Next Article in Journal
Trends of Hepatitis A Virus Infection in Poland: Assessing the Potential Impact of the COVID-19 Pandemic and War in Ukraine
Next Article in Special Issue
Factor-Dependent Internal Ribosome Entry Site and -1 Programmed Frameshifting Signal in the Bemisia-Associated Dicistrovirus 2
Previous Article in Journal
Susceptibility of Mediterranean Buffalo (Bubalus bubalis) following Experimental Infection with Lumpy Skin Disease Virus
Previous Article in Special Issue
Human Betacoronavirus OC43 Interferes with the Integrated Stress Response Pathway in Infected Cells
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Host-like RNA Elements Regulate Virus Translation

Department of Cardiovascular and Metabolic Sciences, Lerner Research Institute, Cleveland Clinic, Cleveland, OH 44195, USA
*
Authors to whom correspondence should be addressed.
Viruses 2024, 16(3), 468; https://doi.org/10.3390/v16030468
Submission received: 7 February 2024 / Revised: 14 March 2024 / Accepted: 17 March 2024 / Published: 20 March 2024

Abstract

:
Viruses are obligate, intracellular parasites that co-opt host cell machineries for propagation. Critical among these machineries are those that translate RNA into protein and their mechanisms of control. Most regulatory mechanisms effectuate their activity by targeting sequence or structural features at the RNA termini, i.e., at the 5′ or 3′ ends, including the untranslated regions (UTRs). Translation of most eukaryotic mRNAs is initiated by 5′ cap-dependent scanning. In contrast, many viruses initiate translation at internal RNA regions at internal ribosome entry sites (IRESs). Eukaryotic mRNAs often contain upstream open reading frames (uORFs) that permit condition-dependent control of downstream major ORFs. To offset genome compression and increase coding capacity, some viruses take advantage of out-of-frame overlapping uORFs (oORFs). Lacking the essential machinery of protein synthesis, for example, ribosomes and other translation factors, all viruses utilize the host apparatus to generate virus protein. In addition, some viruses exhibit RNA elements that bind host regulatory factors that are not essential components of the translation machinery. SARS-CoV-2 is a paradigm example of a virus taking advantage of multiple features of eukaryotic host translation control: the virus mimics the established human GAIT regulatory element and co-opts four host aminoacyl tRNA synthetases to form a stimulatory binding complex. Utilizing discontinuous transcription, the elements are present and identical in all SARS-CoV-2 subgenomic RNAs (and the genomic RNA). Thus, the virus exhibits a post-transcriptional regulon that improves upon analogous eukaryotic regulons, in which a family of functionally related mRNA targets contain elements that are structurally similar but lacking sequence identity. This “thrifty” virus strategy can be exploited against the virus since targeting the element can suppress the expression of all subgenomic RNAs as well as the genomic RNA. Other 3′ end viral elements include 3′-cap-independent translation elements (3′-CITEs) and 3′-tRNA-like structures. Elucidation of virus translation control elements, their binding proteins, and their mechanisms can lead to novel therapeutic approaches to reduce virus replication and pathogenicity.

1. Introduction

The flow of genetic information in biological systems is tightly regulated at all levels, that is, at replication, transcription, and translation. Post-transcriptional gene regulation allows rapid biological responses when challenged by cellular signals and environmental stimuli. A principal component of post-transcriptional gene regulation is translational control [1,2,3]. mRNA translation occurs in four stages: initiation, elongation, termination, and ribosome recycling. Translation initiation is generally the rate-limiting step in protein synthesis and is exquisitely regulated by a variety of mechanisms [1,2,4]. In eukaryotes, translation of most mRNAs is initiated via cap-dependent scanning [5,6]. Viruses are obligate, intracellular parasites that co-opt host cell machineries for propagation. Eukaryotic viruses subvert host translation machinery by employing diverse and elaborate RNA-centered strategies, often mimicking host features [7]. Utilizing such features that draw upon shared resources, viral RNAs can compete with, and even outpace, host mRNAs for selfish benefits. In this review, we present the RNA elements (Figure 1) and their protein partners that are key to shared virus and host translational control mechanisms underlying the pathogenesis of viral diseases and are targetable with a synthesis of understanding of the molecular interactions.

2. Virus 5′ End Elements

An archetype of eukaryotic translation regulation was initially described about 50 years ago in mouse liver extracts, where ferritin mRNA was induced by iron [8]. In the absence of iron, iron responsive elements (IRE) containing 25- to 30-nt stem-loop secondary structures in the ferritin and eALAS mRNA 5′-untranslated regions (UTRs) were shown to recruit iron regulatory protein 1 (IRP1) and block scanning of the 43S pre-initiation complex, and thus translation [9,10,11]. In the presence of iron, the IRP1 does not bind the IRE, thus permitting scanning and translation and representing a paradigm of 5′-end-directed translation regulation.

2.1. Internal Ribosome Entry Sites

Parallel to the landmark studies on the eukaryote IRE, pioneering studies were reported on viral translation regulation through internal ribosome entry sites (IRESs) (Figure 1). Increasing secondary structure complexity and content in 5′ UTRs of eukaryotic mRNAs, as well as viral RNAs, generally decreases translation efficiency [12]. In contrast, IRESs are RNA elements, defined by structure, sequence, or both, that recruit ribosomes at internal regions of mRNAs to initiate translation. IRESs were first discovered in picornaviruses such as poliovirus (PV) and encephalomyocarditis virus (EMCV) [13,14]. Soon afterwards, human immunodeficiency virus (HIV), hepatitis C virus (HCV), and foot and mouth disease virus (FMDV) IRESs were reported [15,16,17], and the inventory of viruses utilizing IRESs continues to expand [18]. Although the viral IRESs contain a diversity of sequences, many have similar secondary structures and initiate translation through similar mechanisms. Importantly, viruses and host cells often share translation control mechanisms, and cellular IRESs were soon reported in mRNAs encoding proteins required in physiological and pathological stress responses, e.g., apoptosis, hypoxia, and starvation [19,20,21,22,23,24]. Generally, cellular IRESs are not well conserved in either sequence or structure and are difficult to predict from sequence ab initio, and require rigorous experimental verification [25,26,27,28,29]. Notwithstanding, high-throughput methods such as mRNA display have revealed the genome-wide presence of cap-independent translation-enhancing elements [30], and nearly 10% of mammalian mRNAs contain RNA elements that potentially function as IRESs [31]. The scope and challenges in cellular IRES annotation and validation have been described in detail [32]. Mis-annotation of 5′UTRs, cryptic promoter activity, presence of 5′ introns and alternate splice sites, confusion of cap-independence with 5′end-independence, and unnatural experimental platforms have led to contested interpretation of cellular IRESs [33,34], and an essential workflow for characterizing cellular IRESs has been postulated [35]. Many viral IRESs function completely independent of the 5′-terminal cap structure, while cap-dependence of some viral IRESs and many cellular IRESs is more nuanced and context-dependent [32,36,37,38,39,40,41]. In addition to common translation initiation factors (eIFs), IRESs often recruit IRES-specific RNA-binding proteins, i.e., IRES trans-acting factors (ITAFs), for translation initiation (reviewed in [20,42,43,44]).
Many viruses, including picornaviruses, deploy a variety of strategies directing global shutdown of host cap-dependent translation while the IRESs promote translation of viral RNAs, thereby conferring a competitive advantage to the virus [7,45,46]. For example, PV infection cleaves eIF4G, preventing end-to-end circularization and translation of host mRNAs, while the IRES utilizes the cleaved carboxy-terminal fragment of eIF4G that binds eIF3 for translation initiation. In PV and EMCV infections, eIF4E is sequestered by hypophosphorylated 4E-BP, reducing cellular cap-dependent translation, without affecting IRES-dependent viral translation [47]. In both cases, viral infection essentially frees up the cellular translation machinery for self-serving protein synthesis.

2.2. Class 1 and 2 IRESs

Viral IRESs can be grouped into four classes by structure and the requirement of eIFs and ITAFs. In general, Class 1 and 2 IRESs do not bind the 40S subunit directly and require all canonical translation initiation factors, except the cap-binding protein eIF4E [48,49]. They generally present as complex secondary structures comprised of multiple domains, each containing stem-loops, bulges, and junctions. These structures are recognized by ITAFs, facilitating eIF binding and 43S pre-initiation complex recruitment. Long-range RNA–RNA interactions are also involved in the function of these IRESs [50,51]. Despite their similarity in length and multi-domain RNA structure, Class 1 and Class 2 IRESs are not identical in their secondary structures, ITAF requirements, or mechanism of initiation (reviewed in [39,48]); however, ribosomal subunit scanning after internal recruitment of the pre-initiation complex may not be considered a distinguishing factor between the two classes. For example, initiation on Class 1 IRESs can occur at the IRES-adjacent upstream codon as well as downstream at the polyprotein initiation codon [52,53,54]. Similarly, initiation on Class 2 IRESs can occur at the IRES-adjacent codon but also occurs downstream in all FMDV isolates [55,56], and in Theiler’s murine encephalomyelitis virus (TMEV), downstream initiation yields the L* protein by translation of an overlapping ORF [57,58]. Class 1 and 2 IRESs are partially or completely refractory to translation inhibition by eIF2α phosphorylation [59,60], suggesting they do not require canonical Met-tRNAiMet delivery by the ternary complex. In general, eIF2-less initiation can be facilitated by eIF2A [61], eIF2D (Ligatin) [62], eIF5B [63], or by the concerted action of MCT-1 and DENR [64]. Multiple ITAFs have been identified for the Class 1 and 2 IRESs; the specific ITAF required is variable [42,43,46]. The Class 2 EMCV IRES binds the ITAF polypyrimidine tract-binding protein (PTB), thus altering RNA conformation [65,66]. The Class 1 PV IRES binds both PTB and poly(rC)-binding protein 2 (PCBP2) as ITAFs [65,67]. Domain V in PV IRES binds cellular glycyl-tRNA synthetase (GARS), promoting association of the IRES in the mRNA binding site of the ribosome [68]. Cellular Unr protein is required for human rhinovirus (HRV) IRES activity and acts synergistically with PTB, while such synergy is absent for PV IRES, highlighting differential ITAF requirements within the same IRES class [69].

2.3. Class 3 IRESs

Class 3 IRESs, exemplified by the hepatitis C virus (HCV) of the family Flaviviridae, are present at the 5′ end of viral RNAs and contain multiple stem-loop structures organized around helical junctions and a pseudoknot with additional requirements of ITAFs, e.g., PCBP2 and La [70,71,72]. Class 3 IRESs require a small set of canonical eIFs and a notable ability for eIF2-less initiation [63,73], but can operate in either eIF2-dependent or eIF2-independent modes depending on availability [62,64,74,75,76]. Incoming 40S subunits take a “land and initiate” approach; the stem-loop structures directly capture the ribosome at the AUG start codon, thereby bypassing scanning [77,78,79]. The 40S ribosomal subunit and eIF3 bind and are stabilized by domains II and III of the IRES RNA [73,80]. The interaction of eIF3 and 40S with the IRES is structurally distinct from the cap-dependent scanning mode of translation initiation [78,79,81,82,83,84]. HCV IRES subdomain IIIb as well as a stem-loop in the specialized translation initiation element in the cellular c-JUN transcript 5′UTR share structural similarity and bind eIF3d [85]. Nonetheless, recruitment of the 60S subunit leads to 80S ribosome formation and subsequent elongation [77,86,87]. An extensive catalog of HCV IRES-40S interactions has been described [88]. This study also demonstrates how eIF5B reorients initiator tRNA prior to ribosome subunit joining, promoting eIF2-independent initiation.

2.4. Class 4 IRESs

Class 4 IRESs have been identified to date only in the Dicistroviridae family of picornaviruses, e.g., cricket paralysis virus (CrPV). The single-stranded viral RNA genome contains two open reading frames (ORFs) separated by an intergenic region (IGR) that harbors an IRES [89,90] (Figure 1). Class 4 IRESs are considered “factorless”, that is, without the requirement of eIFs and ITAFs, yet display strong translation initiation potencies in multiple cell types [91,92,93,94]. These IRESs are folded into a compact, ordered three-dimensional structure consisting of multiple pseudoknots and stem-loops that directly interact with the 40S ribosomal subunit and permit assembly of the 80S ribosome [95,96,97]. Moreover, the IRES structure can directly recruit an 80S ribosome in vitro [98]. Remarkably, the IRES initiates translation from a non-AUG start codon, bypassing the requirement for the eIF2/Met-tRNAi/GTP ternary complex [97,99,100]. An RNA pseudoknot in the Class 4 IRES mimics a cellular tRNA that occupies the ribosome P-site and directs a tRNA-less elongation step, i.e., pseudotranslocation, poising translation to start from the second elongation step [91,101,102,103,104,105]. eIF2α phosphorylation, a hallmark of viral infections and other stresses, can increase translation efficiency of these IRESs, thereby flipping this antiviral response mechanism to the benefit of the virus [106].

2.5. Hybrid IRESs and Other IRESs

In many viruses, IRESs operate outside the bounds of easy classification into the schemes outlined above. Uniquely, the hepatitis A virus IRES depends on eIF4E, requiring its association with eIF4G to unwind RNA [107,108]. A “Class 5” IRES was identified in Kobuvirus, Salivirus, and Paraturdivirus genera of Picornaviridae [109]. Various domains in this IRES class, as typified in Aichivirus, are shared with both Class 1 and Class 2 IRESs, with a unique requirement of cellular DHX29 protein as an ITAF [109]. IRESs that are not classified within the four common types are present in pegiviruses, pestiviruses, and hepaciviruses [110]. Retroviral IRES function, including HIV-1 IRES, is debated as viral transcripts are capped [38,111,112]. In addition, a specialized translation mechanism of HIV-1 mRNAs sustained during eIF4E repression can be directed by a hypermethylated cap instead of being cap-independent, requiring interaction of the primer binding site in viral 5′UTR with cellular nuclear RNA helicase A (RHA)/DHX9 [113]. The IRES in the iflavirus Ectropis obliqua picorna-like virus (EoPV) is similar in several functional and structural aspects to Class 2 IRESs in cardioviruses and aphthoviruses; however, unlike Class 2 IRESs, it lacks a GNRA motif, has a functionally redundant RAAA motif, and utilizes a non-homologous initiation codon [114]. The dicistrovirus CrPV 5′UTR-IRES assembles a functional initiation complex utilizing an upstream start codon within the IRES [115]. In another dicistrovirus, the Halastavi árva virus (HalV), the 5′UTR IRES utilizes an unusual retrograde scanning of the 43S complex for initiation [36].
In summary, IRES structures are diverse and employ equally diverse mechanisms with limited common principles that serve as a framework for features specific to a single IRES or IRES class. Focused reviews contain further information on viral IRESs [39,46,48,49,116]. Beyond IRESs, viral and cellular transcripts can share specialized themes in cap-dependent translation initiation as well, e.g., cis-acting elements in the 5′UTR of transcripts from Mononegavirales family members such as vesicular stomatitis virus (VSV), measles virus, and rabies virus, as well as cellular transcripts such as DDR2, interact uniquely with RPL40 (ribosomal protein 40, large subunit) for cap-dependent translation initiation, suggesting these viruses might have usurped an endogenous translation pathway [117]. It remains to be seen if other alternative pathways of cap-dependent cellular translation initiation, e.g., eIF3d-dependent [118] and eIF4EHP-dependent [119], are utilized by multiple viruses.

3. Upstream Open Reading Frames

3.1. Translation Control by Upstream Open Reading Frames

A series of landmark studies revealed translation repression of GCN4 mRNA in brewer’s yeast (Saccharomyces cerevisiae) through initiation events at small ORFs in its 5′UTR [120,121,122,123]. Upstream open reading frames (uORFs) are cis-regulatory RNA elements prevalent in eukaryotic mRNAs that regulate translation initiation of downstream coding sequences (CDSs) (Figure 1). The pre-initiation complex (PIC) scans the 5′ UTR for a start codon until it recognizes an upstream AUG (uAUG), thereby permitting assembly of 80S ribosomes at the uORF. uORFs serve as “roadblocks”, translation elongation following uORF recognition prevents PIC scanning to the downstream AUG of the main ORF. The Kozak sequence context of an uAUG, as in a canonical CDS, is a critical determinant of translation initiation of an uORF (reviewed in [124]). A weak context at the uAUG leads to “leaky scanning” by 43S PIC and initiation at the main ORF, while a strong context results in translation elongation at the uORF, necessitating re-initiation events by 40S subunits at the downstream AUG of the main ORF. Non-canonical initiation codons, e.g., CUG, GUG, and UUG, can also direct upstream initiation at uORFs [125,126]. uORFs attenuate main ORF translation by multiple mechanisms: by the “first dibs” scheme of translation initiation, by ribosome dissociation at the uORF stop codon, by stalling-mediated mRNA decay, and by encoding regulatory peptides [4,127,128,129]. uORF-mediated repression is widespread across the mammalian transcriptome [130,131,132] and is associated with diseases and stress responses (reviewed in [133,134,135,136]). For example, the integrated stress response depletes pre-initiation complexes to specifically de-repress translation from uORF-containing mRNAs [137,138,139]. A PIC reinitiates at a main ORF initiation site only when an uORF does not overlap with the main ORF. In contrast, an upstream AUG preceding the main ORF start codon can be in an overlapping ORF, with 2 further subclasses: out-of-frame overlapping uORFs (oORFs), the stop codons of which are downstream of main ORF AUGs and in different reading frames, and N-terminal extensions, which are basically overlapping uORFs in-frame with the main ORF [140,141,142,143,144] (Figure 1). The downstream CDS can be engaged by the re-initiating PIC only when the translated uORFs do not overlap with CDSs [145], or in special cases through a TURBS (termination upstream ribosome binding site) RNA element at the end of the preceding ORF (reviewed in [145,146], see below) (Figure 1). In addition to leaky scanning, bi-directional scanning of the PIC between closely spaced start codons, as in oORFs, can influence start site selection [147,148].

3.2. Virus Utilization of uORFs and uAUGs

Viruses take advantage of oORFs to offset genome compression and increase coding capacity [149]. Likewise, utilization of uORFs and uAUGs is commonplace in DNA viruses as well as in positive- and negative-sense RNA viruses [150,151]. Examples representing multiple mechanisms and viruses are summarized here.
Hepatitis B virus (HBV) polymerase ORF is preceded by an overlapping core ORF and by a more upstream AUG in the viral pre-genomic RNA [152]. Translation from the uAUG represses translation of the core ORF while allowing reinitiation of translation at the polymerase ORF [153,154]. Translation of the Human Cytomegalovirus (HCMV) gpUL4 ORF is regulated by an uORF [155]. Kaposi’s sarcoma-associated herpesvirus (KSHV) ORF35 and ORF36 are translated from a polycistronic transcript regulated by a pair of uORFs. The second uORF overlaps with ORF35 and allows translation of ORF36 by a reinitiation mechanism, which is essential for viral propagation [156,157]. An uORF in the ebolavirus (EBOV) L gene suppresses translation of the L ORF under normal conditions and increases it under stress to maintain optimal polymerase activity, and uORF mutations attenuate EBOV growth [158]. Human immunodeficiency virus type 1 (HIV1) tat ORF is followed by rev and nef ORFs, and translation of upstream tat strongly represses Rev production [159]. A minimal uORF, containing only a start and a stop codon within the HIV1 vpu leader, permits efficient translation initiation at the downstream env start codon in viral vpu-env bicistronic mRNAs [160]. In arteriviruses, an uORF is present in the 5′-leader of the genomic RNA (gRNA), as well as in all sub-genomic RNAs (sgRNAs). In related coronaviruses, the uORF maps downstream of the genomic leader and is present only in the gRNA [161]. Bovine coronaviruses are an exception in which the uORF is utilized in all sgRNAs during persistent infection, causing translation attenuation [162]. Simian immunodeficiency virus (SIV) Rev and Env proteins are regulated by up to five uAUGs in multiple splice variants [163,164]. uORF translation enhances ribosomal shunting in prototype foamy virus genomic RNA [165] and in rice tungro virus [166], in a way that enables the mRNA or viral RNA to escape translation inhibition.
Termination-reinitiation events in uORF/main ORF pairs are regulated by TURBS elements extensively characterized in viruses of the Caliciviridae family and in influenza B virus from the Orthomyxoviridae family [167,168,169]. In these cases, the uORF stop codon and downstream ORF start codon are in close proximity, out-of-frame, or overlap in consecutive nucleotides. The viral TURBS regulates reinitiation through base-pairing with 18S rRNA and tethering of the post-termination 40S subunit (reviewed in [145,146]). It has been suggested that uORFs in the 5′UTR regulate translation of cellular SLAMF1 mRNA utilizing a TURBS-like element, highlighting shared strategies in eukaryotic translational control [170]. Additionally, functional polypeptides encoded by viral uORFs can regulate virulence and tropism. For example, a conserved uORF overlaps the viral polyprotein ORF in enteroviruses, including echovirus 7 and poliovirus 1, encoding a transmembrane protein that facilitates virus growth in gut epithelial cells [53].

4. 3′ End Elements

4.1. GAIT and VAIT Elements

The Gamma-interferon-Activated Inhibitor of Translation (GAIT) RNA element was first identified in the 3′UTR of human ceruloplasmin mRNA as an inducible, translation repressive element [171]. Subsequent studies observed similar structurally conserved GAIT elements in VEGFA, DAPK, and ZIPK mRNAs. To date, translation-repressive GAIT elements have been reported in multiple inflammation-related human mRNAs [171,172,173,174]. The GAIT RNA element is characterized by a split stem-loop secondary structure, with generally conserved A and U residues (absent in the ZIPK GAIT element [172]) in an asymmetric bulge separating the stems. The GAIT element recruits a gamma-interferon-inducible, heterotetrameric complex in which the direct RNA-interacting constituent is the bifunctional glutamyl-prolyl tRNA synthetase EPRS1 [175,176]. EPRS1 contains two catalytic domains joined by a linker containing three repeats of an atypical RNA-binding domain termed the WHEP domain, also found in human tryptophanyl, histidyl, glycyl, and methionyl tRNA synthetases. Three other constituents of the human GAIT complex are ribosomal protein L13a, heterogenous ribonucleoprotein Q (or NSAP1), and GAPDH. Eukaryotic translation occurs in a closed loop where the 5′ and 3′ ends of translated RNAs are generally bridged by eIF4G and PABP interactions [177,178,179]. L13a is central to the repressive function of the GAIT complex, as it blocks eIF4G from recruiting the pre-initiation complex [180].
The GAIT system is a classical archetype of a post-transcriptional regulon in which families of functionally related mRNAs are co-regulated by specific RNA-binding proteins that target similar sequences or structural elements [181,182]. Mimicking host RNA structures, GAIT-like RNA elements have been described in the RNA viruses RSV (respiratory syncytial virus) and the pig alphacoronavirus TGEV (transmissible gastroenteritis virus); both repress viral RNA translation in response to viral cues [183,184]. The cellular GAIT element, the RSV GAIT-like element binds L13a and has been termed a VAIT or Virus-Activated Inhibitor of Translation element [184] (Figure 1). Two coding-region VAITs have been reported in SARS-CoV-2, one in ORF1a of the genomic RNA and another in the coding region of Spike subgenomic (sgRNA) [185]. Unlike the VAITs, the inhibitory TGEV GAIT-like element does not bind L13a but binds two aminoacyl-tRNA synthetases (aaRSs), EPRS1 and RARS1, showcasing L13a-independent translation repression [183], and suggesting similar RNA structural motifs can recruit distinct but functionally similar machineries.

4.2. Sarbecoviral Pan-End Activating RNA Element

We reported a novel 39-nucleotide GAIT-like RNA element in the 3′-end of SARS-CoV-2 genomic and sgRNAs [186] (Figure 1). Discontinuous transcription of the viral gRNA generates an ensemble of nested 3′-co-terminal sgRNAs that contain 5′-leader and 3′-end sequences identical to each other and to the genomic RNA (gRNA) (Figure 2). Thus, the element defines a unique post-transcriptional regulon distinct from non-viral systems, in which the elements are structurally related but not identical in sequence [187]. The RNA sequence is conserved in viruses of the subgenus Sarbecovirus of the genus Betacoronavirus, including SARS-CoV-1, suggesting an invariant function. Host-derived insulin and IFN-γ, agents associated with COVID-19 severity and outcome, singly or additively with SARS-CoV-2 spike subunit 1 that contains the host ACE2-receptor binding domain, increase sgRNA expression and translation only when the cis-element is intact [186]. Therefore, the element was termed the Sarbecoviral Pan-End Activating RNA (SPEAR) element. Strikingly, EPRS1 binds SPEAR as part of an unconventional, inducible Tetra-Aminoacyl-tRNA synthetase Sarbecoviral RNA-Interacting (TASRI) complex, containing three other aaRSs: arginyl-, lysyl- and methionyl-tRNA synthetases (RARS1, KARS1, and MARS1, respectively). Of note, SPEAR, such as the repressive TGEV GAIT-like element, does not bind L13a, L13a-interacting RSV, or SARS-CoV-2 VAITs [184,186]. Binding of RARS1 and EPRS1 to the repressive TGEV element [183] suggests the translation-enhancing function of SPEAR might depend on the other binding partners, i.e., KARS1 or MARS1, or both. Potential internal initiation at ORF10, a 3′-end co-terminal feature newly acquired in SARS-CoV-2, appears to play a critical role in SPEAR-mediated induction of viral sgRNA expression [186]. The SPEAR element and TASRI complex together form a host system hijacked by SARS-CoV-2 to direct an outcome, i.e., translation enhancement, opposite to the original host function of translation inhibition.
Surprisingly, the SPEAR element at the 3′-end of gRNA supports ribosomal frameshifting in reporter assays, but only in the presence of the genomic 5′UTR [186], indicative of a 5′-end-dependent role of the SPEAR element. Circularization of eukaryotic mRNAs by interactions of proteins bound to both termini have been proposed to augment translation efficiency, possibly by facilitating ribosome transfer to the initiation site following a round of elongation [177]. In an early example of mRNA circularization influencing translation control, rather than translation efficiency, translational silencing by the 3′UTR GAIT element was shown to require the elements of mRNA circularization, including eIF4G [188,189]. In an alternative control mechanism that takes advantage of the closed loop, the 3′-Cap-Independent Translation Elements (CITEs, described in detail in Section 4.3), structurally diverse elements in the 3′UTRs of many plant viral genomes facilitate translation by recruiting initiation factors, ribosomes, and ITAFs to the 3′UTR, which exhibit trans-terminal effects on initiation at the 5′-end [116,190,191,192,193,194,195]. Indeed, SARS-CoV-2 initiates a global shutdown of translation, possibly necessitating a CITE-like function of SPEAR [196]. A proposed closed-loop mechanism placing the SPEAR element in proximity of the start codon across all viral sgRNAs awaits verification [197]. The 5′UTR of SARS-CoV-2 gRNA exhibits cap-independent translation activity; however, the smaller 5′-leaders of sgRNAs are eIF4E-dependent [198]. Since the SPEAR element is present in all SARS-CoV-2 RNAs and might facilitate 5′-3′ end communication, the difference in translation-initiation modes between 5′UTR of gRNA and 5′-leaders of sgRNAs can further indicate diverse roles of SPEAR in the 5′-3′ communications of gRNA compared to sgRNAs (Figure 2).
A critical feature of the discontinuous transcription program utilized by SARS-CoV-2 is the generation of an ensemble of nested 3′-co-terminal sgRNAs containing 5′ leader and 3′ end sequences identical to each other and to the genomic sequence [186]. Thus, the system forms a “post-transcriptional regulon” analogous to those found in eukaryotes, in which families of functionally related mRNAs are co-regulated by specific RNA-binding proteins that target similar sequences or structural elements [181]. However, the viral system is distinct in that the RNAs regulated are contained within the same genetic unit, versus scattered among chromosomes in eukaryotic systems. Also, the element is identical in the RNAs of the viral system, compared to the similarity of sequence or structure in eukaryotic systems. The “thrifty” virus mechanism that compacts the genome, i.e., reusing the genomic termini in every sgRNA, can be turned against the virus as a therapeutic strategy. Thus, interfering with the SPEAR-TASRI complex interaction is an attractive antiviral strategy to target the entire SPEAR regulon. Importantly, interception of the SPEAR element/TASRI complex interaction with peptide-conjugated phosphorodiamidate morpholine oligonucleotides (PPMOs) antisense to the SPEAR element markedly reduces SARS-CoV-2 replication in Caco-2-hACE2 cells [186].

4.3. 3′-Cap-Independent Translation Elements

The trans-terminal-acting 3′-CITEs are structurally diverse elements in seven identified classes [39] (Figure 1). The 5′ region of the viral RNA is essential for the function of certain 3′-CITEs through 5′-3′ communication, even if the cap is not [199,200,201] (Figure 1). These elements can compete with cellular mRNAs for the eIFs; e.g., barley yellow dwarf virus translation element (BTE) binds eIF4G [202,203,204] and panicum mosaic virus-like translation enhancers (PTEs) bind eIF4E while base-paired with the 5′UTR [205,206]. Crystal structures have shown that a guanosine is extruded from a G-rich stretch in the PTE of the pea-enation mosaic virus (PEMV2), providing a structural basis for eIF4E recognition [207]. Similar studies in the saguaro cactus virus (SCV) revealed a flipped-out G residue ready to dock into the 5′ cap-binding pocket of eIF4E, highlighting 5′ cap mimicry by a 3′-CITE [208]. 3′-CITEs in T-shaped structures (TSSs) in TCV and cardamine chlorotic fleck virus RNAs can recycle post-termination ribosomes or ribosomal subunits directly to the 5′-end [209,210]. The 55-nt Cucurbit aphid-borne yellows virus-Xinjiang-like translation element (CXTE) functions in eIF4E-depleted lysate and is enhanced by the viral 5′UTR; however, detailed molecular mechanisms remain unknown [211]. Interfamilial recombination of 3′-CITEs between viruses has been observed, and they can act in combination with other elements, e.g., the melon necrotic spot virus has a CXTE and an I-shaped structure (ISS) [211], and the PTE and an upstream TSS in PEMV2 [212]. Furthermore, 3′-CITEs are conformationally dynamic [213], e.g., the TCV TSS is disassembled through viral RNA-dependent RNA polymerase binding, a requirement for replication [214]. Structural rearrangement is also proposed for the ORF10-dependent function of the SPEAR element in SARS-CoV-2 sgRNA 3′-end, comparable to 3′-CITEs [186,197]. The TSSs in TCV and PEMV2 are predicted to form internal tRNA-like structures (TLS), ubiquitous host molecules linking the amino acid and RNA worlds [192].

4.4. 3′-tRNA-Like Structures

The plant RNA virus Tobamovirus and Tymovirus genomes contain 3′-terminal TLSs that control their replication [215,216,217] (Figure 1). Like the aaRS-interacting GAIT and TGEV elements, these 3′-TLSs recruit specific aaRSs and are aminoacylated with histidine (e.g., tobacco mosaic virus, TMV), tyrosine (e.g., brome mosaic virus, BMV), or valine (e.g., turnip yellow mosaic virus, TYMV) to stabilize viral RNAs, enhance 5′-cap-dependent translation by communicating with 5′UTR elements in the closed-loop model of translation initiation, and possibly recruit eIFs and ribosomes [215,218,219]. The TLS in TYMV interacts with an upstream pseudoknot domain, possibly constituting a longer, modular structural element that increases the stability of this RNA substrate for valylation [220]. It remains contentious whether the TLS delivers valine as the first amino acid for viral protein synthesis [221,222]. The BMV TLS binds tyrosyl-tRNA synthetase in a conformation that markedly differs from a tRNA [223], further illustrating the dynamics of viral RNA structures. The TLS of the Hordeivirus genus in Virgaviridae, although structurally distinct from BMV TLS, can be tyrsylated, and, more importantly, a chimeric hordeivirus-BMV TLS can be similarly aminoacylated, illustrating modularity in TLSs [224]. Likewise, tRNA mimicry is seen in 5′-localized structures in viral RNAs, e.g., poliovirus IRES domain V resembles glycine tRNA and binds glycyl-tRNA synthetase for translation initiation [68].

5. Conclusions

Viruses employ a diverse array of host-directed translational strategies for efficient expression of viral proteins. Lacking the essential machinery of protein synthesis, for example, ribosomes and other translation factors, all viruses take advantage of the host apparatus to generate virus protein. Layered above the absolute requirement for the translation machinery, viruses have adopted certain features of host translation control to the benefit of the virus. In some cases, the mechanism itself is imitated; for example, virus utilization of IRESs and uORFs is used to increase coding capacity while reducing genome size. Some viruses have co-opted eukaryotic sequence or structural elements generally employed for condition-dependent regulation, often translation repression. Viruses can incorporate similar elements to recruit essential constituents of the eukaryote translation apparatus, generally to enhance translation initiation. Other virus elements recruit host regulatory factors that are not components of the translation machinery. The utilization of multiple aaRSs as trans-acting factors interacting with a variety of virus elements forms an unexpected molecular leitmotif. SARS-CoV-2 is a paradigm example of a virus conjoining several distinct features of eukaryotic host translation control: the virus mimics an established regulatory element and co-opts four aaRSs to form a stimulatory binding complex. Utilizing discontinuous transcription, identical elements are present in all SARS-CoV-2 sgRNAs (and the gRNA). Thus, the virus post-transcriptional regulon improves upon comparable regulons in eukaryotes, in which related but different elements are present in mRNA targets scattered among the chromosomes. The same strategy exploited by the virus can be pitted against the virus, as targeting the element can simultaneously inhibit the expression of all sgRNAs. In-depth elucidation of virus translation control elements, their binding proteins, and their mechanisms can lead to novel therapeutic approaches to reduce virus replication and pathogenicity.

Author Contributions

Conceptualization, D.K. and P.L.F.; writing—original draft preparation, D.K.; writing—review and editing, D.K. and P.L.F.; funding acquisition, P.L.F. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by N.I.H. grants R01 DK130377 and R01 NS124547 (to P.L.F.).

Institutional Review Board Statement

The study did not require ethical approval.

Informed Consent Statement

The study did not require informed consent.

Data Availability Statement

There are no original data in the paper.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. 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] [PubMed]
  2. Sonenberg, N.; Hinnebusch, A.G. Regulation of translation initiation in eukaryotes: Mechanisms and biological targets. Cell 2009, 136, 731–745. [Google Scholar] [CrossRef] [PubMed]
  3. Gebauer, F.; Hentze, M.W. Molecular mechanisms of translational control. Nat. Rev. Mol. Cell Biol. 2004, 5, 827–835. [Google Scholar] [CrossRef] [PubMed]
  4. Hinnebusch, A.G.; Ivanov, I.P.; Sonenberg, N. Translational control by 5′-untranslated regions of eukaryotic mRNAs. Science 2016, 352, 1413–1416. [Google Scholar] [CrossRef] [PubMed]
  5. Hinnebusch, A.G. The scanning mechanism of eukaryotic translation initiation. Annu. Rev. Biochem. 2014, 83, 779–812. [Google Scholar] [CrossRef] [PubMed]
  6. Shirokikh, N.E.; Preiss, T. Translation initiation by cap-dependent ribosome recruitment: Recent insights and open questions. Wiley Interdiscip. Rev. RNA 2018, 9, e1473. [Google Scholar] [CrossRef] [PubMed]
  7. Stern-Ginossar, N.; Thompson, S.R.; Mathews, M.B.; Mohr, I. Translational control in virus-infected cells. Cold Spring Harb. Perspect. Biol. 2019, 11, a033001. [Google Scholar] [CrossRef] [PubMed]
  8. Zahringer, J.; Baliga, B.S.; Munro, H.N. Novel mechanism for translational control in regulation of ferritin synthesis by iron. Proc. Natl. Acad. Sci. USA 1976, 73, 857–861. [Google Scholar] [CrossRef]
  9. Hentze, M.W.; Caughman, S.W.; Rouault, T.A.; Barriocanal, J.G.; Dancis, A.; Harford, J.B.; Klausner, R.D. Identification of the iron-responsive element for the translational regulation of human ferritin mRNA. Science 1987, 238, 1570–1573. [Google Scholar] [CrossRef]
  10. Hentze, M.W.; Rouault, T.A.; Caughman, S.W.; Dancis, A.; Harford, J.B.; Klausner, R.D. A cis-acting element is necessary and sufficient for translational regulation of human ferritin expression in response to iron. Proc. Natl. Acad. Sci. USA 1987, 84, 6730–6734. [Google Scholar] [CrossRef]
  11. Gray, N.K.; Hentze, M.W. Iron regulatory protein prevents binding of the 43S translation pre-initiation complex to ferritin and eALAS mRNAs. EMBO J. 1994, 13, 3882–3891. [Google Scholar] [CrossRef] [PubMed]
  12. Pelletier, J.; Sonenberg, N. Insertion mutagenesis to increase secondary structure within the 5′ noncoding region of a eukaryotic mRNA reduces translational efficiency. Cell 1985, 40, 515–526. [Google Scholar] [CrossRef]
  13. 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] [PubMed]
  14. Pelletier, J.; Sonenberg, N. Internal initiation of translation of eukaryotic mRNA directed by a sequence derived from poliovirus RNA. Nature 1988, 334, 320–325. [Google Scholar] [CrossRef] [PubMed]
  15. Belsham, G.J.; Brangwyn, J.K. A region of the 5′ noncoding region of foot-and-mouth disease virus RNA directs efficient internal initiation of protein synthesis within cells: Involvement with the role of L protease in translational control. J. Virol. 1990, 64, 5389–5395. [Google Scholar] [CrossRef]
  16. 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] [PubMed]
  17. Buck, C.B.; Shen, X.; Egan, M.A.; Pierson, T.C.; Walker, C.M.; Siliciano, R.F. The human immunodeficiency virus type 1 gag gene encodes an internal ribosome entry site. J. Virol. 2001, 75, 181–191. [Google Scholar] [CrossRef]
  18. Zhao, J.; Li, Y.; Wang, C.; Zhang, H.; Zhang, H.; Jiang, B.; Guo, X.; Song, X. IRESbase: A comprehensive database of experimentally validated internal ribosome entry sites. Genom. Proteom. Bioinform. 2020, 18, 129–139. [Google Scholar] [CrossRef]
  19. Komar, A.A.; Hatzoglou, M. Cellular IRES-mediated translation: The war of ITAFs in pathophysiological states. Cell Cycle 2011, 10, 229–240. [Google Scholar] [CrossRef]
  20. King, H.A.; Cobbold, L.C.; Willis, A.E. The role of IRES trans-acting factors in regulating translation initiation. Biochem. Soc. Trans. 2010, 38, 1581–1586. [Google Scholar] [CrossRef]
  21. Spriggs, K.A.; Bushell, M.; Willis, A.E. Translational regulation of gene expression during conditions of cell stress. Mol. Cell 2010, 40, 228–237. [Google Scholar] [CrossRef] [PubMed]
  22. Yoon, A.; Peng, G.; Brandenburger, Y.; Zollo, O.; Xu, W.; Rego, E.; Ruggero, D. Impaired control of IRES-mediated translation in X-linked dyskeratosis congenita. Science 2006, 312, 902–906. [Google Scholar] [CrossRef] [PubMed]
  23. Leprivier, G.; Rotblat, B.; Khan, D.; Jan, E.; Sorensen, P.H. Stress-mediated translational control in cancer cells. Biochim. Biophys. Acta 2015, 1849, 845–860. [Google Scholar] [CrossRef] [PubMed]
  24. Kovalski, J.R.; Kuzuoglu-Ozturk, D.; Ruggero, D. Protein synthesis control in cancer: Selectivity and therapeutic targeting. EMBO J. 2022, 41, e109823. [Google Scholar] [CrossRef] [PubMed]
  25. Martinez-Salas, E.; Pineiro, D.; Fernandez, N. Alternative mechanisms to initiate translation in eukaryotic mRNAs. Comp. Funct. Genom. 2012, 2012, 391546. [Google Scholar] [CrossRef] [PubMed]
  26. Dotu, I.; Lozano, G.; Clote, P.; Martinez-Salas, E. Using RNA inverse folding to identify IRES-like structural subdomains. RNA Biol. 2013, 10, 1842–1852. [Google Scholar] [CrossRef] [PubMed]
  27. Van Eden, M.E.; Byrd, M.P.; Sherrill, K.W.; Lloyd, R.E. Demonstrating internal ribosome entry sites in eukaryotic mRNAs using stringent RNA test procedures. RNA 2004, 10, 720–730. [Google Scholar] [CrossRef] [PubMed]
  28. Kolekar, P.; Pataskar, A.; Kulkarni-Kale, U.; Pal, J.; Kulkarni, A. IRESPred: Web server for prediction of cellular and viral internal ribosome entry site (IRES). Sci. Rep. 2016, 6, 27436. [Google Scholar] [CrossRef]
  29. Gritsenko, A.A.; Weingarten-Gabbay, S.; Elias-Kirma, S.; Nir, R.; de Ridder, D.; Segal, E. Sequence features of viral and human Internal Ribosome Entry Sites predictive of their activity. PLoS Comput. Biol. 2017, 13, e1005734. [Google Scholar] [CrossRef]
  30. Wellensiek, B.P.; Larsen, A.C.; Stephens, B.; Kukurba, K.; Waern, K.; Briones, N.; Liu, L.; Snyder, M.; Jacobs, B.L.; Kumar, S.; et al. Genome-wide profiling of human cap-independent translation-enhancing elements. Nat. Methods 2013, 10, 747–750. [Google Scholar] [CrossRef]
  31. Weingarten-Gabbay, S.; Elias-Kirma, S.; Nir, R.; Gritsenko, A.A.; Stern-Ginossar, N.; Yakhini, Z.; Weinberger, A.; Segal, E. Comparative genetics. Systematic discovery of cap-independent translation sequences in human and viral genomes. Science 2016, 351, aad4939. [Google Scholar] [CrossRef] [PubMed]
  32. Lozano, G.; Francisco-Velilla, R.; Martinez-Salas, E. Deconstructing internal ribosome entry site elements: An update of structural motifs and functional divergences. Open Biol. 2018, 8, 180155. [Google Scholar] [CrossRef]
  33. Xue, S.; Tian, S.; Fujii, K.; Kladwang, W.; Das, R.; Barna, M. RNA regulons in Hox 5′ UTRs confer ribosome specificity to gene regulation. Nature 2015, 517, 33–38. [Google Scholar] [CrossRef] [PubMed]
  34. Akirtava, C.; May, G.E.; McManus, C.J. False-positive IRESes from Hoxa9 and other genes resulting from errors in mammalian 5′ UTR annotations. Proc. Natl. Acad. Sci. USA 2022, 119, e2122170119. [Google Scholar] [CrossRef] [PubMed]
  35. Terenin, I.M.; Smirnova, V.V.; Andreev, D.E.; Dmitriev, S.E.; Shatsky, I.N. A researcher’s guide to the galaxy of IRESs. Cell. Mol. Life Sci. 2017, 74, 1431–1455. [Google Scholar] [CrossRef] [PubMed]
  36. Abaeva, I.S.; Pestova, T.V.; Hellen, C.U. Attachment of ribosomal complexes and retrograde scanning during initiation on the Halastavi arva virus IRES. Nucleic Acids Res. 2016, 44, 2362–2377. [Google Scholar] [CrossRef] [PubMed]
  37. Thompson, S.R. Tricks an IRES uses to enslave ribosomes. Trends Microbiol. 2012, 20, 558–566. [Google Scholar] [CrossRef] [PubMed]
  38. Smirnova, V.V.; Terenin, I.M.; Khutornenko, A.A.; Andreev, D.E.; Dmitriev, S.E.; Shatsky, I.N. Does HIV-1 mRNA 5′-untranslated region bear an internal ribosome entry site? Biochimie 2016, 121, 228–237. [Google Scholar] [CrossRef]
  39. Jaafar, Z.A.; Kieft, J.S. Viral RNA structure-based strategies to manipulate translation. Nat. Rev. Microbiol. 2019, 17, 110–123. [Google Scholar] [CrossRef]
  40. Jackson, R.J. The current status of vertebrate cellular mRNA IRESs. Cold Spring Harb. Perspect. Biol. 2013, 5, a011569. [Google Scholar] [CrossRef]
  41. Gilbert, W.V. Alternative ways to think about cellular internal ribosome entry. J. Biol. Chem. 2010, 285, 29033–29038. [Google Scholar] [CrossRef] [PubMed]
  42. Lopez-Ulloa, B.; Fuentes, Y.; Pizarro-Ortega, M.S.; Lopez-Lastra, M. RNA-binding proteins as regulators of internal initiation of viral mRNA translation. Viruses 2022, 14, 188. [Google Scholar] [CrossRef] [PubMed]
  43. Lee, K.M.; Chen, C.J.; Shih, S.R. Regulation mechanisms of viral IRES-driven translation. Trends Microbiol. 2017, 25, 546–561. [Google Scholar] [CrossRef] [PubMed]
  44. 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] [PubMed]
  45. Walsh, D.; Mathews, M.B.; Mohr, I. Tinkering with translation: Protein synthesis in virus-infected cells. Cold Spring Harb. Perspect. Biol. 2013, 5, a012351. [Google Scholar] [CrossRef]
  46. Martinez-Salas, E.; Francisco-Velilla, R.; Fernandez-Chamorro, J.; Lozano, G.; Diaz-Toledano, R. Picornavirus IRES elements: RNA structure and host protein interactions. Virus Res. 2015, 206, 62–73. [Google Scholar] [CrossRef]
  47. Gingras, A.C.; Svitkin, Y.; Belsham, G.J.; Pause, A.; Sonenberg, N. Activation of the translational suppressor 4E-BP1 following infection with encephalomyocarditis virus and poliovirus. Proc. Natl. Acad. Sci. USA 1996, 93, 5578–5583. [Google Scholar] [CrossRef]
  48. Lozano, G.; Martinez-Salas, E. Structural insights into viral IRES-dependent translation mechanisms. Curr. Opin. Virol. 2015, 12, 113–120. [Google Scholar] [CrossRef]
  49. Mailliot, J.; Martin, F. Viral internal ribosomal entry sites: Four classes for one goal. Wiley Interdiscip. Rev. RNA 2018, 9, e1458. [Google Scholar] [CrossRef]
  50. Fernandez-Miragall, O.; Ramos, R.; Ramajo, J.; Martinez-Salas, E. Evidence of reciprocal tertiary interactions between conserved motifs involved in organizing RNA structure essential for internal initiation of translation. RNA 2006, 12, 223–234. [Google Scholar] [CrossRef]
  51. Fernandez, N.; Fernandez-Miragall, O.; Ramajo, J.; Garcia-Sacristan, A.; Bellora, N.; Eyras, E.; Briones, C.; Martinez-Salas, E. Structural basis for the biological relevance of the invariant apical stem in IRES-mediated translation. Nucleic Acids Res. 2011, 39, 8572–8585. [Google Scholar] [CrossRef] [PubMed]
  52. Pestova, T.V.; Hellen, C.U.; Wimmer, E. A conserved AUG triplet in the 5′ nontranslated region of poliovirus can function as an initiation codon in vitro and in vivo. Virology 1994, 204, 729–737. [Google Scholar] [CrossRef] [PubMed]
  53. Lulla, V.; Dinan, A.M.; Hosmillo, M.; Chaudhry, Y.; Sherry, L.; Irigoyen, N.; Nayak, K.M.; Stonehouse, N.J.; Zilbauer, M.; Goodfellow, I.; et al. An upstream protein-coding region in enteroviruses modulates virus infection in gut epithelial cells. Nat. Microbiol. 2019, 4, 280–292. [Google Scholar] [CrossRef] [PubMed]
  54. Guo, H.; Li, Y.; Liu, G.; Jiang, Y.; Shen, S.; Bi, R.; Huang, H.; Cheng, T.; Wang, C.; Wei, W. A second open reading frame in human enterovirus determines viral replication in intestinal epithelial cells. Nat. Commun. 2019, 10, 4066. [Google Scholar] [CrossRef] [PubMed]
  55. Clarke, B.E.; Sangar, D.V.; Burroughs, J.N.; Newton, S.E.; Carroll, A.R.; Rowlands, D.J. Two initiation sites for foot-and-mouth disease virus polyprotein in vivo. J. Gen. Virol. 1985, 66 Pt 12, 2615–2626. [Google Scholar] [CrossRef] [PubMed]
  56. 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]
  57. Yamasaki, K.; Weihl, C.C.; Roos, R.P. Alternative translation initiation of Theiler’s murine encephalomyelitis virus. J. Virol. 1999, 73, 8519–8526. [Google Scholar] [CrossRef]
  58. van Eyll, O.; Michiels, T. Non-AUG-initiated internal translation of the L* protein of Theiler’s virus and importance of this protein for viral persistence. J. Virol. 2002, 76, 10665–10673. [Google Scholar] [CrossRef]
  59. Meurs, E.F.; Watanabe, Y.; Kadereit, S.; Barber, G.N.; Katze, M.G.; Chong, K.; Williams, B.R.; Hovanessian, A.G. Constitutive expression of human double-stranded RNA-activated p68 kinase in murine cells mediates phosphorylation of eukaryotic initiation factor 2 and partial resistance to encephalomyocarditis virus growth. J. Virol. 1992, 66, 5805–5814. [Google Scholar] [CrossRef]
  60. 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]
  61. Ventoso, I.; Sanz, M.A.; Molina, S.; Berlanga, J.J.; Carrasco, L.; Esteban, M. Translational resistance of late alphavirus mRNA to eIF2alpha phosphorylation: A strategy to overcome the antiviral effect of protein kinase PKR. Genes Dev. 2006, 20, 87–100. [Google Scholar] [CrossRef] [PubMed]
  62. Dmitriev, S.E.; Terenin, I.M.; Andreev, D.E.; Ivanov, P.A.; Dunaevsky, J.E.; Merrick, W.C.; Shatsky, I.N. GTP-independent tRNA delivery to the ribosomal P-site by a novel eukaryotic translation factor. J. Biol. Chem. 2010, 285, 26779–26787. [Google Scholar] [CrossRef] [PubMed]
  63. 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]
  64. 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]
  65. Kafasla, P.; Morgner, N.; Poyry, T.A.; Curry, S.; Robinson, C.V.; Jackson, R.J. Polypyrimidine tract binding protein stabilizes the encephalomyocarditis virus IRES structure via binding multiple sites in a unique orientation. Mol. Cell 2009, 34, 556–568. [Google Scholar] [CrossRef] [PubMed]
  66. Borovjagin, A.; Pestova, T.; Shatsky, I. Pyrimidine tract binding protein strongly stimulates in vitro encephalomyocarditis virus RNA translation at the level of preinitiation complex formation. FEBS Lett. 1994, 351, 299–302. [Google Scholar] [CrossRef] [PubMed]
  67. 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] [PubMed]
  68. Andreev, D.E.; Hirnet, J.; Terenin, I.M.; Dmitriev, S.E.; Niepmann, M.; Shatsky, I.N. Glycyl-tRNA synthetase specifically binds to the poliovirus IRES to activate translation initiation. Nucleic Acids Res. 2012, 40, 5602–5614. [Google Scholar] [CrossRef]
  69. 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]
  70. Isoyama, T.; Kamoshita, N.; Yasui, K.; Iwai, A.; Shiroki, K.; Toyoda, H.; Yamada, A.; Takasaki, Y.; Nomoto, A. Lower concentration of La protein required for internal ribosome entry on hepatitis C virus RNA than on poliovirus RNA. J. Gen. Virol. 1999, 80 Pt 9, 2319–2327. [Google Scholar] [CrossRef]
  71. Wang, L.; Jeng, K.S.; Lai, M.M. Poly(C)-binding protein 2 interacts with sequences required for viral replication in the hepatitis C virus (HCV) 5′ untranslated region and directs HCV RNA replication through circularizing the viral genome. J. Virol. 2011, 85, 7954–7964. [Google Scholar] [CrossRef] [PubMed]
  72. Kumar, A.; Ray, U.; Das, S. Human La protein interaction with GCAC near the initiator AUG enhances hepatitis C Virus RNA replication by promoting linkage between 5′ and 3′ untranslated regions. J. Virol. 2013, 87, 6713–6726. [Google Scholar] [CrossRef] [PubMed]
  73. 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] [CrossRef] [PubMed]
  74. Kim, J.H.; Park, S.M.; Park, J.H.; Keum, S.J.; Jang, S.K. eIF2A mediates translation of hepatitis C viral mRNA under stress conditions. EMBO J. 2011, 30, 2454–2464. [Google Scholar] [CrossRef] [PubMed]
  75. Skabkin, M.A.; Skabkina, O.V.; Dhote, V.; Komar, A.A.; Hellen, C.U.; Pestova, T.V. Activities of Ligatin and MCT-1/DENR in eukaryotic translation initiation and ribosomal recycling. Genes Dev. 2010, 24, 1787–1801. [Google Scholar] [CrossRef]
  76. de Breyne, S.; Yu, Y.; Pestova, T.V.; Hellen, C.U. Factor requirements for translation initiation on the Simian picornavirus internal ribosomal entry site. RNA 2008, 14, 367–380. [Google Scholar] [CrossRef] [PubMed]
  77. Khawaja, A.; Vopalensky, V.; Pospisek, M. Understanding the potential of hepatitis C virus internal ribosome entry site domains to modulate translation initiation via their structure and function. Wiley Interdiscip. Rev. RNA 2015, 6, 211–224. [Google Scholar] [CrossRef]
  78. Filbin, M.E.; Vollmar, B.S.; Shi, D.; Gonen, T.; Kieft, J.S. HCV IRES manipulates the ribosome to promote the switch from translation initiation to elongation. Nat. Struct. Mol. Biol. 2013, 20, 150–158. [Google Scholar] [CrossRef]
  79. Spahn, C.M.; Kieft, J.S.; Grassucci, R.A.; Penczek, P.A.; Zhou, K.; Doudna, J.A.; Frank, J. Hepatitis C virus IRES RNA-induced changes in the conformation of the 40s ribosomal subunit. Science 2001, 291, 1959–1962. [Google Scholar] [CrossRef]
  80. Kieft, J.S.; Zhou, K.; Jubin, R.; Doudna, J.A. Mechanism of ribosome recruitment by hepatitis C IRES RNA. RNA 2001, 7, 194–206. [Google Scholar] [CrossRef]
  81. 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] [PubMed]
  82. 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] [PubMed]
  83. Filbin, M.E.; Kieft, J.S. HCV IRES domain IIb affects the configuration of coding RNA in the 40S subunit’s decoding groove. RNA 2011, 17, 1258–1273. [Google Scholar] [CrossRef] [PubMed]
  84. 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]
  85. Walker, M.J.; Shortridge, M.D.; Albin, D.D.; Cominsky, L.Y.; Varani, G. Structure of the RNA specialized translation initiation element that recruits eIF3 to the 5′-UTR of c-Jun. J. Mol. Biol. 2020, 432, 1841–1855. [Google Scholar] [CrossRef]
  86. Yamamoto, H.; Unbehaun, A.; Loerke, J.; Behrmann, E.; Collier, M.; Burger, J.; Mielke, T.; Spahn, C.M. Structure of the mammalian 80S initiation complex with initiation factor 5B on HCV-IRES RNA. Nat. Struct. Mol. Biol. 2014, 21, 721–727. [Google Scholar] [CrossRef] [PubMed]
  87. Quade, N.; Boehringer, D.; Leibundgut, M.; van den Heuvel, J.; Ban, N. Cryo-EM structure of Hepatitis C virus IRES bound to the human ribosome at 3.9-A resolution. Nat. Commun. 2015, 6, 7646. [Google Scholar] [CrossRef] [PubMed]
  88. 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]
  89. Wilson, J.E.; Powell, M.J.; Hoover, S.E.; Sarnow, P. Naturally occurring dicistronic cricket paralysis virus RNA is regulated by two internal ribosome entry sites. Mol. Cell. Biol. 2000, 20, 4990–4999. [Google Scholar] [CrossRef]
  90. Khong, A.; Bonderoff, J.M.; Spriggs, R.V.; Tammpere, E.; Kerr, C.H.; Jackson, T.J.; Willis, A.E.; Jan, E. Temporal regulation of distinct internal ribosome entry sites of the Dicistroviridae cricket paralysis virus. Viruses 2016, 8, 25. [Google Scholar] [CrossRef]
  91. Jan, E.; Sarnow, P. Factorless ribosome assembly on the internal ribosome entry site of cricket paralysis virus. J. Mol. Biol. 2002, 324, 889–902. [Google Scholar] [CrossRef] [PubMed]
  92. Kerr, C.H.; Jan, E. Commandeering the ribosome: Lessons learned from dicistroviruses about translation. J. Virol. 2016, 90, 5538–5540. [Google Scholar] [CrossRef] [PubMed]
  93. Hertz, M.I.; Thompson, S.R. Mechanism of translation initiation by Dicistroviridae IGR IRESs. Virology 2011, 411, 355–361. [Google Scholar] [CrossRef] [PubMed]
  94. Colussi, T.M.; Costantino, D.A.; Zhu, J.; Donohue, J.P.; Korostelev, A.A.; Jaafar, Z.A.; Plank, T.D.; Noller, H.F.; Kieft, J.S. Initiation of translation in bacteria by a structured eukaryotic IRES RNA. Nature 2015, 519, 110–113. [Google Scholar] [CrossRef] [PubMed]
  95. Pfingsten, J.S.; Costantino, D.A.; Kieft, J.S. Structural basis for ribosome recruitment and manipulation by a viral IRES RNA. Science 2006, 314, 1450–1454. [Google Scholar] [CrossRef] [PubMed]
  96. 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]
  97. Spahn, C.M.; Jan, E.; Mulder, A.; Grassucci, R.A.; Sarnow, P.; Frank, J. Cryo-EM visualization of a viral internal ribosome entry site bound to human ribosomes: The IRES functions as an RNA-based translation factor. Cell 2004, 118, 465–475. [Google Scholar] [CrossRef]
  98. Petrov, A.; Grosely, R.; Chen, J.; O’Leary, S.E.; Puglisi, J.D. Multiple parallel pathways of translation initiation on the CrPV IRES. Mol. Cell 2016, 62, 92–103. [Google Scholar] [CrossRef]
  99. 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]
  100. 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]
  101. Kanamori, Y.; Nakashima, N. A tertiary structure model of the internal ribosome entry site (IRES) for methionine-independent initiation of translation. RNA 2001, 7, 266–274. [Google Scholar] [CrossRef]
  102. Nishiyama, T.; Yamamoto, H.; Shibuya, N.; Hatakeyama, Y.; Hachimori, A.; Uchiumi, T.; Nakashima, N. Structural elements in the internal ribosome entry site of Plautia stali intestine virus responsible for binding with ribosomes. Nucleic Acids Res. 2003, 31, 2434–2442. [Google Scholar] [CrossRef] [PubMed]
  103. 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] [PubMed]
  104. Ruehle, M.D.; Zhang, H.; Sheridan, R.M.; Mitra, S.; Chen, Y.; Gonzalez, R.L.; Cooperman, B.S.; Kieft, J.S. A dynamic RNA loop in an IRES affects multiple steps of elongation factor-mediated translation initiation. eLife 2015, 4, e08146. [Google Scholar] [CrossRef] [PubMed]
  105. Butcher, S.E.; Jan, E. tRNA-mimicry in IRES-mediated translation and recoding. RNA Biol. 2016, 13, 1068–1074. [Google Scholar] [CrossRef] [PubMed]
  106. Fernandez, J.; Yaman, I.; Sarnow, P.; Snider, M.D.; Hatzoglou, M. Regulation of internal ribosomal entry site-mediated translation by phosphorylation of the translation initiation factor eIF2alpha. J. Biol. Chem. 2002, 277, 19198–19205. [Google Scholar] [CrossRef] [PubMed]
  107. 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] [CrossRef] [PubMed]
  108. Avanzino, B.C.; Fuchs, G.; Fraser, C.S. Cellular cap-binding protein, eIF4E, promotes picornavirus genome restructuring and translation. Proc. Natl. Acad. Sci. USA 2017, 114, 9611–9616. [Google Scholar] [CrossRef]
  109. Sweeney, T.R.; Dhote, V.; Yu, Y.; Hellen, C.U. A distinct class of internal ribosomal entry site in members of the Kobuvirus and proposed Salivirus and Paraturdivirus genera of the Picornaviridae. J. Virol. 2012, 86, 1468–1486. [Google Scholar] [CrossRef]
  110. 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]
  111. Plank, T.D.; Whitehurst, J.T.; Kieft, J.S. Cell type specificity and structural determinants of IRES activity from the 5′ leaders of different HIV-1 transcripts. Nucleic Acids Res. 2013, 41, 6698–6714. [Google Scholar] [CrossRef] [PubMed]
  112. Ohlmann, T.; Mengardi, C.; Lopez-Lastra, M. Translation initiation of the HIV-1 mRNA. Translation (Austin) 2014, 2, e960242. [Google Scholar]
  113. Singh, G.; Seufzer, B.; Song, Z.; Zucko, D.; Heng, X.; Boris-Lawrie, K. HIV-1 hypermethylated guanosine cap licenses specialized translation unaffected by mTOR. Proc. Natl. Acad. Sci. USA 2022, 119, e2105153118. [Google Scholar] [CrossRef] [PubMed]
  114. Lu, J.; Zhang, J.; Wang, X.; Jiang, H.; Liu, C.; Hu, Y. In vitro and in vivo identification of structural and sequence elements in the 5′ untranslated region of Ectropis obliqua picorna-like virus required for internal initiation. J. Gen. Virol. 2006, 87, 3667–3677. [Google Scholar] [CrossRef]
  115. 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] [CrossRef] [PubMed]
  116. Geng, G.; Wang, D.; Liu, Z.; Wang, Y.; Zhu, M.; Cao, X.; Yu, C.; Yuan, X. Translation of plant RNA viruses. Viruses 2021, 13, 2499. [Google Scholar] [CrossRef]
  117. Lee, A.S.; Burdeinick-Kerr, R.; Whelan, S.P. A ribosome-specialized translation initiation pathway is required for cap-dependent translation of vesicular stomatitis virus mRNAs. Proc. Natl. Acad. Sci. USA 2013, 110, 324–329. [Google Scholar] [CrossRef]
  118. Lee, A.S.; Kranzusch, P.J.; Doudna, J.A.; Cate, J.H. eIF3d is an mRNA cap-binding protein that is required for specialized translation initiation. Nature 2016, 536, 96–99. [Google Scholar] [CrossRef]
  119. Liang, M.; Hody, C.; Yammine, V.; Soin, R.; Sun, Y.; Lin, X.; Tian, X.; Meurs, R.; Perdrau, C.; Delacourt, N.; et al. eIF4EHP promotes Ldh mRNA translation in and fruit fly adaptation to hypoxia. EMBO Rep. 2023, 24, e56460. [Google Scholar] [CrossRef]
  120. Mueller, P.P.; Hinnebusch, A.G. Multiple upstream AUG codons mediate translational control of GCN4. Cell 1986, 45, 201–207. [Google Scholar] [CrossRef]
  121. Mueller, P.P.; Jackson, B.M.; Miller, P.F.; Hinnebusch, A.G. The first and fourth upstream open reading frames in GCN4 mRNA have similar initiation efficiencies but respond differently in translational control to change in length and sequence. Mol. Cell. Biol. 1988, 8, 5439–5447. [Google Scholar] [PubMed]
  122. Miller, P.F.; Hinnebusch, A.G. Sequences that surround the stop codons of upstream open reading frames in GCN4 mRNA determine their distinct functions in translational control. Genes Dev. 1989, 3, 1217–1225. [Google Scholar] [CrossRef] [PubMed]
  123. Mueller, P.P.; Harashima, S.; Hinnebusch, A.G. A segment of GCN4 mRNA containing the upstream AUG codons confers translational control upon a heterologous yeast transcript. Proc. Natl. Acad. Sci. USA 1987, 84, 2863–2867. [Google Scholar] [CrossRef] [PubMed]
  124. Dever, T.E.; Ivanov, I.P.; Hinnebusch, A.G. Translational regulation by uORFs and start codon selection stringency. Genes Dev. 2023, 37, 474–489. [Google Scholar] [CrossRef]
  125. Ingolia, N.T.; Lareau, L.F.; Weissman, J.S. Ribosome profiling of mouse embryonic stem cells reveals the complexity and dynamics of mammalian proteomes. Cell 2011, 147, 789–802. [Google Scholar] [CrossRef] [PubMed]
  126. Lee, S.; Liu, B.; Lee, S.; Huang, S.X.; Shen, B.; Qian, S.B. Global mapping of translation initiation sites in mammalian cells at single-nucleotide resolution. Proc. Natl. Acad. Sci. USA 2012, 109, E2424–E2432. [Google Scholar] [CrossRef] [PubMed]
  127. Morris, D.R.; Geballe, A.P. Upstream open reading frames as regulators of mRNA translation. Mol. Cell. Biol. 2000, 20, 8635–8642. [Google Scholar] [CrossRef] [PubMed]
  128. Uchiyama-Kadokura, N.; Murakami, K.; Takemoto, M.; Koyanagi, N.; Murota, K.; Naito, S.; Onouchi, H. Polyamine-responsive ribosomal arrest at the stop codon of an upstream open reading frame of the AdoMetDC1 gene triggers nonsense-mediated mRNA decay in Arabidopsis thaliana. Plant Cell Physiol. 2014, 55, 1556–1567. [Google Scholar] [CrossRef]
  129. Dever, T.E.; Ivanov, I.P.; Sachs, M.S. Conserved upstream open reading frame nascent peptides that control translation. Annu. Rev. Genet. 2020, 54, 237–264. [Google Scholar] [CrossRef]
  130. Calvo, S.E.; Pagliarini, D.J.; Mootha, V.K. Upstream open reading frames cause widespread reduction of protein expression and are polymorphic among humans. Proc. Natl. Acad. Sci. USA 2009, 106, 7507–7512. [Google Scholar] [CrossRef]
  131. McGillivray, P.; Ault, R.; Pawashe, M.; Kitchen, R.; Balasubramanian, S.; Gerstein, M. A comprehensive catalog of predicted functional upstream open reading frames in humans. Nucleic Acids Res. 2018, 46, 3326–3338. [Google Scholar] [CrossRef] [PubMed]
  132. Manske, F.; Ogoniak, L.; Jurgens, L.; Grundmann, N.; Makalowski, W.; Wethmar, K. The new uORFdb: Integrating literature, sequence, and variation data in a central hub for uORF research. Nucleic Acids Res. 2023, 51, D328–D336. [Google Scholar] [CrossRef]
  133. Barbosa, C.; Peixeiro, I.; Romao, L. Gene expression regulation by upstream open reading frames and human disease. PLoS Genet. 2013, 9, e1003529. [Google Scholar] [CrossRef] [PubMed]
  134. Silva, J.; Fernandes, R.; Romao, L. Translational regulation by upstream open reading frames and human diseases. Adv. Exp. Med. Biol. 2019, 1157, 99–116. [Google Scholar] [PubMed]
  135. Zhang, H.; Wang, Y.; Lu, J. Function and evolution of upstream ORFs in eukaryotes. Trends Biochem. Sci. 2019, 44, 782–794. [Google Scholar] [CrossRef] [PubMed]
  136. Ruiz-Orera, J.; Alba, M.M. Translation of small open reading frames: Roles in regulation and evolutionary innovation. Trends Genet. 2019, 35, 186–198. [Google Scholar] [CrossRef] [PubMed]
  137. Vattem, K.M.; Wek, R.C. Reinitiation involving upstream ORFs regulates ATF4 mRNA translation in mammalian cells. Proc. Natl. Acad. Sci. USA 2004, 101, 11269–11274. [Google Scholar] [CrossRef] [PubMed]
  138. Andreev, D.E.; Arnold, M.; Kiniry, S.J.; Loughran, G.; Michel, A.M.; Rachinskii, D.; Baranov, P.V. TASEP modelling provides a parsimonious explanation for the ability of a single uORF to derepress translation during the integrated stress response. eLife 2018, 7, e32563. [Google Scholar] [CrossRef]
  139. Vasu, K.; Ramachandiran, I.; Chechi, A.; Khan, K.; Khan, D.; Kaufman, R.; Fox, P.L. Translational control of murine adiponectin expression by an upstream open reading frame element. RNA Biol. 2023, 20, 737–749. [Google Scholar] [CrossRef]
  140. Johnstone, T.G.; Bazzini, A.A.; Giraldez, A.J. Upstream ORFs are prevalent translational repressors in vertebrates. EMBO J. 2016, 35, 706–723. [Google Scholar] [CrossRef]
  141. Torrance, V.; Lydall, D. Overlapping open reading frames strongly reduce human and yeast STN1 gene expression and affect telomere function. PLoS Genet. 2018, 14, e1007523. [Google Scholar] [CrossRef] [PubMed]
  142. Wright, B.W.; Yi, Z.; Weissman, J.S.; Chen, J. The dark proteome: Translation from noncanonical open reading frames. Trends Cell Biol. 2022, 32, 243–258. [Google Scholar] [CrossRef]
  143. Vasu, K.; Khan, D.; Ramachandiran, I.; Blankenberg, D.; Fox, P.L. Analysis of nested alternate open reading frames and their encoded proteins. NAR Genom. Bioinform. 2022, 4, lqac076. [Google Scholar] [CrossRef] [PubMed]
  144. Vanderperre, B.; Lucier, J.F.; Roucou, X. HAltORF: A database of predicted out-of-frame alternative open reading frames in human. Database 2012, 2012, bas025. [Google Scholar] [CrossRef] [PubMed]
  145. Gunisova, S.; Hronova, V.; Mohammad, M.P.; Hinnebusch, A.G.; Valasek, L.S. Please do not recycle! Translation reinitiation in microbes and higher eukaryotes. FEMS Microbiol. Rev. 2018, 42, 165–192. [Google Scholar] [CrossRef] [PubMed]
  146. Sherlock, M.E.; Baquero Galvis, L.; Vicens, Q.; Kieft, J.S.; Jagannathan, S. Principles, mechanisms, and biological implications of translation termination-reinitiation. RNA 2023, 29, 865–884. [Google Scholar] [CrossRef] [PubMed]
  147. Gu, Y.; Mao, Y.; Jia, L.; Dong, L.; Qian, S.B. Bi-directional ribosome scanning controls the stringency of start codon selection. Nat. Commun. 2021, 12, 6604. [Google Scholar] [CrossRef] [PubMed]
  148. Matsuda, D.; Dreher, T.W. Close spacing of AUG initiation codons confers dicistronic character on a eukaryotic mRNA. RNA 2006, 12, 1338–1349. [Google Scholar] [CrossRef]
  149. Chirico, N.; Vianelli, A.; Belshaw, R. Why genes overlap in viruses. Proc. Biol. Sci. 2010, 277, 3809–3817. [Google Scholar] [CrossRef]
  150. Ho, J.S.Y.; Zhu, Z.; Marazzi, I. Unconventional viral gene expression mechanisms as therapeutic targets. Nature 2021, 593, 362–371. [Google Scholar] [CrossRef]
  151. Finkel, Y.; Stern-Ginossar, N.; Schwartz, M. Viral short ORFs and their possible functions. Proteomics 2018, 18, e1700255. [Google Scholar] [CrossRef] [PubMed]
  152. Hwang, W.L.; Su, T.S. Translational regulation of hepatitis B virus polymerase gene by termination-reinitiation of an upstream minicistron in a length-dependent manner. J. Gen. Virol. 1998, 79 Pt 9, 2181–2189. [Google Scholar] [CrossRef] [PubMed]
  153. Chen, A.; Kao, Y.F.; Brown, C.M. Translation of the first upstream ORF in the hepatitis B virus pregenomic RNA modulates translation at the core and polymerase initiation codons. Nucleic Acids Res. 2005, 33, 1169–1181. [Google Scholar] [CrossRef] [PubMed]
  154. Zong, L.; Qin, Y.; Jia, H.; Ye, L.; Wang, Y.; Zhang, J.; Wands, J.R.; Tong, S.; Li, J. Differential regulation of hepatitis B virus core protein expression and genome replication by a small upstream open reading frame and naturally occurring mutations in the precore region. Virology 2017, 505, 155–161. [Google Scholar] [CrossRef]
  155. Degnin, C.R.; Schleiss, M.R.; Cao, J.; Geballe, A.P. Translational inhibition mediated by a short upstream open reading frame in the human cytomegalovirus gpUL4 (gp48) transcript. J. Virol. 1993, 67, 5514–5521. [Google Scholar] [CrossRef] [PubMed]
  156. Kronstad, L.M.; Brulois, K.F.; Jung, J.U.; Glaunsinger, B.A. Dual short upstream open reading frames control translation of a herpesviral polycistronic mRNA. PLoS Pathog. 2013, 9, e1003156. [Google Scholar] [CrossRef] [PubMed]
  157. Kronstad, L.M.; Brulois, K.F.; Jung, J.U.; Glaunsinger, B.A. Reinitiation after translation of two upstream open reading frames (ORF) governs expression of the ORF35-37 Kaposi’s sarcoma-associated herpesvirus polycistronic mRNA. J. Virol. 2014, 88, 6512–6518. [Google Scholar] [CrossRef]
  158. Shabman, R.S.; Hoenen, T.; Groseth, A.; Jabado, O.; Binning, J.M.; Amarasinghe, G.K.; Feldmann, H.; Basler, C.F. An upstream open reading frame modulates ebola virus polymerase translation and virus replication. PLoS Pathog. 2013, 9, e1003147. [Google Scholar] [CrossRef]
  159. Luukkonen, B.G.; Tan, W.; Schwartz, S. Efficiency of reinitiation of translation on human immunodeficiency virus type 1 mRNAs is determined by the length of the upstream open reading frame and by intercistronic distance. J. Virol. 1995, 69, 4086–4094. [Google Scholar] [CrossRef]
  160. Krummheuer, J.; Johnson, A.T.; Hauber, I.; Kammler, S.; Anderson, J.L.; Hauber, J.; Purcell, D.F.; Schaal, H. A minimal uORF within the HIV-1 vpu leader allows efficient translation initiation at the downstream env AUG. Virology 2007, 363, 261–271. [Google Scholar] [CrossRef]
  161. Wu, H.Y.; Guan, B.J.; Su, Y.P.; Fan, Y.H.; Brian, D.A. Reselection of a genomic upstream open reading frame in mouse hepatitis coronavirus 5′-untranslated-region mutants. J. Virol. 2014, 88, 846–858. [Google Scholar] [CrossRef] [PubMed]
  162. Hofmann, M.A.; Senanayake, S.D.; Brian, D.A. A translation-attenuating intraleader open reading frame is selected on coronavirus mRNAs during persistent infection. Proc. Natl. Acad. Sci. USA 1993, 90, 11733–11737. [Google Scholar] [CrossRef] [PubMed]
  163. van der Velden, G.J.; Klaver, B.; Das, A.T.; Berkhout, B. Upstream AUG codons in the simian immunodeficiency virus SIVmac239 genome regulate Rev and Env protein translation. J. Virol. 2012, 86, 12362–12371. [Google Scholar] [CrossRef] [PubMed]
  164. van der Velden, G.J.; Vink, M.A.; Klaver, B.; Das, A.T.; Berkhout, B. An AUG codon upstream of rev and env open reading frames ensures optimal translation of the simian immunodeficiency virus Env protein. Virology 2013, 436, 191–200. [Google Scholar] [CrossRef] [PubMed]
  165. Schepetilnikov, M.; Schott, G.; Katsarou, K.; Thiebeauld, O.; Keller, M.; Ryabova, L.A. Molecular dissection of the prototype foamy virus (PFV) RNA 5′-UTR identifies essential elements of a ribosomal shunt. Nucleic Acids Res. 2009, 37, 5838–5847. [Google Scholar] [CrossRef] [PubMed]
  166. Pooggin, M.M.; Rajeswaran, R.; Schepetilnikov, M.V.; Ryabova, L.A. Short ORF-dependent ribosome shunting operates in an RNA picorna-like virus and a DNA pararetrovirus that cause rice tungro disease. PLoS Pathog. 2012, 8, e1002568. [Google Scholar] [CrossRef] [PubMed]
  167. Luttermann, C.; Meyers, G. A bipartite sequence motif induces translation reinitiation in feline calicivirus RNA. J. Biol. Chem. 2007, 282, 7056–7065. [Google Scholar] [CrossRef] [PubMed]
  168. Meyers, G. Characterization of the sequence element directing translation reinitiation in RNA of the calicivirus rabbit hemorrhagic disease virus. J. Virol. 2007, 81, 9623–9632. [Google Scholar] [CrossRef]
  169. Zinoviev, A.; Hellen, C.U.T.; Pestova, T.V. Multiple mechanisms of reinitiation on bicistronic calicivirus mRNAs. Mol. Cell 2015, 57, 1059–1073. [Google Scholar] [CrossRef]
  170. Putlyaeva, L.V.; Schwartz, A.M.; Korneev, K.V.; Covic, M.; Uroshlev, L.A.; Makeev, V.Y.; Dmitriev, S.E.; Kuprash, D.V. Upstream open reading frames regulate translation of the long isoform of SLAMF1 mRNA that encodes costimulatory receptor CD150. Biochemistry (Moscow) 2014, 79, 1405–1411. [Google Scholar] [CrossRef]
  171. Sampath, P.; Mazumder, B.; Seshadri, V.; Fox, P.L. Transcript-selective translational silencing by gamma interferon is directed by a novel structural element in the ceruloplasmin mRNA 3′ untranslated region. Mol. Cell. Biol. 2003, 23, 1509–1519. [Google Scholar] [CrossRef] [PubMed]
  172. Mukhopadhyay, R.; Ray, P.S.; Arif, A.; Brady, A.K.; Kinter, M.; Fox, P.L. DAPK-ZIPK-L13a axis constitutes a negative-feedback module regulating inflammatory gene expression. Mol. Cell 2008, 32, 371–382. [Google Scholar] [CrossRef] [PubMed]
  173. Ray, P.S.; Fox, P.L. A post-transcriptional pathway represses monocyte VEGF-A expression and angiogenic activity. EMBO J. 2007, 26, 3360–3372. [Google Scholar] [CrossRef] [PubMed]
  174. Vyas, K.; Chaudhuri, S.; Leaman, D.W.; Komar, A.A.; Musiyenko, A.; Barik, S.; Mazumder, B. Genome-wide polysome profiling reveals an inflammation-responsive post-transcriptional operon in IFN-gamma-activated monocytes. Mol. Cell. Biol. 2009, 29, 458–470. [Google Scholar] [CrossRef] [PubMed]
  175. Sampath, P.; Mazumder, B.; Seshadri, V.; Gerber, C.A.; Chavatte, L.; Kinter, M.; Ting, S.M.; Dignam, J.D.; Kim, S.; Driscoll, D.M.; et al. Noncanonical function of glutamyl-prolyl-tRNA synthetase: Gene-specific silencing of translation. Cell 2004, 119, 195–208. [Google Scholar] [CrossRef] [PubMed]
  176. Mazumder, B.; Sampath, P.; Seshadri, V.; Maitra, R.K.; DiCorleto, P.; Fox, P.L. Regulated release of L13a from the 60S ribosomal subunit as a mechanism of transcript-specific translational control. Cell 2003, 115, 187–198. [Google Scholar] [CrossRef] [PubMed]
  177. Wells, S.E.; Hillner, P.E.; Vale, R.D.; Sachs, A.B. Circularization of mRNA by eukaryotic translation initiation factors. Mol. Cell 1998, 2, 135–140. [Google Scholar] [CrossRef]
  178. Vicens, Q.; Kieft, J.S.; Rissland, O.S. Revisiting the closed-loop model and the nature of mRNA 5′-3′ communication. Mol. Cell 2018, 72, 805–812. [Google Scholar] [CrossRef]
  179. Filbin, M.E.; Kieft, J.S. Linking Alpha to Omega: Diverse and dynamic RNA-based mechanisms to regulate gene expression by 5′-to-3′ communication. F1000Research 2016, 5. [Google Scholar] [CrossRef]
  180. Kapasi, P.; Chaudhuri, S.; Vyas, K.; Baus, D.; Komar, A.A.; Fox, P.L.; Merrick, W.C.; Mazumder, B. L13a blocks 48S assembly: Role of a general initiation factor in mRNA-specific translational control. Mol. Cell 2007, 25, 113–126. [Google Scholar] [CrossRef]
  181. Keene, J.D. RNA regulons: Coordination of post-transcriptional events. Nat. Rev. Genet. 2007, 8, 533–543. [Google Scholar] [CrossRef] [PubMed]
  182. Mukhopadhyay, R.; Jia, J.; Arif, A.; Ray, P.S.; Fox, P.L. The GAIT system: A gatekeeper of inflammatory gene expression. Trends Biochem. Sci. 2009, 34, 324–331. [Google Scholar] [CrossRef] [PubMed]
  183. Marquez-Jurado, S.; Nogales, A.; Zuniga, S.; Enjuanes, L.; Almazan, F. Identification of a gamma interferon-activated inhibitor of translation-like RNA motif at the 3′ end of the transmissible gastroenteritis coronavirus genome modulating innate immune response. mBio 2015, 6, e00105. [Google Scholar] [CrossRef] [PubMed]
  184. Mazumder, B.; Poddar, D.; Basu, A.; Kour, R.; Verbovetskaya, V.; Barik, S. Extraribosomal L13a is a specific innate immune factor for antiviral defense. J. Virol. 2014, 88, 9100–9110. [Google Scholar] [CrossRef] [PubMed]
  185. Basu, A.; Penumutchu, S.; Nguyen, K.; Mbonye, U.; Tolbert, B.S.; Karn, J.; Komar, A.A.; Mazumder, B. A structurally conserved RNA element within SARS-CoV-2 ORF1a RNA and S mRNA regulates translation in response to viral S protein-induced signaling in human lung cells. J. Virol. 2022, 96, e0167821. [Google Scholar] [CrossRef] [PubMed]
  186. Khan, D.; Terenzi, F.; Liu, G.; Ghosh, P.K.; Ye, F.; Nguyen, K.; China, A.; Ramachandiran, I.; Chakraborty, S.; Stefan, J.; et al. A viral pan-end RNA element and host complex define a SARS-CoV-2 regulon. Nat. Commun. 2023, 14, 3385. [Google Scholar] [CrossRef] [PubMed]
  187. Arif, A.; Yao, P.; Terenzi, F.; Jia, J.; Ray, P.S.; Fox, P.L. The GAIT translational control system. Wiley Interdiscip. Rev. RNA 2017, 9, e1441. [Google Scholar] [CrossRef] [PubMed]
  188. Mazumder, B.; Seshadri, V.; Imataka, H.; Sonenberg, N.; Fox, P.L. Translational silencing of ceruloplasmin requires the essential elements of mRNA circularization: Poly(A) tail, Poly(A)-binding protein, and eukaryotic translation initiation factor 4G. Mol. Cell. Biol. 2001, 21, 6440–6449. [Google Scholar] [CrossRef]
  189. Mazumder, B.; Seshadri, V.; Fox, P.L. Translational control by the 3′-UTR: The ends specify the means. Trends Biochem. Sci. 2003, 28, 91–98. [Google Scholar] [CrossRef]
  190. Nicholson, B.L.; White, K.A. Functional long-range RNA-RNA interactions in positive-strand RNA viruses. Nat. Rev. Microbiol. 2014, 12, 493–504. [Google Scholar] [CrossRef]
  191. 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]
  192. Simon, A.E.; Miller, W.A. 3′ cap-independent translation enhancers of plant viruses. Annu. Rev. Microbiol. 2013, 67, 21–42. [Google Scholar] [CrossRef] [PubMed]
  193. Miras, M.; Miller, W.A.; Truniger, V.; Aranda, M.A. Non-canonical translation in plant RNA viruses. Front. Plant Sci. 2017, 8, 494. [Google Scholar] [CrossRef] [PubMed]
  194. 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] [PubMed]
  195. Bernardes, W.S.; Menossi, M. Plant 3′ regulatory regions from mRNA-encoding genes and their uses to modulate expression. Front. Plant Sci. 2020, 11, 1252. [Google Scholar] [CrossRef] [PubMed]
  196. Banerjee, A.K.; Blanco, M.R.; Bruce, E.A.; Honson, D.D.; Chen, L.M.; Chow, A.; Bhat, P.; Ollikainen, N.; Quinodoz, S.A.; Loney, C.; et al. SARS-CoV-2 disrupts splicing, translation, and protein trafficking to suppress host defenses. Cell 2020, 183, 1325–1339. [Google Scholar] [CrossRef] [PubMed]
  197. Khan, D.; Fox, P.L. Aminoacyl-tRNA synthetase interactions in SARS-CoV-2 infection. Biochem. Soc. Trans. 2023, 51, 2127–2141. [Google Scholar] [CrossRef]
  198. Slobodin, B.; Sehrawat, U.; Lev, A.; Hayat, D.; Zuckerman, B.; Fraticelli, D.; Ogran, A.; Ben-Shmuel, A.; Bar-David, E.; Levy, H.; et al. Cap-independent translation and a precisely located RNA sequence enable SARS-CoV-2 to control host translation and escape anti-viral response. Nucleic Acids Res. 2022, 50, 8080–8092. [Google Scholar] [CrossRef]
  199. Guo, L.; Allen, E.M.; Miller, W.A. Base-pairing between untranslated regions facilitates translation of uncapped, nonpolyadenylated viral RNA. Mol. Cell 2001, 7, 1103–1109. [Google Scholar] [CrossRef]
  200. 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]
  201. 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] [PubMed]
  202. Sharma, S.D.; Kraft, J.J.; Miller, W.A.; Goss, D.J. Recruitment of the 40S ribosome subunit to the 3′-untranslated region (UTR) of a viral mRNA, via the eIF4 complex, facilitates cap-independent translation. J. Biol. Chem. 2015, 290, 11268–11281. [Google Scholar] [CrossRef] [PubMed]
  203. Kraft, J.J.; Treder, K.; Peterson, M.S.; Miller, W.A. Cation-dependent folding of 3′ cap-independent translation elements facilitates interaction of a 17-nucleotide conserved sequence with eIF4G. Nucleic Acids Res. 2013, 41, 3398–3413. [Google Scholar] [CrossRef] [PubMed]
  204. 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] [PubMed]
  205. 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] [PubMed]
  206. 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] [PubMed]
  207. 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]
  208. 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]
  209. 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]
  210. 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]
  211. 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] [CrossRef] [PubMed]
  212. 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] [PubMed]
  213. Yuan, X.; Shi, K.; Meskauskas, A.; Simon, A.E. The 3′ end of Turnip crinkle virus contains a highly interactive structure including a translational enhancer that is disrupted by binding to the RNA-dependent RNA polymerase. RNA 2009, 15, 1849–1864. [Google Scholar] [CrossRef] [PubMed]
  214. Le, M.T.; Kasprzak, W.K.; Kim, T.; Gao, F.; Young, M.Y.; Yuan, X.; Shapiro, B.A.; Seog, J.; Simon, A.E. Folding behavior of a T-shaped, ribosome-binding translation enhancer implicated in a wide-spread conformational switch. eLife 2017, 6, e22883. [Google Scholar] [CrossRef]
  215. Dreher, T.W. Viral tRNAs and tRNA-like structures. Wiley Interdiscip. Rev. RNA 2010, 1, 402–414. [Google Scholar] [CrossRef] [PubMed]
  216. Yot, P.; Pinck, M.; Haenni, A.L.; Duranton, H.M.; Chapeville, F. Valine-specific tRNA-like structure in turnip yellow mosaic virus RNA. Proc. Natl. Acad. Sci. USA 1970, 67, 1345–1352. [Google Scholar] [CrossRef] [PubMed]
  217. Sherlock, M.E.; Hartwick, E.W.; MacFadden, A.; Kieft, J.S. Structural diversity and phylogenetic distribution of valyl tRNA-like structures in viruses. RNA 2021, 27, 27–39. [Google Scholar] [CrossRef]
  218. Dreher, T.W.; Uhlenbeck, O.C.; Browning, K.S. Quantitative assessment of EF-1alpha.GTP binding to aminoacyl-tRNAs, aminoacyl-viral RNA, and tRNA shows close correspondence to the RNA binding properties of EF-Tu. J. Biol. Chem. 1999, 274, 666–672. [Google Scholar] [CrossRef]
  219. Colussi, T.M.; Costantino, D.A.; Hammond, J.A.; Ruehle, G.M.; Nix, J.C.; Kieft, J.S. The structural basis of transfer RNA mimicry and conformational plasticity by a viral RNA. Nature 2014, 511, 366–369. [Google Scholar] [CrossRef]
  220. Hammond, J.A.; Rambo, R.P.; Kieft, J.S. Multi-domain packing in the aminoacylatable 3′ end of a plant viral RNA. J. Mol. Biol. 2010, 399, 450–463. [Google Scholar] [CrossRef]
  221. Barends, S.; Bink, H.H.; van den Worm, S.H.; Pleij, C.W.; Kraal, B. Entrapping ribosomes for viral translation: tRNA mimicry as a molecular Trojan horse. Cell 2003, 112, 123–129. [Google Scholar] [CrossRef]
  222. Matsuda, D.; Dreher, T.W. Cap- and initiator tRNA-dependent initiation of TYMV polyprotein synthesis by ribosomes: Evaluation of the Trojan horse model for TYMV RNA translation. RNA 2007, 13, 129–137. [Google Scholar] [CrossRef]
  223. Bonilla, S.L.; Sherlock, M.E.; MacFadden, A.; Kieft, J.S. A viral RNA hijacks host machinery using dynamic conformational changes of a tRNA-like structure. Science 2021, 374, 955–960. [Google Scholar] [CrossRef]
  224. Sherlock, M.E.; Langeberg, C.J.; Kieft, J.S. Diversity and modularity of tyrosine-accepting tRNA-like structures. RNA 2024, 30, 213–222. [Google Scholar] [CrossRef]
Figure 1. Translation in viruses is regulated by themes shared with hosts. Various RNA elements shown here are IRES (Internal Ribosome Entry Site), uORF (upstream Open Reading Frame), oORF (overlapping Open Reading Frame), TURBS (Termination Upstream Ribosome Binding Site), VAIT (Virus Activated Inhibitor of Translation element), 3′-CITE (3′-Cap Independent Translation Enhancer), SPEAR (Sarbecoviral Pan-End Activating RNA element), and 3′TLS (3′ tRNA-Like Structure) that can be charged with amino acid (aa). IRESs and 3′-CITEs are diverse, as discussed in the text. 3′-CITEs and VAIT activate and inhibit virus translation in a 5′-end dependent manner, as shown with dashed lines.
Figure 1. Translation in viruses is regulated by themes shared with hosts. Various RNA elements shown here are IRES (Internal Ribosome Entry Site), uORF (upstream Open Reading Frame), oORF (overlapping Open Reading Frame), TURBS (Termination Upstream Ribosome Binding Site), VAIT (Virus Activated Inhibitor of Translation element), 3′-CITE (3′-Cap Independent Translation Enhancer), SPEAR (Sarbecoviral Pan-End Activating RNA element), and 3′TLS (3′ tRNA-Like Structure) that can be charged with amino acid (aa). IRESs and 3′-CITEs are diverse, as discussed in the text. 3′-CITEs and VAIT activate and inhibit virus translation in a 5′-end dependent manner, as shown with dashed lines.
Viruses 16 00468 g001
Figure 2. A unique regulon in viruses that employ discontinuous transcription can regulate genomic and sub-genomic RNAs. As an example, a simplified schematic of the SPEAR regulon of SARS-CoV-2 is shown. L: 5′-leader (within solid black arrows) that resides in the 5′UTR, gRNA: genomic RNA, sgRNA: subgenomic RNA. Frameshift element between ORF1a and ORF1b is shown in grey. Green, brown and blue dashed lines represent sgRNA-generating events between identical transcription regulatory sequences contained in 5′-leader and in nucleotides 5′ to ORFs encoded in sgRNA1 (green), sgRNA2 (blue), sgRNA3 (brown) and so on. The genomic terminus, common to gRNA and sgRNAs, reportedly pairs with the 5′-leader (darker gray dashed line). 5′-3′ communication through the SPEAR element is yet unknown (lighter gray dashed line).
Figure 2. A unique regulon in viruses that employ discontinuous transcription can regulate genomic and sub-genomic RNAs. As an example, a simplified schematic of the SPEAR regulon of SARS-CoV-2 is shown. L: 5′-leader (within solid black arrows) that resides in the 5′UTR, gRNA: genomic RNA, sgRNA: subgenomic RNA. Frameshift element between ORF1a and ORF1b is shown in grey. Green, brown and blue dashed lines represent sgRNA-generating events between identical transcription regulatory sequences contained in 5′-leader and in nucleotides 5′ to ORFs encoded in sgRNA1 (green), sgRNA2 (blue), sgRNA3 (brown) and so on. The genomic terminus, common to gRNA and sgRNAs, reportedly pairs with the 5′-leader (darker gray dashed line). 5′-3′ communication through the SPEAR element is yet unknown (lighter gray dashed line).
Viruses 16 00468 g002
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

Khan, D.; Fox, P.L. Host-like RNA Elements Regulate Virus Translation. Viruses 2024, 16, 468. https://doi.org/10.3390/v16030468

AMA Style

Khan D, Fox PL. Host-like RNA Elements Regulate Virus Translation. Viruses. 2024; 16(3):468. https://doi.org/10.3390/v16030468

Chicago/Turabian Style

Khan, Debjit, and Paul L. Fox. 2024. "Host-like RNA Elements Regulate Virus Translation" Viruses 16, no. 3: 468. https://doi.org/10.3390/v16030468

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