Sarnowski, C.P.; Kohlbrecher, J.; et al. Integrative solution structure of PTBP1-IRES complex reveals strong compaction and ordering with residual conformational flexibility. Nat. Commun. 2023, 14, 6429. [Google Scholar] [CrossRef]
  209. Perard, J.; Leyrat, C.; Baudin, F.; Drouet, E.; Jamin, M. Structure of the full-length HCV IRES in solution. Nat. Commun. 2013, 4, 1612. [Google Scholar] [CrossRef] [PubMed]
  210. Lukavsky, P.J.; Kim, I.; Otto, G.A.; Puglisi, J.D. Structure of HCV IRES domain II determined by NMR. Nat. Struct. Biol. 2003, 10, 1033–1038. [Google Scholar] [CrossRef]
  211. Neupane, R.; Pisareva, V.P.; Rodriguez, C.F.; Pisarev, A.V.; Fernandez, I.S. A complex IRES at the 5′-UTR of a viral mRNA assembles a functional 48S complex via an uAUG intermediate. eLife 2020, 9, e54575. [Google Scholar]
  212. Johnson, P.Z.; Kasprzak, W.K.; Shapiro, B.A.; Simon, A.E. Structural characterization of a new subclass of panicum mosaic virus-like 3′ cap-independent translation enhancer. Nucleic Acids Res. 2022, 50, 1601–1619. [Google Scholar]
  213. Bera, S.; Ilyas, M.; Mikkelsen, A.A.; Simon, A.E. Conserved Structure Associated with Different 3′CITEs Is Important for Translation of Umbraviruses. Viruses 2023, 15, 638. [Google Scholar] [CrossRef]
  214. Venkatesan, A.; Dasgupta, A. Novel fluorescence-based screen to identify small synthetic internal ribosome entry site elements. Mol. Cell. Biol. 2001, 21, 2826–2837. [Google Scholar] [CrossRef]
  215. Lozano, G.; Francisco-Velilla, R.; Martinez-Salas, E. Ribosome-dependent conformational flexibility changes and RNA dynamics of IRES domains revealed by differential SHAPE. Sci. Rep. 2018, 8, 5545. [Google Scholar] [CrossRef]
  216. García-Sacristán, A.; Moreno, M.; Ariza-Mateos, A.; López-Camacho, E.; Jáudenes, R.M.; Vázquez, L.; Gómez, J.; Martín-Gago, J.; Briones, C. A magnesium-induced RNA conformational switch at the internal ribosome entry site of hepatitis C virus genome visualized by atomic force microscopy. Nucleic Acids Res. 2015, 43, 565–580. [Google Scholar] [PubMed]
  217. Fernandez-Miragall, O.; Martinez-Salas, E. Structural organization of a viral IRES depends on the integrity of the GNRA motif. RNA 2003, 9, 1333–1344. [Google Scholar] [CrossRef] [PubMed]
  218. Gross, L.; Vicens, Q.; Einhorn, E.; Noireterre, A.; Schaeffer, L.; Kuhn, L.; Imler, J.L.; Eriani, G.; Meignin, C.; Martin, F. The IRES5’UTR of the dicistrovirus cricket paralysis virus is a type III IRES containing an essential pseudoknot structure. Nucleic Acids Res. 2017, 45, 8993–9004. [Google Scholar] [PubMed]
  219. Mahmud, B.; Horn, C.M.; Tapprich, W.E. Structure of the 5′ Untranslated Region of Enteroviral Genomic RNA. J. Virol. 2019, 93, e01288-19. [Google Scholar] [CrossRef]
  220. Merino, E.J.; Wilkinson, K.A.; Coughlan, J.L.; Weeks, K.M. RNA structure analysis at single nucleotide resolution by selective 2′-hydroxyl acylation and primer extension (SHAPE). J. Am. Chem. Soc. 2005, 127, 4223–4231. [Google Scholar]
  221. Low, J.T.; Garcia-Miranda, P.; Mouzakis, K.D.; Gorelick, R.J.; Butcher, S.E.; Weeks, K.M. Structure and dynamics of the HIV-1 frameshift element RNA. Biochemistry 2014, 53, 4282–4291. [Google Scholar] [CrossRef]
  222. Boerneke, M.A.; Ehrhardt, J.E.; Weeks, K.M. Physical and Functional Analysis of Viral RNA Genomes by SHAPE. Annu. Rev. Virol. 2019, 6, 93–117. [Google Scholar] [CrossRef]
  223. Madden, E.A.; Plante, K.S.; Morrison, C.R.; Kutchko, K.M.; Sanders, W.; Long, K.M.; Taft-Benz, S.; Cruz Cisneros, M.C.; White, A.M.; Sarkar, S.; et al. Using SHAPE-MaP To Model RNA Secondary Structure and Identify 3′UTR Variation in Chikungunya Virus. J. Virol. 2020, 94, e00701-20. [Google Scholar] [CrossRef]
  224. Diaz-Toledano, R.; Lozano, G.; Martinez-Salas, E. In-cell SHAPE uncovers dynamic interactions between the untranslated regions of the foot-and-mouth disease virus RNA. Nucleic Acids Res. 2017, 45, 1416–1432. [Google Scholar]
  225. Gosavi, D.; Wower, I.; Beckmann, I.K.; Hofacker, I.L.; Wower, J.; Wolfinger, M.T.; Sztuba-Solinska, J. Insights into the secondary and tertiary structure of the Bovine Viral Diarrhea Virus Internal Ribosome Entry Site. RNA Biol. 2022, 19, 496–506. [Google Scholar] [CrossRef]
  226. 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]
  227. Etchison, D.; Milburn, S.C.; Edery, I.; Sonenberg, N.; Hershey, J.W. Inhibition of HeLa cell protein synthesis following poliovirus infection correlates with the proteolysis of a 220,000-dalton polypeptide associated with eucaryotic initiation factor 3 and a cap binding protein complex. J. Biol. Chem. 1982, 257, 14806–14810. [Google Scholar] [CrossRef]
  228. Gradi, A.; Svitkin, Y.V.; Imataka, H.; Sonenberg, N. Proteolysis of human eukaryotic translation initiation factor eIF4GII, but not eIF4GI, coincides with the shutoff of host protein synthesis after poliovirus infection. Proc. Natl. Acad. Sci. USA 1998, 95, 11089–11094. [Google Scholar]
  229. 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]
  230. Yamamoto, H.; Unbehaun, A.; Spahn, C.M.T. Ribosomal Chamber Music: Toward an Understanding of IRES Mechanisms. Trends Biochem. Sci. 2017, 42, 655–668. [Google Scholar] [CrossRef]
  231. 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]
  232. Das, N.K.; Vogt, J.; Patel, A.; Banna, H.A.; Koirala, D. Structural basis for a highly conserved RNA-mediated enteroviral genome replication. Nucleic Acids Res. 2024, 52, 11218–11233. [Google Scholar] [CrossRef] [PubMed]
  233. Zust, R.; Miller, T.B.; Goebel, S.J.; Thiel, V.; Masters, P.S. Genetic interactions between an essential 3′ cis-acting RNA pseudoknot, replicase gene products, and the extreme 3′ end of the mouse coronavirus genome. J. Virol. 2008, 82, 1214–1228. [Google Scholar] [CrossRef] [PubMed]
  234. Brierley, I.; Dos Ramos, F.J. Programmed ribosomal frameshifting in HIV-1 and the SARS-CoV. Virus Res. 2006, 119, 29–42. [Google Scholar] [CrossRef] [PubMed]
  235. Alexandersen, S.; Chamings, A.; Bhatta, T.R. SARS-CoV-2 genomic and subgenomic RNAs in diagnostic samples are not an indicator of active replication. Nat. Commun. 2020, 11, 6059. [Google Scholar] [CrossRef]
  236. Romero-Lopez, C.; Roda-Herreros, M.; Berzal-Herranz, B.; Ramos-Lorente, S.E.; Berzal-Herranz, A. Inter- and Intramolecular RNA-RNA Interactions Modulate the Regulation of Translation Mediated by the 3′ UTR in West Nile Virus. Int. J. Mol. Sci. 2023, 24, 5337. [Google Scholar] [CrossRef] [PubMed]
  237. Barrera, A.; Ramos, H.; Vera-Otarola, J.; Fernandez-Garcia, L.; Angulo, J.; Olguin, V.; Pino, K.; Mouland, A.J.; Lopez-Lastra, M. Post-translational modifications of hnRNP A1 differentially modulate retroviral IRES-mediated translation initiation. Nucleic Acids Res. 2020, 48, 10479–10499. [Google Scholar] [CrossRef]
  238. Soto-Rifo, R.; Rubilar, P.S.; Limousin, T.; de Breyne, S.; Decimo, D.; Ohlmann, T. DEAD-box protein DDX3 associates with eIF4F to promote translation of selected mRNAs. EMBO J. 2012, 31, 3745–3756. [Google Scholar] [CrossRef]
  239. Su, Y.S.; Tsai, A.H.; Ho, Y.F.; Huang, S.Y.; Liu, Y.C.; Hwang, L.H. Stimulation of the Internal Ribosome Entry Site (IRES)-Dependent Translation of Enterovirus 71 by DDX3X RNA Helicase and Viral 2A and 3C Proteases. Front. Microbiol. 2018, 9, 1324. [Google Scholar] [CrossRef] [PubMed]
  240. Han, S.; Sun, S.; Li, P.; Liu, Q.; Zhang, Z.; Dong, H.; Sun, M.; Wu, W.; Wang, X.; Guo, H. Ribosomal Protein L13 Promotes IRES-Driven Translation of Foot-and-Mouth Disease Virus in a Helicase DDX3-Dependent Manner. J. Virol. 2020, 94, e01679-19. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Types of Picornaviral IRESs and organization of cellular mRNA (cap-dependent translation) and a (+) sense RNA viral genome. (a) Types of IRESs found in members of the Picornaviridae family demonstrate IRES diversity among various (+) sense RNA viruses. However, this IRES classification extends beyond picornaviruses. For instance, type IV IRESs are also present in Hepatitis C Virus, which belongs to the Flaviviridae family. (b) Organization of a typical eukaryotic mRNA, which contains a 5′ m7G cap and 3′ poly(A) tail, both of which protect and activate mature RNA (mRNA) for translation. (c) Organization of a typical (+) sense RNA viral genome, which promotes a cap-independent translation mechanism in host cells through unique RNA elements within its viral genome. Along with the attachment of VPg, viral genomic RNA contains cis-acting elements such as an IRES in the 5′ UTR and 3′ CITEs in the 3′ UTR, enabling interaction with host initiation factors and the ribosome for viral protein translation.
Figure 1. Types of Picornaviral IRESs and organization of cellular mRNA (cap-dependent translation) and a (+) sense RNA viral genome. (a) Types of IRESs found in members of the Picornaviridae family demonstrate IRES diversity among various (+) sense RNA viruses. However, this IRES classification extends beyond picornaviruses. For instance, type IV IRESs are also present in Hepatitis C Virus, which belongs to the Flaviviridae family. (b) Organization of a typical eukaryotic mRNA, which contains a 5′ m7G cap and 3′ poly(A) tail, both of which protect and activate mature RNA (mRNA) for translation. (c) Organization of a typical (+) sense RNA viral genome, which promotes a cap-independent translation mechanism in host cells through unique RNA elements within its viral genome. Along with the attachment of VPg, viral genomic RNA contains cis-acting elements such as an IRES in the 5′ UTR and 3′ CITEs in the 3′ UTR, enabling interaction with host initiation factors and the ribosome for viral protein translation.
Viruses 17 01404 g001
Figure 2. Life cycle of a typical (+) sense RNA virus. The virus is internalized into the host cell via endocytosis, releasing its (+) sense RNA genome into the cytoplasm. The genome directly functions as mRNA, enabling the immediate translation of essential viral proteins by exploiting the host’s ribosomes, amino acids, and other necessary cellular machinery. Translation produces a single polyprotein that, upon proteolysis, produces capsid-related proteins (Vp4, Vp2, Vp3), which assist in viral packaging and encapsidation, enabling further enteroviral infection. Replication-related proteins (2A, 2B, 2C, 3A, 3B, 3C, and 3D) are also generated during proteolytic processing. With the 3D protein acting as a polymerase, a (−) sense RNA is formed using (+) sense genomic RNA as a template. Subsequently, this intermediate (−) RNA serves as a template to synthesize the (+) sense genomic RNA again, which can then undergo translation and replication or be packaged into new virions.
Figure 2. Life cycle of a typical (+) sense RNA virus. The virus is internalized into the host cell via endocytosis, releasing its (+) sense RNA genome into the cytoplasm. The genome directly functions as mRNA, enabling the immediate translation of essential viral proteins by exploiting the host’s ribosomes, amino acids, and other necessary cellular machinery. Translation produces a single polyprotein that, upon proteolysis, produces capsid-related proteins (Vp4, Vp2, Vp3), which assist in viral packaging and encapsidation, enabling further enteroviral infection. Replication-related proteins (2A, 2B, 2C, 3A, 3B, 3C, and 3D) are also generated during proteolytic processing. With the 3D protein acting as a polymerase, a (−) sense RNA is formed using (+) sense genomic RNA as a template. Subsequently, this intermediate (−) RNA serves as a template to synthesize the (+) sense genomic RNA again, which can then undergo translation and replication or be packaged into new virions.
Viruses 17 01404 g002
Figure 3. Mechanisms of translation initiation and regulation in virus-infected eukaryotic cells. (a) Cap-dependent translation in uninfected eukaryotic cells depends on the binding of initiation factors and a poly-A binding protein. The eIF4E in the eIF4F complex (comprising eIF4E, eIF4A, eIF4G, and eIF3) attaches to the 5′ m7G cap to facilitate the attachment of the eIF-bound 40S ribosomal subunit to the mRNA for translation. Cap-independent translation in (+) sense RNA involves using (b) IRES or (c) 3′ CITE to recruit host factors. Utilizing the IRES or 3′ CITE bypasses the need for a cap, as these elements can bind to eIFs and interact with the host ribosome without this feature. (d) Frameshifting elements embedded in the (+) sense RNA can influence ribosomal activity, which regulates the levels of specific viral proteins being translated that may be required at a particular stage of the viral lifecycle.
Figure 3. Mechanisms of translation initiation and regulation in virus-infected eukaryotic cells. (a) Cap-dependent translation in uninfected eukaryotic cells depends on the binding of initiation factors and a poly-A binding protein. The eIF4E in the eIF4F complex (comprising eIF4E, eIF4A, eIF4G, and eIF3) attaches to the 5′ m7G cap to facilitate the attachment of the eIF-bound 40S ribosomal subunit to the mRNA for translation. Cap-independent translation in (+) sense RNA involves using (b) IRES or (c) 3′ CITE to recruit host factors. Utilizing the IRES or 3′ CITE bypasses the need for a cap, as these elements can bind to eIFs and interact with the host ribosome without this feature. (d) Frameshifting elements embedded in the (+) sense RNA can influence ribosomal activity, which regulates the levels of specific viral proteins being translated that may be required at a particular stage of the viral lifecycle.
Viruses 17 01404 g003
Figure 4. Various types of IRESs and their predicted secondary structures. The subdomains and major secondary structural features are labeled. Types I–III IRESs (a–c) are prevalent among members of the Picornaviridae family. The type IV IRESs (d) were initially found in members of Flaviviridae, such as HCV and CSFV, but they were later identified in the porcine teschovirus, a member of the Picornaviridae. Type V IRES (e) was recognized in Aichi viruses. The Intergenic region (IGR) IRES (f) is found in viruses such as cricket paralysis virus (CrPV), with some structural and ribosome recruitment features similar to type IV IRESs.
Figure 4. Various types of IRESs and their predicted secondary structures. The subdomains and major secondary structural features are labeled. Types I–III IRESs (a–c) are prevalent among members of the Picornaviridae family. The type IV IRESs (d) were initially found in members of Flaviviridae, such as HCV and CSFV, but they were later identified in the porcine teschovirus, a member of the Picornaviridae. Type V IRES (e) was recognized in Aichi viruses. The Intergenic region (IGR) IRES (f) is found in viruses such as cricket paralysis virus (CrPV), with some structural and ribosome recruitment features similar to type IV IRESs.
Viruses 17 01404 g004
Figure 5. Various types of 3′ CITEs and their predicted secondary structures. BTE was identified in Barley yellow dwarf virus [135]. TED was among the first 3′ CITEs discovered in the satellite Tobacco necrosis virus [128]. PTE and ISS are found in panicum mosaic virus and maize necrotic streak virus and bind eIF4E through a flipped-out guanine within their structures [136,137]. YSS has been reported in Tomato bushy stunt virus and tombusviruses [131,138]. They typically contain three stem-loops essential for function. TSS are characteristic of Turnip crinkle virus and fold into intricate structures containing pseudoknots [139,140]. The CXTEs have complex stem-loop structures and are named based on the geographical isolate of the Cucurbit aphid-borne yellows virus, with CXTE originating from Xiajiang [141].
Figure 5. Various types of 3′ CITEs and their predicted secondary structures. BTE was identified in Barley yellow dwarf virus [135]. TED was among the first 3′ CITEs discovered in the satellite Tobacco necrosis virus [128]. PTE and ISS are found in panicum mosaic virus and maize necrotic streak virus and bind eIF4E through a flipped-out guanine within their structures [136,137]. YSS has been reported in Tomato bushy stunt virus and tombusviruses [131,138]. They typically contain three stem-loops essential for function. TSS are characteristic of Turnip crinkle virus and fold into intricate structures containing pseudoknots [139,140]. The CXTEs have complex stem-loop structures and are named based on the geographical isolate of the Cucurbit aphid-borne yellows virus, with CXTE originating from Xiajiang [141].
Viruses 17 01404 g005
Table 3. Some crucial ITAFs and their regulatory roles in genome translation of (+) sense RNA viruses.
Table 3. Some crucial ITAFs and their regulatory roles in genome translation of (+) sense RNA viruses.
ITAFIRESsITAF’s RolesIRES ActivitiesReferences
PTB (hnRNP I)PV, HRVRNA chaperone and stabilizer; alters IRES secondary structureActivation, inhibition (context-dependent)References [56,70,172]
PCBP2 (hnRNP E)PV, HAVBinds IRES sites; restricts conformational flexibility; assists eIFsActivationReferences [168,169]
hnRNP A1HRV-2 RNA chaperone; nucleocytoplasmic shuttling; affects structureActivation, inhibition (context-dependent)Reference [177]
La autoantigenPV, HCVAssists translation initiation; stabilizes IRES conformationActivation, inhibition (HAV IRES)References [57,178,179]
UnrHRVRNA chaperone; alters IRES secondary structureActivationReference [58]
RACK1HCV, DCV, CrPV40S ribosomal subunit-associated protein, mediates IRES activityActivationReferences [180,181]
Gemin5FMDV, HCVRibosome-associated protein, competes for ITAF bindingInhibitionReference [182]
Sam68FMDV, EV71Nucleocytoplasmic shuttling, stimulates IRES activityActivationReferences [82,183]
Table 4. Structures of translation-related RNAs (IRESs, 3′ CITEs and frameshifting elements) determined by various biochemical and biophysical methods.
Table 4. Structures of translation-related RNAs (IRESs, 3′ CITEs and frameshifting elements) determined by various biochemical and biophysical methods.
MethodsStrengthsLimitationsStructures
X-ray CrystallographyAtomic-resolution 3D structures; precise protein-RNA contact sitesRequires crystal formation (difficult for large, flexible RNAs); static naturePEMV2 PTE [199], SCV PTE [198], Donggang virus dumbbells [201]
HAV IRES domain V [195], HCV IRES subdomains [174,202,203], IGR IRES [204]
Nuclear Magnetic Resonance (NMR)Probes dynamic conformational changes; solution-state interactions; binding affinitiesSize limitations for large complexes; spectral complexityEMCV IRES + eIF4G/eIF4A complex [205], EV71 IRES SLII + DMA-135 [206], EMCV IRES J-K domain [65,207], PTBP1-EMCV IRES fragment [208], HCV IRES [209,210]
Cryo-Electron Microscopy (Cryo-EM)High-resolution structures of large complexes; captures multiple conformational statesRequires large complexes; data processing complexityHCV IRES-40S [101], CrPV IRES-40S [211], EMCV IRES-eIF4G/eIF4A [205]
Biochemical MethodsMaps RNA secondary structures in solution; identifies flexible regions; guides 3D modelingDoes not provide atomic resolution; mostly limited to secondary structure informationPTE structures [199,212], BTE 3′ CITE [213], Poliovirus IRES [214,215], HCV IRES [216], FMDV IRES [217], CrPV IRES [218], Coxsackievirus B3 IRES [219]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Lu, G.; Beyene, B.G.; Camacho, J.M.; Koirala, D. Roles of RNA Structures in the Genome Translation of (+) Sense RNA Viruses. Viruses 2025, 17, 1404. https://doi.org/10.3390/v17111404

AMA Style

Lu G, Beyene BG, Camacho JM, Koirala D. Roles of RNA Structures in the Genome Translation of (+) Sense RNA Viruses. Viruses. 2025; 17(11):1404. https://doi.org/10.3390/v17111404

Chicago/Turabian Style

Lu, Guangming, Bethel G. Beyene, Joshua Miguele Camacho, and Deepak Koirala. 2025. "Roles of RNA Structures in the Genome Translation of (+) Sense RNA Viruses" Viruses 17, no. 11: 1404. https://doi.org/10.3390/v17111404

APA Style

Lu, G., Beyene, B. G., Camacho, J. M., & Koirala, D. (2025). Roles of RNA Structures in the Genome Translation of (+) Sense RNA Viruses. Viruses, 17(11), 1404. https://doi.org/10.3390/v17111404

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