Next Article in Journal
Different Profiles of Cytokines, Chemokines and Coagulation Mediators Associated with Severity in Brazilian Patients Infected with Dengue Virus
Next Article in Special Issue
Interplay between Host tRNAs and HIV-1: A Structural Perspective
Previous Article in Journal
De-Escalating Strategies in HPV-Associated Head and Neck Squamous Cell Carcinoma
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

RNA Structures and Their Role in Selective Genome Packaging

1
Genome Architecture and Evolution of RNA Viruses, Helmholtz Institute for RNA-Based Infection Research, Helmholtz-Center for Infection Research, 97080 Würzburg, Germany
2
Faculty of Medicine, University of Würzburg, 97080 Würzburg, Germany
*
Author to whom correspondence should be addressed.
Viruses 2021, 13(9), 1788; https://doi.org/10.3390/v13091788
Submission received: 29 July 2021 / Revised: 31 August 2021 / Accepted: 2 September 2021 / Published: 8 September 2021
(This article belongs to the Special Issue RNA Packaging 2.0)

Abstract

:
To generate infectious viral particles, viruses must specifically select their genomic RNA from milieu that contains a complex mixture of cellular or non-genomic viral RNAs. In this review, we focus on the role of viral encoded RNA structures in genome packaging. We first discuss how packaging signals are constructed from local and long-range base pairings within viral genomes, as well as inter-molecular interactions between viral and host RNAs. Then, how genome packaging is regulated by the biophysical properties of RNA. Finally, we examine the impact of RNA packaging signals on viral evolution.

1. Introduction

Genome packaging is the process whereby viruses assemble their genomes into capsids [1]. The primary purpose of the capsid is to protect the genome from a hostile cellular and extracellular environment until its cargo can be released into a new host for a further round of replication. For faithful replication it is essential that genome packaging occurs with high fidelity. In the case of RNA viruses, this is particularly challenging because viral RNA must be specifically selected from a complex mix of cellular RNA, which often includes non-genomic viral RNA. Furthermore, genome packaging must be tightly regulated as it is often in competition with other essential functions, such as genome replication or translation. RNA viruses have solved this problem by exploiting the capacity of RNA to fold into three-dimensional structures that are recognised by viral packaging machinery [2,3,4,5,6,7,8]. RNA structures are formed from local intra-molecular and long-range base pairings within the same molecule, as well as inter-molecular RNA-RNA interactions [9]. Because RNA structures are rarely static, packaging can be dynamically regulated by intrinsic RNA structural switches, binding of viral factors, or in some cases, inter-molecular interactions with host RNAs.
Interestingly, many RNA viruses have evolved genome organizations that greatly complicate viral assembly and packaging. Segmented viruses, such as influenza and rotavirus, need to incorporate multiple genome segments for their virions to be infectious [10]. On the other hand, retroviruses package two copies of their genome, even though the total genetic material of only one genome is replicated [11]. In exchange for this increased complexity, RNA viruses enhance their evolvability through recombination or reassortment. Here, genome packaging uses specific inter-molecular RNA interactions that bring together different segments or genomes for assembly. In this review, we discuss how RNA viruses exploit the properties of RNA structure to regulate their packaging and explain how RNA based packaging mechanisms can influence viral evolution.

2. Packaging Signals in RNA Viruses

RNA viruses distinguish their genomes from cellular RNAs using cis acting packaging signals that serve as high affinity binding sites for the viral capsid (or nucleocapsid) proteins [12]. RNA readily folds back on itself to form secondary and tertiary structures, which are complex enough to enable the specific recognition of RNA by proteins [13]. A canonical example is the 19 nucleotide stem-loop, also known as TR, in the genome of the MS2 bacteriophage (Figure 1a) [14,15,16]. The MS2 coat protein (CP) dimer specifically recognizes this short stem loop structure to initiate assembly. Despite the simplicity of the structure, which is only formed from local base-pairings, MS2 CP binds with high affinity and specificity. This has led to the extensive repurposing of the TR-CP interaction in applications such as single molecule live cell imaging [17,18]. Remarkably, stem-loops having a C at position −5 in the loop have a higher affinity for CP than the wild-type U [19]. Whilst this is useful for biotechnology purposes, it also neatly demonstrates that increased affinity is not always beneficial for viruses, presumably because genomes must eventually be released during the early steps of the next replication cycle. Interestingly, high-throughput RNA structure-function analyses reveal that nucleotides in the stem can be exchanged without impairing binding to CP. In contrast, specific single stranded residues were required for function [20,21]. This is because, for the most part, RNA binding proteins (RBPs) make non-specific interactions with double stranded RNA (dsRNA) through generic contacts with the 2′hydroxyl groups of the ribose or the phosphodiester backbone [22]. For RNA viruses, non-specific affinity is usually driven by electrostatic interactions between positively charged patches on capsid protein subunits and the negatively charged nucleic acid genome. On the other hand, unique structural features created by loops and mismatches can be more readily recognized through extensive sequence-specific contacts [23].
MS2 bacteriophage genome packaging depends not only on the high affinity TR interaction site, but also on lower affinity pseudo-packaging sites that are dispersed throughout the genome (Figure 1b) [16,24]. This strategy is proposed to enhance the specificity and efficiency of assembly through cooperative interactions with viral capsid proteins [25]. This can be seen in other RNA viruses, such as tobacco necrosis virus (TNV) [26], hepatitis B virus (HBV) [27], and alphavirus [28]. Moreover, similar features emerged in directed evolution experiments that successfully converted a bacterial enzyme into a nucleocapsid that packages and protects its own encoding mRNA [29]. On the contrary, HIV-1 has no described packaging signals outside of the 5′ end of the genome. Instead, assembly on viral genomic RNA is driven by changes in the specificity of the structural protein Gag during its multimerization at the plasma membrane [30] (Figure 1c). This change in specificity enhances its affinity for A-rich sequences that are enriched in the HIV-1 genome compared to cellular RNA [30,31]. The prevailing model is that specific and high affinity binding sites nucleate viral assembly, whereas lower affinity binding sites and/or non-specific interactions are needed to drive growth and assembly of the capsid structure from capsid subunits [25,32]. Indeed, under physiological conditions, electrostatic repulsion between positively charged capsid subunits inhibit assembly to prevent the wasteful formation of empty capsids. This repulsion is counteracted by electrostatic attraction between capsid subunits and the negatively charged RNA genome. Essentially, electrostatic interactions provide the thermodynamic driving force for capsid assembly, converting the reversible capsid protein(s)-genome interaction into an irreversible assembly reaction [33,34]. RNA may even be considered as a structural component of virions. In the absence of genomic RNA, retroviruses package cellular RNAs, and retroviral cores can be disrupted upon treatment with RNase [35]. Moreover, recent work with the non-enveloped brome mosaic virus (BMV) has shown that when capsid subunits assemble on small non-viral RNAs it forms unstable shells with non-icosahedral structures that may be rapidly recycled into icosahedral capsids when the genomic RNA becomes available [36]. Furthermore, impressive cryo-electron microscopy (cryo-EM) reconstruction of the MS2 bacteriophage demonstrate how distinct, but non-uniformly distributed RNA structures favour assembly of the correct icosahedral capsid structure over non-productive octahedral structures [16]. Evidently, optimal viral assembly depends on the biophysical properties of the genomic RNA, which in turn is driven by a myriad of factors, including RNA sequence, structure [37,38] and size [36].
Complex viruses that express sub-genomic or spliced viral RNAs have an additional challenge: they must not only distinguish their genomic RNA from cellular RNA, but also from non-genomic viral RNAs (Figure 2). One simple way to achieve this selectivity is the removal of the packaging signal from the non-genomic RNA during its production. This mechanism occurs in certain retroviruses, such as Moloney murine leukemia virus (MoMLV), which contains a packaging signal with high affinity binding sites for viral nucleocapsid (NC) composed of three stem-loop structures (DIS-2, SL-C and SL-D) [39,40]. All these RNA structures lie downstream of the major splice donor site and are thus removed from spliced viral RNAs (Figure 2a). Another retrovirus, HIV-1, recognizes its genomic RNA through specific interactions between the viral Gag protein and packaging signals present at the 5′ end of the genome [41,42,43,44,45,46,47]. Early deletion mutagenesis studies identified SL3 (Ψ), which lies downstream of the major splice donor SL2, as the major packaging motif [46,47,48]. This genome organization was originally thought to explain the selectivity for genomic over spliced viral RNA [49,50]. However, an abundance of evidence has now revised this picture. Specifically, the basal part and internal loop of SL1, which lies upstream of SL2, is now recognized as the primary Gag binding site [21,51,52,53,54]. Notably, deletion or mutagenesis of SL1 has a more drastic effect on Gag binding and genome packaging compared to SL3, and deletion of sequences downstream of SL2 has only modest effects on binding [52,53]. This revision in understanding resurrected the problem of how HIV-1 discriminates between spliced viral RNA and genomic RNA. Surprisingly, genome fragments from the first nucleotide through to SL3—containing SL1—are not efficiently bound by Gag unless they contain sequences downstream of SL3 [51]. A model was proposed whereby a long-range interaction between sequences downstream of the splice donor site counteracts a negative regulatory element upstream of the high affinity binding site in SL1 [51] (Figure 2b). As this interaction can only be formed in genomic RNA, it enables the selectivity of Gag for genomic RNA over spliced viral RNA at the initial binding step [51,54,55,56,57]. A good candidate for this long-range interaction is the so-called U5-AUG base pairing, which is thought to promote genome packaging (discussed in the next section). However, it is not excluded that additional structural features within the packaging region SL1 to SL3, which are present only in genomic RNA, collaborate to promote nucleation of the Gag-RNA complex [21,41,57,58].
Coronaviruses (CoV) have extraordinarily large genomes (~30 kb) that presumably pose additional difficulties for packaging, yet genomic RNA is efficiently and selectively incorporated into virions [5,59,60,61]. Accumulating evidence suggests that SARS-CoV-2 exploits liquid-liquid phase separation (LLPS) during its replication [62,63,64,65,66,67,68,69,70] (Figure 3). LLPS occurs when biological molecules condense into a phase resembling a liquid droplet, and is an emerging paradigm for organizing membrane-less viral factories [71,72]. It is a common property of RBPs containing intrinsically disordered regions (IDR), such as the SARS-CoV-2 nucleocapsid (N) [64,67,68,73]. LLPS of N protein is enhanced in the presence of viral RNA [62,65], and even though N protein binds throughout the genomic RNA [62,70], LLPS is specifically promoted by RNA sequences at the 5′ and 3′ of the SARS-CoV-2 genome (Figure 3) [62]. Interestingly, other sequences, such as the CoV frameshift site, were found to disperse condensates [62], and importantly, sub-genomic RNA was efficiently excluded from preformed droplets [62]. This demonstrates that for LLPS mediated packaging, the biophysical properties of RNA-protein interaction are as important as the protein-RNA affinities. LLPS likely promotes viral assembly by enhancing interaction between RNA and N protein within a privileged site [62], but may have other roles in viral replication, such as hiding viral RNA from cellular immune sensors [74,75].
Packaging sites may also include motifs necessary for the correct presentation of the RNA molecule in time and space. For example, influenza viruses have a segmented genome of negative sense viral RNAs (vRNAs) that are replicated in the nucleus via complementary RNA (cRNA) intermediates. Long-range interactions between the 5′ and 3′ termini, in some cases over distances of thousands of nucleotides, construct the promoter structure that is involved in transcription, replication and packaging [76]. Interestingly, cRNAs and vRNAs are both complexed into ribonucleoproteins (vRNPs) with very similar protein compositions, but only vRNPs are packaged into virions. Slight differences in promoter structures between vRNPs and cRNPs, due to imperfect complementarity between the terminal sequences, affect its interaction with the viral M1 protein that acts as a bridge between vRNPs and the nuclear export machinery. This structural difference allows the virus to discriminate between cRNPs and vRNPs by either preventing nuclear export of the cRNP [77] or by changing nuclear export pathways [78] (Figure 3). In the same vein, several studies show that the binding of the HIV-1 Rev protein to its cognate RNA structure, the Rev Response Element (RRE), enhances genome packaging [79,80,81]. Surprisingly, this enhancement effect seems to be unrelated to the role of Rev/RRE in increasing cytoplasmic RNA levels. Rather, the Rev/RRE is proposed to enhance packaging by defining the correct nuclear export pathway and subcellular localization of the genomic RNA [79,80,81,82] (Figure 3).

3. RNA Structure as a Regulator of Genome Packaging

Genome packaging occurs during the late stages of replication when sufficient genomes and structural proteins have been replicated and produced to ensure effective viral assembly. Sometimes even, the same viral RNA molecule must carry out several competing functions. It is not surprising therefore that viruses heavily regulate the translation, replication, and packaging of their genomes. RNA viruses achieve this, in part, by exploiting the dynamic and flexible properties of RNA. Namely, RNA molecules can spontaneously fold into multiple, mutually exclusive structures, acting as riboswitches with each structure having a different function [83]. RNA can also respond to the binding of cellular or viral biomolecules, which can act as a regulatory trigger for further remodelling of ribonucleoprotein complexes [84].
Hepatitis C virus (HCV) is a model of such complex RNA based regulation [85,86,87]. The cis-acting replicating element (CRE) in the coding region of the NS5B protein forms a long distance base pairing with the highly conserved X-region in 3′ UTR [88,89,90,91,92] (Figure 4a). This interaction is required for replication, but also acts as a regulatory switch between replication and packaging by masking the core protein binding sites present in the 3′UTR [92] (Figure 4a). At the same time, the CRE regulates HCV genome translation via a long-range intra-molecular interaction with the internal ribosome entry site (IRES) in the 5′UTR [86,93] (Figure 4a). Finally, the 3′UTR X-region contains a palindromic sequence that promotes homo-dimerization of the HCV genome via a kissing loop inter-molecular RNA-RNA interaction [85,94,95] (Figure 4a). Since homo-dimerization is incompatible with the CRE-X interaction, and because it is likely tied to the concentration of genomes and viral chaperones in the cell, this mechanism is predicted to inhibit genome replication in favour of packaging late in the replication cycle. In this way, HCV elegantly fine tunes its replication using a complex network of dynamic and mutually exclusive RNA-RNA interactions [85,86,87].
Unsurprisingly, other viruses use similar principles to regulate their replication. HIV-1 genomic RNA is transcribed by the host cell and exported into the cytoplasm as a single pool of RNA that can be either selected by the viral Gag protein for packaging into viral particles or translated by host cell ribosomes [6]. A long-standing hypothesis is that the HIV-1 5′UTR adopts two alternative structural conformations to regulate the balance between genome translation and packaging (Figure 4b). Many structural models have been proposed, but all of them have the common feature that SL1 is presented in one conformation and sequestered in another [96,97,98,99,100,101,102,103]. As previously noted, SL1 is a key packaging motif in HIV-1 because the stem of SL1 contains the major Gag binding motif [51,104]. In addition to the Gag binding site, SL1 contains a six-nucleotide palindromic loop sequence that mediates an inter-molecular kissing loop interaction leading to the formation of genome dimers [105,106,107]. Unlike HCV, which produces homodimers that remain in the cell, HIV-1 dimers are packaged into virions [108]. This process, known as dimerization, is a conserved feature of retroviral replication that is assumed, but not formally proven, to be a pre-requisite for packaging [109]. A series of NMR studies have identified a region, U5, in the 5′UTR that base pairs with the loop sequence of SL1, or alternatively with a region surrounding the AUG start site [98,100,101,110]. When the SL1 loop sequence is base paired with U5, genomic RNA is monomeric, which promotes translation (Figure 4b). When U5 is base paired with a region surrounding the AUG start codon, the SL1 loop is available for dimerization and packaging [111] (Figure 4b). Remarkably, transcription start site heterogeneity inherent to the HIV-1 promotor strongly influences the equilibrium between these two structures [98,112]. HIV-1 genomes transcribed with a single guanosine favour the dimer conformation and are packaged into viral particles, while genomes transcribed with two or three guanosines form monomers that are preferentially translated [98,112] (Figure 4b). The fact that a single GC base-pair perturbs the monomer-dimer equilibrium provides striking proof that viruses exploit metastable RNA structures in their regulation.
Added complexity comes from the fact that RNA viruses also regulate their replication using inter-molecular interactions between host RNAs and their genomes. The HCV 5′UTR contains binding sites for the host micro-RNA miR-122 [113,114,115] (Figure 4a). miR-122 is essential for the stability of HCV genomic RNA by inhibiting RNA decay by Xrn exonucleases [116,117,118]. Binding of miR-122 also increases HCV genome translation [119,120,121,122,123] and replication [124,125] by other mechanisms. Several lines of evidence suggest that miR-122 can act by inducing RNA structural changes in the 5′UTR. Specifically, miR-122 either alone or in partnership with Argonaute proteins, enhances translation by promoting the folding of a functional IRES and suppressing alternative folds of the 5′ UTR that interfere with IRES function [121,123]. Others have proposed that miR-122 enhances translation by facilitating cyclization of the genome, by promoting stranded separation of the replication intermediates, or by bringing or displacing protein co-factors to the genome [124]. Similarly, the HIV-1 5′UTR contains a binding site for a host cellular tRNALys3, which is used as a primer for reverse transcription [126,127] (Figure 4b). Its binding results in RNA conformational changes that favour dimerization, and presumably packaging [103,128].

4. Intermolecular RNA-RNA Interactions in Segmented Viruses

Many viruses split their genome into smaller independent segments. This causes problems for genome assembly, which is solved using one of two strategies: random vs selective packaging (Figure 5). Tri-segmented bunyaviruses, such as the rift valley fever virus (RVFV) and Schmallenberg virus (SBV), use the simpler strategy of random incorporation [129,130] (Figure 5a). Single molecule fluorescent in situ hybridisation (smFISH) revealed only 1 in 10 RVFV particles contain the full complement of genome segments due to the inherent heterogeneity in this packaging strategy [129]. Intuitively, as the number of segments increase, the probability of packaging one copy per particle decreases rapidly unless a large number of genome segments are incorporated per particle [131]. As this is not very efficient, many segmented viruses have overcome the genome assembly problem with specific packaging signals, which allow each distinct segment to be identified and packaged (Figure 5b).
A well-studied example is influenza A virus of the Orthomyxoviridae family. Its genome consists of eight negative-sense viral RNAs (vRNAs) that are packaged into viral particles as viral ribonucleoprotein (vRNP) complexes [131]. Because each vRNA encodes for an essential protein, every infectious viral particle must contain at least one copy of each segment. Indeed, smFISH experiments prove that most viral particles contain precisely one of each segment [132,133]. Furthermore, numerous electron tomography studies demonstrate that influenza vRNPs in budding viruses adopt an arrangement, also known as ‘1 + 7’ conformation, in which seven vRNPs surround a central one [134,135,136]. Altogether, these data argue for a selective packaging process. Defective interfering (DI) RNA, which naturally arise in cell culture at high multiplicity of infection (MOI), retain 100–300 nucleotides from their terminal sequences indicating that these regions contain packaging signals [137,138,139,140]. Indeed, deletion and mutagenesis studies have grossly defined terminal packaging regions within all eight vRNAs [141,142,143,144,145,146,147,148,149,150]. Terminal packaging signals are proposed to be bipartite, containing a non-specific “incorporation signal” in the UTR/promoter region, and a specific “bundling signal” in the terminal coding regions [150]. The hypothesized incorporation signal directs vRNP packaging into virions, whereas the bundling signal allows discrimination between vRNPs. The mechanism mediating this phenomenon is still not completely understood, but the most popular explanation is that packaging signals discriminate between segments by defining direct and segment specific inter-molecular RNA-RNA interactions (Figure 5c). In support of this idea, electron microscopy studies show frequent physical contacts through the entire length of each vRNP with a string-like form reminiscent of RNA [135,136,151,152], and vRNAs are able to form RNA-RNA interactions in vitro [136,153,154]. The prevailing model is that influenza vRNPs are packaged as a supramolecular complex that are held together through a network of interactions where each vRNA contacts at least one other vRNA [8]. This would help explain why mutations to packaging signals in one vRNA often affected the packaging of other vRNAs [146,149,155]. Furthermore, the capacity of RNA to tolerate mutations without disrupting structure and function would explain why packaging site mutations do not always give rise to phenotypic effects. Importantly, several vRNA-vRNA interactions have been characterised at the nucleotide level proving that at least some packaging signals define direct RNA-RNA contacts [153,156,157]. These recent results have spurred efforts to map more completely inter-segment interactions in influenza using high-throughput sequencing and RNA proximity ligation technologies [157,158]. Collectively, these studies have revealed that the inter-molecular RNA-RNA interactions are extensive, with frequent contacts seen throughout vRNAs, including in the central coding regions. Nevertheless, comprehensive maps of direct vRNA-vRNA contacts have been surprisingly difficult to interpret, with many interactions having no apparent functional role. One major conclusion could be that vRNA packaging signals are complex and redundant, but it could also reflect deficiencies in contact map technology which may miss important interactions because of known inefficiencies and biases imposed by crosslinking reagents [159] or ligation strategies [160]. Furthermore, the role of protein-RNA and protein-protein interactions in this process is not excluded. As a matter of fact, influenza nucleoprotein (NP) provides an additional layer of complexity to this process as it incompletely coats the vRNA and helps to define which vRNA sequences are available to form inter-molecular interactions [161,162,163].
Similar principles seem to apply to dsRNA segmented viruses of the Reoviridae family, which includes rotavirus and bluetongue virus. Rotavirus has a genome composed of eleven dsRNA segments of different sequences and lengths (0.7 to 3.1 kb) [164]. Paralleling recent results in influenza, inter-molecular interactions between segments 9, 10 and 11 of the rotavirus RNAs were observed in vitro [165]. Disruption of the putative interaction sites by mutation or with oligoribonucleotides inhibited complex formation and viral replication in cell culture [165]. Studies with bluetongue virus, which has 10 dsRNA segments, suggest a model whereby assembly begins with the formation of an initial complex built of the small RNA segments [166,167,168]. This complex would then serve as a base for sequentially recruiting the remaining RNA segments to ultimately generate a complete complex that is packaged into virions. In the case of influenza, smFISH studies reveal that sequential vRNP-vRNP interactions occur en route to the plasma membrane where packaging takes place [133,169,170]. However, current evidence indicates that there is not a single assembly pathway, but a number of alternative preferred pathways that nevertheless prevent the incorporation of more than one copy of each segment [133]. How this is achieved at the mechanistic level is still an open question.

5. RNA Packaging and Evolution

RNA based packaging signals play a much broader role in viral life cycles than assembly, and it is now appreciated that RNA virus genome structures enhance virus ‘evolvability’ by facilitating gene exchange during co-infection [171]. One widespread strategy is template switching during replication leading to recombination and the formation of genome chimeras, which is a conserved phenomenon in retroviruses [11,172] (Figure 6a). Another common strategy is genome segmentation leading to reassortment, which can be seen in rotaviruses and influenza viruses [173] (Figure 6b). Recombination and reassortment are both non-random processes that are heavily biased by RNA sequence and structure.
Retroviruses package two near identical copies of the genome as a non-covalently associated dimer [108]. One evolutionary advantage for this dimeric genome organization is that it brings together two templates for packaging into virions. Template switching during subsequent infection and reverse transcription generates a recombinant virus that is genetically distinct from the two parental viruses [174,175,176,177,178] (Figure 6a). Retroviral recombination is a major mechanism by which retroviruses escape selective pressures imposed by the immune system or antiretroviral therapy [11]. As previously noted, HIV-1 dimerization is mediated by the palindromic loop sequence of SL1 [105,106,107]. Sequence variations that are unable to form the kissing loop interaction are also defective in recombination due to their inability to be co-packaged into virions [179]. Indeed, the loop sequences in subtype B (GCGCGC), and subtypes A, C and G (GUGCAC) are incompatible. Inter-subtype recombination is thus much lower compared to intra-subtype recombination [105,179,180]. Interestingly, HIV-1 genomes containing deletions in SL1 are still packaged into virions as dimers, albeit at a lower level than wild-type viruses [181,182]. This provides strong evidence that so-far undetected inter-molecular interaction exist throughout the HIV-1 genome that may enable the formation inter-subtype recombinants even viruses are unable to form the kissing-loop interaction at SL1.
Packaging signal incompatibilities are also thought to be a major restriction to reassortment in segmented viruses [156,157,183,184,185,186]. This is especially important for influenza where introductions of sequence variation from animal reservoirs have led to pandemics in the past [187]. Fortunately, divergence in packaging signals between human and animal viruses is one of many steps that may block reassortment [184,185]. The molecular mechanism restricting reassortment probably lies in the inability of divergent sequences to form intermolecular vRNA-vRNA interactions required for packaging [156,157,183]. This provides hope that in the future, better knowledge of viral structures may be repurposed to predict or even direct viral evolution to combat both emerging and endemic RNA viruses.

6. Outlook

Generally, viral RNA packaging is assumed to be a process dependent on a few clearly defined RNA structural motifs specifically recognized by a viral protein. However, when evaluating binding affinities and specificities of those RNA-protein complexes in vitro, they often do not show the specificity that is observed for the packaging process in vivo [25]. Thus, it may be reasonable to think of packaging as an integrative process that involves multiple co-occurrent interactions that must also take place at the correct time and subcellular localization for genome packaging to occur.
Excitingly, methods to characterize RNA virus packaging signals are being developed and refined to better resolve the details of these integrative processes. Approaches to study viral RNA packaging spans disciplines and can now shed light on this process across multiple scales. For example, advanced cryo-EM techniques promise to determine RNA structures at high resolution in three-dimensions [188,189,190]. Furthermore, cryo-EM [191,192,193,194] and X-ray scattering [195,196] may reveal RNA-protein interaction sites inside of viral capsids. In parallel, RNA structural probing techniques are being developed that enable the detection of structural changes in RNA that may be the result of RNA packaging, e.g., by identifying alternative structures [197,198] and/or mapping RNA-protein interaction sites on the RNA [199,200]. Continual improvements in quantitative live, super resolution, and expansion microscopy will be key for understanding mechanisms of viral assembly in cells [133,201,202,203,204,205]. These improvements are beginning to reveal how inherent variability in viral assembly allow viruses to replicate and evolve in the face of complex and unpredictable environments [133,203]. Finally, comparative high throughput sequencing can identify RNA packaging signals, be it historically from identifying genomic constraints of packaging-competent defective viral genomes [206,207,208,209,210], or more recently by reverse genetics systems that quantify relative packaging efficiencies of large pools of mutants in parallel [53]. Together, these technical revolutions are sure to dramatically improve our understanding of the molecular mechanisms of viral RNA genome packaging across virus families. In the near future, these insights can be pivoted into novel antiviral drugs and vaccines for controlling these important human pathogens.

Author Contributions

Conceptualization, L.Y. and R.P.S.; writing—original draft preparation, L.Y. and R.P.S.; writing—review and editing, all authors; visualization, all authors; supervision, R.P.S.; funding acquisition, R.P.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Helmholtz Association (VH-NG-1347 to R.S.) and the Bundesministerium für Bildung und Forschung (BMBF) (COMPLS-182 to R.S.). A.S.G. was supported with a fellowship from the Peter und Traudl Engelhorn Stiftung. U.A. was supported by a fellowship from the German Academic Exchange Service (DAAD).

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the writing of the manuscript, or in the decision to publish.

References

  1. Sun, S.; Rao, V.B.; Rossmann, M.G. Genome Packaging in Viruses. Curr. Opin. Struct. Biol. 2010, 20, 114–120. [Google Scholar] [CrossRef] [Green Version]
  2. Adams, R.L.; Pirakitikulr, N.; Pyle, A.M. Functional RNA Structures throughout the Hepatitis C Virus Genome. Curr. Opin. Virol. 2017, 24, 79–86. [Google Scholar] [CrossRef] [PubMed]
  3. Wilkinson, K.; Gorelick, R.J.; Vasa, S.M.; Guex, N.; Rein, A.; Mathews, D.H.; Giddings, M.C.; Weeks, K.M. High-Throughput SHAPE Analysis Reveals Structures in HIV-1 Genomic RNA Strongly Conserved across Distinct Biological States. PLoS Biol. 2008, 6, e96. [Google Scholar] [CrossRef] [Green Version]
  4. Rangan, R.; Zheludev, I.N.; Hagey, R.J.; Pham, E.A.; Wayment-Steele, H.K.; Glenn, J.S.; Das, R. RNA Genome Conservation and Secondary Structure in SARS-CoV-2 and SARS-Related Viruses: A First Look. RNA 2020, 26, 937–959. [Google Scholar] [CrossRef] [PubMed]
  5. Masters, P.S. Coronavirus Genomic RNA Packaging. Virology 2019, 537, 198–207. [Google Scholar] [CrossRef]
  6. Mailler, E.; Bernacchi, S.; Marquet, R.; Paillart, J.-C.; Vivet-Boudou, V.; Smyth, R.P.; Mailler, E.; Bernacchi, S.; Marquet, R.; Paillart, J.-C.; et al. The Life-Cycle of the HIV-1 Gag-RNA Complex. Viruses 2016, 8, 248. [Google Scholar] [CrossRef] [PubMed]
  7. Borodavka, A.; Desselberger, U.; Patton, J.T. Genome Packaging in Multi-Segmented DsRNA Viruses: Distinct Mechanisms with Similar Outcomes. Curr. Opin. Virol. 2018, 33, 106–112. [Google Scholar] [CrossRef] [PubMed]
  8. Gerber, M.; Isel, C.; Moules, V.; Marquet, R. Selective Packaging of the Influenza A Genome and Consequences for Genetic Reassortment. Trends Microbiol. 2014, 22, 446–455. [Google Scholar] [CrossRef] [PubMed]
  9. Cruz, J.A.; Westhof, E. The Dynamic Landscapes of RNA Architecture. Cell 2009, 136, 604–609. [Google Scholar] [CrossRef] [PubMed]
  10. Newburn, L.R.; White, K.A. Trans-Acting RNA-RNA Interactions in Segmented RNA Viruses. Viruses 2019, 11, 751. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  11. Smyth, R.P.; Davenport, M.P.; Mak, J. The Origin of Genetic Diversity in HIV-1. Virus Res. 2012, 169, 415–429. [Google Scholar] [CrossRef] [PubMed]
  12. Comas-Garcia, M. Packaging of Genomic RNA in Positive-Sense Single-Stranded RNA Viruses: A Complex Story. Viruses 2019, 11, 253. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Sanchez de Groot, N.; Armaos, A.; Graña-Montes, R.; Alriquet, M.; Calloni, G.; Vabulas, R.M.; Tartaglia, G.G. RNA Structure Drives Interaction with Proteins. Nat. Commun. 2019, 10, 1–13. [Google Scholar] [CrossRef] [PubMed]
  14. Peabody, D.S. The RNA Binding Site of Bacteriophage MS2 Coat Protein. EMBO J. 1993, 12, 595–600. [Google Scholar] [CrossRef] [PubMed]
  15. Ni, C.Z.; Syed, R.; Kodandapani, R.; Wickersham, J.; Peabody, D.S.; Ely, K.R. Crystal Structure of the MS2 Coat Protein Dimer: Implications for RNA Binding and Virus Assembly. Structure 1995, 3, 255–263. [Google Scholar] [CrossRef] [Green Version]
  16. Dai, X.; Li, Z.; Lai, M.; Shu, S.; Du, Y.; Zhou, Z.H.; Sun, R. In Situ Structures of the Genome and Genome-Delivery Apparatus in a Single-Stranded RNA Virus. Nature 2017, 541, 112–116. [Google Scholar] [CrossRef] [Green Version]
  17. Halstead, J.M.; Lionnet, T.; Wilbertz, J.H.; Wippich, F.; Ephrussi, A.; Singer, R.H.; Chao, J.A. Translation. An RNA Biosensor for Imaging the First Round of Translation from Single Cells to Living Animals. Science 2015, 347, 1367–1671. [Google Scholar] [CrossRef] [Green Version]
  18. Bertrand, E.; Chartrand, P.; Schaefer, M.; Shenoy, S.M.; Singer, R.H.; Long, R.M. Localization of ASH1 MRNA Particles in Living Yeast. Mol. Cell 1998, 2, 437–445. [Google Scholar] [CrossRef] [Green Version]
  19. Horn, W.T.; Convery, M.A.; Stonehouse, N.J.; Adams, C.J.; Liljas, L.; Phillips, S.E.V.; Stockley, P.G. The Crystal Structure of a High Affinity RNA Stem-Loop Complexed with the Bacteriophage MS2 Capsid: Further Challenges in the Modeling of Ligand-RNA Interactions. RNA 2004, 10, 1776–1782. [Google Scholar] [CrossRef] [Green Version]
  20. Buenrostro, J.D.; Araya, C.L.; Chircus, L.M.; Layton, C.J.; Chang, H.Y.; Snyder, M.P.; Greenleaf, W.J. Quantitative Analysis of RNA-Protein Interactions on a Massively Parallel Array Reveals Biophysical and Evolutionary Landscapes. Nat. Biotechnol. 2014, 32, 562–568. [Google Scholar] [CrossRef]
  21. Smyth, R.P.; Despons, L.; Huili, G.; Bernacchi, S.; Hijnen, M.; Mak, J.; Jossinet, F.; Weixi, L.; Paillart, J.C.; Von Kleist, M.; et al. Mutational Interference Mapping Experiment (MIME) for Studying RNA Structure and Function. Nat. Methods 2015, 12, 866–872. [Google Scholar] [CrossRef] [PubMed]
  22. Masliah, G.; Barraud, P.; Allain, F.H.-T. RNA Recognition by Double-Stranded RNA Binding Domains: A Matter of Shape and Sequence. Cell. Mol. Life Sci. 2013, 70, 1875–1895. [Google Scholar] [CrossRef] [Green Version]
  23. Treger, M.; Westhof, E. Statistical Analysis of Atomic Contacts at RNA-Protein Interfaces. J. Mol. Recognit. 2001, 14, 199–214. [Google Scholar] [CrossRef] [PubMed]
  24. Rolfsson, Ó.; Middleton, S.; Manfield, I.W.; White, S.J.; Fan, B.; Vaughan, R.; Ranson, N.A.; Dykeman, E.; Twarock, R.; Ford, J.; et al. Direct Evidence for Packaging Signal-Mediated Assembly of Bacteriophage MS2. J. Mol. Biol. 2016, 428, 431–448. [Google Scholar] [CrossRef] [Green Version]
  25. Twarock, R.; Stockley, P.G. RNA-Mediated Virus Assembly: Mechanisms and Consequences for Viral Evolution and Therapy. Annu. Rev. Biophys. 2019, 48, 495–514. [Google Scholar] [CrossRef] [PubMed]
  26. Patel, N.; Wroblewski, E.; Leonov, G.; Phillips, S.E.V.; Tuma, R.; Twarock, R.; Stockley, P.G. Rewriting Nature’s Assembly Manual for a SsRNA Virus. Proc. Natl. Acad. Sci. USA 2017, 114, 12255–12260. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Patel, N.; White, S.J.; Thompson, R.F.; Bingham, R.; Weiß, E.U.; Maskell, D.P.; Zlotnick, A.; Dykeman, E.; Tuma, R.; Twarock, R.; et al. HBV RNA Pre-Genome Encodes Specific Motifs That Mediate Interactions with the Viral Core Protein That Promote Nucleocapsid Assembly. Nat. Microbiol. 2017, 2, 17098. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Brown, R.S.; Anastasakis, D.G.; Hafner, M.; Kielian, M. Multiple Capsid Protein Binding Sites Mediate Selective Packaging of the Alphavirus Genomic RNA. Nat. Commun. 2020, 11, 4693. [Google Scholar] [CrossRef] [PubMed]
  29. Tetter, S.; Terasaka, N.; Steinauer, A.; Bingham, R.J.; Clark, S.; Scott, A.J.P.; Patel, N.; Leibundgut, M.; Wroblewski, E.; Ban, N.; et al. Evolution of a Virus-like Architecture and Packaging Mechanism in a Repurposed Bacterial Protein. Science 2021, 372, 1220–1224. [Google Scholar] [CrossRef]
  30. Kutluay, S.B.; Zang, T.; Blanco-Melo, D.; Powell, C.; Jannain, D.; Errando, M.; Bieniasz, P.D. Global Changes in the RNA Binding Specificity of HIV-1 Gag Regulate Virion Genesis. Cell 2014, 159, 1096–1109. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  31. Keating, C.P.; Hill, M.K.; Hawkes, D.J.; Smyth, R.P.; Isel, C.; Le, S.-Y.Y.; Palmenberg, A.C.; Marshall, J.A.; Marquet, R.; Nabel, G.J.; et al. The A-Rich RNA Sequences of HIV-1 Pol Are Important for the Synthesis of Viral CDNA. Nucleic Acids Res. 2009, 37, 945–956. [Google Scholar] [CrossRef] [PubMed]
  32. Zandi, R.; van der Schoot, P.; Reguera, D.; Kegel, W.; Reiss, H. Classical Nucleation Theory of Virus Capsids. Biophys. J. 2006, 90, 1939–1948. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. van der Schoot, P.; Bruinsma, R. Electrostatics and the Assembly of an RNA Virus. Phys. Rev. E Stat. Nonlin. Soft Matter Phys. 2005, 71, 061928. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Tanwar, H.S.; Khoo, K.K.; Garvey, M.; Waddington, L.; Leis, A.; Hijnen, M.; Velkov, T.; Dumsday, G.J.; McKinstry, W.J.; Mak, J. The Thermodynamics of Pr55Gag-RNA Interaction Regulate the Assembly of HIV. PLoS Pathog. 2017, 13, e1006221. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Muriaux, D.; Mirro, J.; Harvin, D.; Rein, A. RNA Is a Structural Element in Retrovirus Particles. Proc. Natl. Acad. Sci. USA 2001, 98, 5246–5251. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Bond, K.; Tsvetkova, I.B.; Wang, J.C.-Y.; Jarrold, M.F.; Dragnea, B. Virus Assembly Pathways: Straying Away but Not Too Far. Small 2020, 16, e2004475. [Google Scholar] [CrossRef]
  37. Perlmutter, J.D.; Qiao, C.; Hagan, M.F. Viral Genome Structures Are Optimal for Capsid Assembly. Elife 2013, 2, e00632. [Google Scholar] [CrossRef] [Green Version]
  38. Erdemci-Tandogan, G.; Wagner, J.; van der Schoot, P.; Podgornik, R.; Zandi, R. Effects of RNA Branching on the Electrostatic Stabilization of Viruses. Phys. Rev. E 2016, 94, 022408. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  39. Mougel, M.; Zhang, Y.; Barklis, E. Cis-Active Structural Motifs Involved in Specific Encapsidation of Moloney Murine Leukemia Virus RNA. J. Virol. 1996, 70, 5043–5050. [Google Scholar] [CrossRef] [Green Version]
  40. Murphy, J.E.; Goff, S.P. Construction and Analysis of Deletion Mutations in the U5 Region of Moloney Murine Leukemia Virus: Effects on RNA Packaging and Reverse Transcription. J. Virol. 1989, 63, 319–327. [Google Scholar] [CrossRef] [Green Version]
  41. Ding, P.; Kharytonchyk, S.; Waller, A.; Mbaekwe, U.; Basappa, S.; Kuo, N.; Frank, H.M.; Quasney, C.; Kidane, A.; Swanson, C.; et al. Identification of the Initial Nucleocapsid Recognition Element in the HIV-1 RNA Packaging Signal. Proc. Natl. Acad. Sci. USA 2020, 117, 17737–17746. [Google Scholar] [CrossRef]
  42. Lu, K.; Heng, X.; Summers, M.F. Structural Determinants and Mechanism of HIV-1 Genome Packaging. J. Mol. Biol. 2011, 410, 609–633. [Google Scholar] [CrossRef] [Green Version]
  43. Luban, J.; Goff, S.P. Binding of Human Immunodeficiency Virus Type 1 (HIV-1) RNA to Recombinant HIV-1 Gag Polyprotein. J. Virol. 1991, 65, 3203–3212. [Google Scholar] [CrossRef] [Green Version]
  44. Luban, J.; Goff, S.P. Mutational Analysis of Cis-Acting Packaging Signals in Human Immunodeficiency Virus Type 1 RNA. J. Virol. 1994, 68, 3784–3793. [Google Scholar] [CrossRef] [Green Version]
  45. McBride, M.S.; Panganiban, A.T. The Human Immunodeficiency Virus Type 1 Encapsidation Site Is a Multipartite RNA Element Composed of Functional Hairpin Structures. J. Virol. 1996, 70, 2963–2973. [Google Scholar] [CrossRef] [Green Version]
  46. Lever, A.; Gottlinger, H.; Haseltine, W.; Sodroski, J. Identification of a Sequence Required for Efficient Packaging of Human Immunodeficiency Virus Type 1 RNA into Virions. J. Virol. 1989, 63, 4085–4087. [Google Scholar] [CrossRef] [Green Version]
  47. Aldovini, A.; Young, R.A. Mutations of RNA and Protein Sequences Involved in Human Immunodeficiency Virus Type 1 Packaging Result in Production of Noninfectious Virus. J. Virol. 1990, 64, 1920–1926. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  48. Clavel, F.; Orenstein, J.M. A Mutant of Human Immunodeficiency Virus with Reduced RNA Packaging and Abnormal Particle Morphology. J. Virol. 1990, 64, 5230–5234. [Google Scholar] [CrossRef] [Green Version]
  49. De Guzman, R.N.; Wu, Z.R.; Stalling, C.C.; Pappalardo, L.; Borer, P.N.; Summers, M.F. Structure of the HIV-1 Nucleocapsid Protein Bound to the SL3 Psi-RNA Recognition Element. Science 1998, 279, 384–388. [Google Scholar] [CrossRef] [PubMed]
  50. Amarasinghe, G.K.; De Guzman, R.N.; Turner, R.B.; Chancellor, K.J.; Wu, Z.R.; Summers, M.F. NMR Structure of the HIV-1 Nucleocapsid Protein Bound to Stem-Loop SL2 of the Psi-RNA Packaging Signal. Implications for Genome Recognition. J. Mol. Biol. 2000, 301, 491–511. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  51. Abd El-Wahab, E.W.; Smyth, R.P.; Mailler, E.; Bernacchi, S.; Vivet-Boudou, V.; Hijnen, M.; Jossinet, F.; Mak, J.; Paillart, J.-C.; Marquet, R. Specific Recognition of the HIV-1 Genomic RNA by the Gag Precursor. Nat. Commun. 2014, 5, 4304. [Google Scholar] [CrossRef] [Green Version]
  52. Houzet, L.; Paillart, J.C.; Smagulova, F.; Maurel, S.; Morichaud, Z.; Marquet, R.; Mougel, M. HIV Controls the Selective Packaging of Genomic, Spliced Viral and Cellular RNAs into Virions through Different Mechanisms. Nucleic Acids Res. 2007, 35, 2695–2704. [Google Scholar] [CrossRef]
  53. Smyth, R.P.; Smith, M.R.; Jousset, A.-C.; Despons, L.; Laumond, G.; Decoville, T.; Cattenoz, P.; Moog, C.; Jossinet, F.; Mougel, M.; et al. In Cell Mutational Interference Mapping Experiment (in Cell MIME) Identifies the 5′ Polyadenylation Signal as a Dual Regulator of HIV-1 Genomic RNA Production and Packaging. Nucleic Acids Res. 2018, 46, e57. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Bernacchi, S.; Abd El-Wahab, E.W.; Dubois, N.; Hijnen, M.; Smyth, R.P.; Mak, J.; Marquet, R.; Paillart, J.-C. HIV-1 Pr55Gag Binds Genomic and Spliced RNAs with Different Affinity and Stoichiometry. RNA Biol. 2017, 14, 90–103. [Google Scholar] [CrossRef] [Green Version]
  55. Comas-Garcia, M.; Datta, S.A.; Baker, L.; Varma, R.; Gudla, P.R.; Rein, A. Dissection of Specific Binding of HIV-1 Gag to the “packaging Signal” in Viral RNA. Elife 2017, 6, e27055. [Google Scholar] [CrossRef] [PubMed]
  56. Rein, A. RNA Packaging in HIV. Trends Microbiol. 2019, 27, 715–723. [Google Scholar] [CrossRef] [PubMed]
  57. Webb, J.; Jones, C.P.; Parent, L.J.; Rouzina, I.; Musier-Forsyth, K. Distinct Binding Interactions of HIV-1 Gag to Psi and Non-Psi RNAs: Implications for Viral Genomic RNA Packaging. RNA 2013, 19, 1078–1088. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. Comas-Garcia, M.; Kroupa, T.; Datta, S.A.; Harvin, D.P.; Hu, W.-S.; Rein, A. Efficient Support of Virus-like Particle Assembly by the HIV-1 Packaging Signal. Elife 2018, 7, e38438. [Google Scholar] [CrossRef]
  59. Hsin, W.-C.; Chang, C.-H.; Chang, C.-Y.; Peng, W.-H.; Chien, C.-L.; Chang, M.-F.; Chang, S.C. Nucleocapsid Protein-Dependent Assembly of the RNA Packaging Signal of Middle East Respiratory Syndrome Coronavirus. J. Biomed. Sci. 2018, 25, 47. [Google Scholar] [CrossRef]
  60. Hsieh, P.-K.; Chang, S.C.; Huang, C.-C.; Lee, T.-T.; Hsiao, C.-W.; Kou, Y.-H.; Chen, I.-Y.; Chang, C.-K.; Huang, T.-H.; Chang, M.-F. Assembly of Severe Acute Respiratory Syndrome Coronavirus RNA Packaging Signal into Virus-like Particles Is Nucleocapsid Dependent. J. Virol. 2005, 79, 13848–13855. [Google Scholar] [CrossRef] [Green Version]
  61. Klein, S.; Cortese, M.; Winter, S.L.; Wachsmuth-Melm, M.; Neufeldt, C.J.; Cerikan, B.; Stanifer, M.L.; Boulant, S.; Bartenschlager, R.; Chlanda, P. SARS-CoV-2 Structure and Replication Characterized by in Situ Cryo-Electron Tomography. Nat. Commun. 2020, 11, 5885. [Google Scholar] [CrossRef]
  62. Iserman, C.; Roden, C.A.; Boerneke, M.A.; Sealfon, R.S.G.; McLaughlin, G.A.; Jungreis, I.; Fritch, E.J.; Hou, Y.J.; Ekena, J.; Weidmann, C.A.; et al. Genomic RNA Elements Drive Phase Separation of the SARS-CoV-2 Nucleocapsid. Mol. Cell 2020, 80, 1078–1091.e6. [Google Scholar] [CrossRef]
  63. Wang, J.; Shi, C.; Xu, Q.; Yin, H. SARS-CoV-2 Nucleocapsid Protein Undergoes Liquid-Liquid Phase Separation into Stress Granules through Its N-Terminal Intrinsically Disordered Region. Cell Discov. 2021, 7, 5. [Google Scholar] [CrossRef]
  64. Savastano, A.; Ibáñez de Opakua, A.; Rankovic, M.; Zweckstetter, M. Nucleocapsid Protein of SARS-CoV-2 Phase Separates into RNA-Rich Polymerase-Containing Condensates. Nat. Commun. 2020, 11, 6041. [Google Scholar] [CrossRef]
  65. Chen, H.; Cui, Y.; Han, X.; Hu, W.; Sun, M.; Zhang, Y.; Wang, P.-H.; Song, G.; Chen, W.; Lou, J. Liquid-Liquid Phase Separation by SARS-CoV-2 Nucleocapsid Protein and RNA. Cell Res. 2020, 30, 1143–1145. [Google Scholar] [CrossRef]
  66. Carlson, C.R.; Asfaha, J.B.; Ghent, C.M.; Howard, C.J.; Hartooni, N.; Safari, M.; Frankel, A.D.; Morgan, D.O. Phosphoregulation of Phase Separation by the SARS-CoV-2 N Protein Suggests a Biophysical Basis for Its Dual Functions. Mol. Cell 2020, 80, 1092–1103.e4. [Google Scholar] [CrossRef]
  67. Perdikari, T.M.; Murthy, A.C.; Ryan, V.H.; Watters, S.; Naik, M.T.; Fawzi, N.L. SARS-CoV-2 Nucleocapsid Protein Phase-Separates with RNA and with Human HnRNPs. EMBO J. 2020, 39, e106478. [Google Scholar] [CrossRef] [PubMed]
  68. Wu, Y.; Ma, L.; Cai, S.; Zhuang, Z.; Zhao, Z.; Jin, S.; Xie, W.; Zhou, L.; Zhang, L.; Zhao, J.; et al. RNA-Induced Liquid Phase Separation of SARS-CoV-2 Nucleocapsid Protein Facilitates NF-ΚB Hyper-Activation and Inflammation. Signal Transduct. Target. Ther. 2021, 6, 167. [Google Scholar] [CrossRef] [PubMed]
  69. Cubuk, J.; Alston, J.J.; Incicco, J.J.; Singh, S.; Stuchell-Brereton, M.D.; Ward, M.D.; Zimmerman, M.I.; Vithani, N.; Griffith, D.; Wagoner, J.A.; et al. The SARS-CoV-2 Nucleocapsid Protein Is Dynamic, Disordered, and Phase Separates with RNA. Nat. Commun. 2021, 12, 1936. [Google Scholar] [CrossRef] [PubMed]
  70. Lu, S.; Ye, Q.; Singh, D.; Cao, Y.; Diedrich, J.K.; Yates, J.R.; Villa, E.; Cleveland, D.W.; Corbett, K.D. The SARS-CoV-2 Nucleocapsid Phosphoprotein Forms Mutually Exclusive Condensates with RNA and the Membrane-Associated M Protein. Nat. Commun. 2021, 12, 502. [Google Scholar] [CrossRef]
  71. Guseva, S.; Milles, S.; Jensen, M.R.; Salvi, N.; Kleman, J.-P.; Maurin, D.; Ruigrok, R.W.H.; Blackledge, M. Measles Virus Nucleo- and Phosphoproteins Form Liquid-like Phase-Separated Compartments That Promote Nucleocapsid Assembly. Sci. Adv. 2020, 6, eaaz7095. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Etibor, T.A.; Yamauchi, Y.; Amorim, M.J. Liquid Biomolecular Condensates and Viral Lifecycles: Review and Perspectives. Viruses 2021, 13, 366. [Google Scholar] [CrossRef] [PubMed]
  73. Protter, D.S.W.; Rao, B.S.; Van Treeck, B.; Lin, Y.; Mizoue, L.; Rosen, M.K.; Parker, R. Intrinsically Disordered Regions Can Contribute Promiscuous Interactions to RNP Granule Assembly. Cell Rep. 2018, 22, 1401–1412. [Google Scholar] [CrossRef] [Green Version]
  74. Wang, S.; Dai, T.; Qin, Z.; Pan, T.; Chu, F.; Lou, L.; Zhang, L.; Yang, B.; Huang, H.; Lu, H.; et al. Targeting Liquid-Liquid Phase Separation of SARS-CoV-2 Nucleocapsid Protein Promotes Innate Antiviral Immunity by Elevating MAVS Activity. Nat. Cell Biol. 2021, 23, 718–732. [Google Scholar] [CrossRef] [PubMed]
  75. Zhao, M.; Yu, Y.; Sun, L.-M.; Xing, J.-Q.; Li, T.; Zhu, Y.; Wang, M.; Yu, Y.; Xue, W.; Xia, T.; et al. GCG Inhibits SARS-CoV-2 Replication by Disrupting the Liquid Phase Condensation of Its Nucleocapsid Protein. Nat. Commun. 2021, 12, 2114. [Google Scholar] [CrossRef] [PubMed]
  76. Pflug, A.; Guilligay, D.; Reich, S.; Cusack, S. Structure of Influenza A Polymerase Bound to the Viral RNA Promoter. Nature 2014, 516, 355–360. [Google Scholar] [CrossRef] [PubMed]
  77. Tchatalbachev, S.; Flick, R.; Hobom, G. The Packaging Signal of Influenza Viral RNA Molecules. RNA 2001, 7, 979–989. [Google Scholar] [CrossRef] [Green Version]
  78. Chaimayo, C.; Hayashi, T.; Underwood, A.; Hodges, E.; Takimoto, T. Selective Incorporation of VRNP into Influenza A Virions Determined by Its Specific Interaction with M1 Protein. Virology 2017, 505, 23–32. [Google Scholar] [CrossRef]
  79. Brandt, S.; Blissenbach, M.; Grewe, B.; Konietzny, R.; Grunwald, T.; Uberla, K. Rev Proteins of Human and Simian Immunodeficiency Virus Enhance RNA Encapsidation. PLoS Pathog. 2007, 3, e54. [Google Scholar] [CrossRef] [PubMed]
  80. Blissenbach, M.; Grewe, B.; Hoffmann, B.; Brandt, S.; Uberla, K. Nuclear RNA Export and Packaging Functions of HIV-1 Rev Revisited. J. Virol. 2010, 84, 6598–6604. [Google Scholar] [CrossRef] [Green Version]
  81. Pocock, G.M.; Becker, J.T.; Swanson, C.M.; Ahlquist, P.; Sherer, N.M. HIV-1 and M-PMV RNA Nuclear Export Elements Program Viral Genomes for Distinct Cytoplasmic Trafficking Behaviors. PLoS Pathog. 2016, 12, e1005565. [Google Scholar] [CrossRef] [Green Version]
  82. Grewe, B.; Ehrhardt, K.; Hoffmann, B.; Blissenbach, M.; Brandt, S.; Uberla, K. The HIV-1 Rev Protein Enhances Encapsidation of Unspliced and Spliced, RRE-Containing Lentiviral Vector RNA. PLoS ONE 2012, 7, e48688. [Google Scholar] [CrossRef] [PubMed]
  83. Serganov, A.; Nudler, E. A Decade of Riboswitches. Cell 2013, 152, 17–24. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Lee, S.R.; Lykke-Andersen, J. Emerging Roles for Ribonucleoprotein Modification and Remodeling in Controlling RNA Fate. Trends Cell Biol. 2013, 23, 504–510. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Shetty, S.; Stefanovic, S.; Mihailescu, M.R. Hepatitis C Virus RNA: Molecular Switches Mediated by Long-Range RNA-RNA Interactions? Nucleic Acids Res. 2013, 41, 2526–2540. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Romero-López, C.; Barroso-Deljesus, A.; García-Sacristán, A.; Briones, C.; Berzal-Herranz, A. End-to-End Crosstalk within the Hepatitis C Virus Genome Mediates the Conformational Switch of the 3′X-Tail Region. Nucleic Acids Res. 2014, 42, 567–582. [Google Scholar] [CrossRef] [PubMed]
  87. Pirakitikulr, N.; Kohlway, A.; Lindenbach, B.D.; Pyle, A.M. The Coding Region of the HCV Genome Contains a Network of Regulatory RNA Structures. Mol. Cell 2016, 62, 111–120. [Google Scholar] [CrossRef] [Green Version]
  88. Friebe, P.; Boudet, J.; Simorre, J.-P.; Bartenschlager, R. Kissing-Loop Interaction in the 3′ End of the Hepatitis C Virus Genome Essential for RNA Replication. J. Virol. 2005, 79, 380–392. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  89. You, S.; Stump, D.D.; Branch, A.D.; Rice, C.M. A Cis-Acting Replication Element in the Sequence Encoding the NS5B RNA-Dependent RNA Polymerase Is Required for Hepatitis C Virus RNA Replication. J. Virol. 2004, 78, 1352–1366. [Google Scholar] [CrossRef] [Green Version]
  90. Lee, H.; Shin, H.; Wimmer, E.; Paul, A. V Cis-Acting RNA Signals in the NS5B C-Terminal Coding Sequence of the Hepatitis C Virus Genome. J. Virol. 2004, 78, 10865–10877. [Google Scholar] [CrossRef] [Green Version]
  91. Tuplin, A.; Struthers, M.; Simmonds, P.; Evans, D.J. A Twist in the Tail: SHAPE Mapping of Long-Range Interactions and Structural Rearrangements of RNA Elements Involved in HCV Replication. Nucleic Acids Res. 2012, 40, 6908–6921. [Google Scholar] [CrossRef] [Green Version]
  92. Shi, G.; Ando, T.; Suzuki, R.; Matsuda, M.; Nakashima, K.; Ito, M.; Omatsu, T.; Oba, M.; Ochiai, H.; Kato, T.; et al. Involvement of the 3′ Untranslated Region in Encapsidation of the Hepatitis C Virus. PLoS Pathog. 2016, 12, e1005441. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Romero-López, C.; Berzal-Herranz, A. The Functional RNA Domain 5BSL3.2 within the NS5B Coding Sequence Influences Hepatitis C Virus IRES-Mediated Translation. Cell. Mol. Life Sci. 2012, 69, 103–113. [Google Scholar] [CrossRef] [PubMed]
  94. Shetty, S.; Kim, S.; Shimakami, T.; Lemon, S.M.; Mihailescu, M.-R. Hepatitis C Virus Genomic RNA Dimerization Is Mediated via a Kissing Complex Intermediate. RNA 2010, 16, 913–925. [Google Scholar] [CrossRef] [Green Version]
  95. Ivanyi-Nagy, R.; Kanevsky, I.; Gabus, C.; Lavergne, J.-P.; Ficheux, D.; Penin, F.; Fossé, P.; Darlix, J.-L. Analysis of Hepatitis C Virus RNA Dimerization and Core-RNA Interactions. Nucleic Acids Res. 2006, 34, 2618–2633. [Google Scholar] [CrossRef]
  96. Huthoff, H.; Berkhout, B. Two Alternating Structures of the HIV-1 Leader RNA. RNA 2001, 7, 143–157. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Ooms, M.; Huthoff, H.; Russell, R.; Liang, C.; Berkhout, B. A Riboswitch Regulates RNA Dimerization and Packaging in Human Immunodeficiency Virus Type 1 Virions. J. Virol. 2004, 78, 10814–10819. [Google Scholar] [CrossRef] [Green Version]
  98. Brown, J.D.; Kharytonchyk, S.; Chaudry, I.; Iyer, A.S.; Carter, H.; Becker, G.; Desai, Y.; Glang, L.; Choi, S.H.; Singh, K.; et al. Structural Basis for Transcriptional Start Site Control of HIV-1 RNA Fate. Science 2020, 368, 413–417. [Google Scholar] [CrossRef]
  99. Abbink, T.E.M.; Berkhout, B. A Novel Long Distance Base-Pairing Interaction in Human Immunodeficiency Virus Type 1 RNA Occludes the Gag Start Codon. J. Biol. Chem. 2003, 278, 11601–11611. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  100. Keane, S.C.; Heng, X.; Lu, K.; Kharytonchyk, S.; Ramakrishnan, V.; Carter, G.; Barton, S.; Hosic, A.; Florwick, A.; Santos, J.; et al. RNA Structure. Structure of the HIV-1 RNA Packaging Signal. Science 2015, 348, 917–921. [Google Scholar] [CrossRef] [Green Version]
  101. Lu, K.; Heng, X.; Garyu, L.; Monti, S.; Garcia, E.L.; Kharytonchyk, S.; Dorjsuren, B.; Kulandaivel, G.; Jones, S.; Hiremath, A.; et al. NMR Detection of Structures in the HIV-1 5′-Leader RNA That Regulate Genome Packaging. Science 2011, 334, 242–245. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  102. Abbink, T.E.M.; Ooms, M.; Haasnoot, P.C.J.; Berkhout, B. The HIV-1 Leader RNA Conformational Switch Regulates RNA Dimerization but Does Not Regulate MRNA Translation †. Biochemistry 2005, 44, 9058–9066. [Google Scholar] [CrossRef]
  103. Brigham, B.S.; Kitzrow, J.P.; Reyes, J.-P.C.; Musier-Forsyth, K.; Munro, J.B. Intrinsic Conformational Dynamics of the HIV-1 Genomic RNA 5′UTR. Proc. Natl. Acad. Sci. USA 2019, 116, 10372–10381. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Jones, K.L.K.L.; Smyth, R.P.R.P.; Pereira, C.F.C.F.; Cameron, P.U.; Lewin, S.R.S.R.; Jaworowski, A.; Mak, J. Early Events of HIV-1 Infection: Can Signaling Be the Next Therapeutic Target? J. Neuroimmune Pharmacol. 2011, 6, 269–283. [Google Scholar] [CrossRef] [PubMed]
  105. Skripkin, E.; Paillart, J.C.; Marquet, R.; Ehresmann, B.; Ehresmann, C.; Lina, B.; Moulès, V.; Marquet, R. Identification of the Primary Site of the Human Immunodeficiency Virus Type 1 RNA Dimerization in Vitro. Proc. Natl. Acad. Sci. USA 1994, 91, 4945–4949. [Google Scholar] [CrossRef] [Green Version]
  106. Paillart, J.C.; Marquet, R.; Skripkin, E.; Ehresmann, B.; Ehresmann, C. Mutational Analysis of the Bipartite Dimer Linkage Structure of Human Immunodeficiency Virus Type 1 Genomic RNA. J. Biol. Chem. 1994, 269, 27486–27493. [Google Scholar] [CrossRef]
  107. Paillart, J.C.; Berthoux, L.; Ottmann, M.; Darlix, J.L.; Marquet, R.; Ehresmann, B.; Ehresmann, C. A Dual Role of the Putative RNA Dimerization Initiation Site of Human Immunodeficiency Virus Type 1 in Genomic RNA Packaging and Proviral DNA Synthesis. J. Virol. 1996, 70, 8348–8354. [Google Scholar] [CrossRef] [Green Version]
  108. Paillart, J.-C.; Shehu-Xhilaga, M.; Marquet, R.; Mak, J. Dimerization of Retroviral RNA Genomes: An Inseparable Pair. Nat. Rev. Microbiol. 2004, 2, 461–472. [Google Scholar] [CrossRef]
  109. Russell, R.S.; Liang, C.; Wainberg, M.A. Is HIV-1 RNA Dimerization a Prerequisite for Packaging? Yes, No, Probably? Retrovirology 2004, 1, 23. [Google Scholar] [CrossRef] [Green Version]
  110. Keane, S.C.; Van, V.; Frank, H.M.; Sciandra, C.A.; McCowin, S.; Santos, J.; Heng, X.; Summers, M.F. NMR Detection of Intermolecular Interaction Sites in the Dimeric 5′-Leader of the HIV-1 Genome. Proc. Natl. Acad. Sci. USA 2016, 113, 13033–13038. [Google Scholar] [CrossRef] [Green Version]
  111. Boyd, P.S.; Brown, J.B.; Brown, J.D.; Catazaro, J.; Chaudry, I.; Ding, P.; Dong, X.; Marchant, J.; O’Hern, C.T.; Singh, K.; et al. NMR Studies of Retroviral Genome Packaging. Viruses 2020, 12, 1115. [Google Scholar] [CrossRef] [PubMed]
  112. Kharytonchyk, S.; Monti, S.; Smaldino, P.J.; Van, V.; Bolden, N.C.; Brown, J.D.; Russo, E.; Swanson, C.; Shuey, A.; Telesnitsky, A.; et al. Transcriptional Start Site Heterogeneity Modulates the Structure and Function of the HIV-1 Genome. Proc. Natl. Acad. Sci. USA 2016, 113, 13378–13383. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Jopling, C.L.; Yi, M.; Lancaster, A.M.; Lemon, S.M.; Sarnow, P. Modulation of Hepatitis C Virus RNA Abundance by a Liver-Specific MicroRNA. Science 2005, 309, 1577–1581. [Google Scholar] [CrossRef] [Green Version]
  114. Luna, J.M.; Scheel, T.K.H.; Danino, T.; Shaw, K.S.; Mele, A.; Fak, J.J.; Nishiuchi, E.; Takacs, C.N.; Catanese, M.T.; de Jong, Y.P.; et al. Hepatitis C Virus RNA Functionally Sequesters MiR-122. Cell 2015, 160, 1099–1110. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Jopling, C.L.; Schütz, S.; Sarnow, P. Position-Dependent Function for a Tandem MicroRNA MiR-122-Binding Site Located in the Hepatitis C Virus RNA Genome. Cell Host Microbe 2008, 4, 77–85. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Li, Y.; Masaki, T.; Yamane, D.; McGivern, D.R.; Lemon, S.M. Competing and Noncompeting Activities of MiR-122 and the 5′ Exonuclease Xrn1 in Regulation of Hepatitis C Virus Replication. Proc. Natl. Acad. Sci. USA 2013, 110, 1881–1886. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Sedano, C.D.; Sarnow, P. Hepatitis C Virus Subverts Liver-Specific MiR-122 to Protect the Viral Genome from Exoribonuclease Xrn2. Cell Host Microbe 2014, 16, 257–264. [Google Scholar] [CrossRef] [Green Version]
  118. Shimakami, T.; Yamane, D.; Jangra, R.K.; Kempf, B.J.; Spaniel, C.; Barton, D.J.; Lemon, S.M. Stabilization of Hepatitis C Virus RNA by an Ago2-MiR-122 Complex. Proc. Natl. Acad. Sci. USA 2012, 109, 941–946. [Google Scholar] [CrossRef] [Green Version]
  119. Niepmann, M. Activation of Hepatitis C Virus Translation by a Liver-Specific MicroRNA. Cell Cycle 2009, 8, 1473–1477. [Google Scholar] [CrossRef] [Green Version]
  120. Roberts, A.P.E.; Lewis, A.P.; Jopling, C.L. MiR-122 Activates Hepatitis C Virus Translation by a Specialized Mechanism Requiring Particular RNA Components. Nucleic Acids Res. 2011, 39, 7716–7729. [Google Scholar] [CrossRef] [Green Version]
  121. Schult, P.; Roth, H.; Adams, R.L.; Mas, C.; Imbert, L.; Orlik, C.; Ruggieri, A.; Pyle, A.M.; Lohmann, V. MicroRNA-122 Amplifies Hepatitis C Virus Translation by Shaping the Structure of the Internal Ribosomal Entry Site. Nat. Commun. 2018, 9, 2613. [Google Scholar] [CrossRef]
  122. Henke, J.I.; Goergen, D.; Zheng, J.; Song, Y.; Schüttler, C.G.; Fehr, C.; Jünemann, C.; Niepmann, M. MicroRNA-122 Stimulates Translation of Hepatitis C Virus RNA. EMBO J. 2008, 27, 3300–3310. [Google Scholar] [CrossRef] [Green Version]
  123. Kunden, R.D.; Ghezelbash, S.; Khan, J.Q.; Wilson, J.A. Location Specific Annealing of MiR-122 and Other Small RNAs Defines an Hepatitis C Virus 5′ UTR Regulatory Element with Distinct Impacts on Virus Translation and Genome Stability. Nucleic Acids Res. 2020, 48, 9235–9249. [Google Scholar] [CrossRef] [PubMed]
  124. Masaki, T.; Arend, K.C.; Li, Y.; Yamane, D.; McGivern, D.R.; Kato, T.; Wakita, T.; Moorman, N.J.; Lemon, S.M. MiR-122 Stimulates Hepatitis C Virus RNA Synthesis by Altering the Balance of Viral RNAs Engaged in Replication versus Translation. Cell Host Microbe 2015, 17, 217–228. [Google Scholar] [CrossRef] [Green Version]
  125. Fukuhara, T.; Kambara, H.; Shiokawa, M.; Ono, C.; Katoh, H.; Morita, E.; Okuzaki, D.; Maehara, Y.; Koike, K.; Matsuura, Y. Expression of MicroRNA MiR-122 Facilitates an Efficient Replication in Nonhepatic Cells upon Infection with Hepatitis C Virus. J. Virol. 2012, 86, 7918–7933. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. Jiang, M.; Mak, J.; Wainberg, M.A.; Parniak, M.A.; Cohen, E.; Kleiman, L. Variable TRNA Content in HIV-1IIIB. Biochem. Biophys. Res. Commun. 1992, 185, 1005–1015. [Google Scholar] [CrossRef]
  127. Jiang, M.; Mak, J.; Ladha, A.; Cohen, E.; Klein, M.; Rovinski, B.; Kleiman, L. Identification of TRNAs Incorporated into Wild-Type and Mutant Human Immunodeficiency Virus Type 1. J. Virol. 1993, 67, 3246–3253. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  128. Seif, E.; Niu, M.; Kleiman, L. Annealing to Sequences within the Primer Binding Site Loop Promotes an HIV-1 RNA Conformation Favoring RNA Dimerization and Packaging. RNA 2013, 19, 1384–1393. [Google Scholar] [CrossRef] [Green Version]
  129. Wichgers Schreur, P.J.; Kortekaas, J. Single-Molecule FISH Reveals Non-Selective Packaging of Rift Valley Fever Virus Genome Segments. PLoS Pathog. 2016, 12, e1005800. [Google Scholar] [CrossRef]
  130. Bermúdez-Méndez, E.; Katrukha, E.A.; Spruit, C.M.; Kortekaas, J.; Wichgers Schreur, P.J. Visualizing the Ribonucleoprotein Content of Single Bunyavirus Virions Reveals More Efficient Genome Packaging in the Arthropod Host. Commun. Biol. 2021, 4, 345. [Google Scholar] [CrossRef] [PubMed]
  131. Hutchinson, E.C.; von Kirchbach, J.C.; Gog, J.R.; Digard, P. Genome Packaging in Influenza A Virus. J. Gen. Virol. 2010, 91, 313–328. [Google Scholar] [CrossRef]
  132. Chou, Y.-Y.; Vafabakhsh, R.; Doganay, S.; Gao, Q.; Ha, T.; Palese, P.; Doğanay, S.; Gao, Q.; Ha, T.; Palese, P. One Influenza Virus Particle Packages Eight Unique Viral RNAs as Shown by FISH Analysis. Proc. Natl. Acad. Sci. USA 2012, 109, 9101–9106. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. Haralampiev, I.; Prisner, S.; Nitzan, M.; Schade, M.; Jolmes, F.; Schreiber, M.; Loidolt-Krüger, M.; Jongen, K.; Chamiolo, J.; Nilson, N.; et al. Selective Flexible Packaging Pathways of the Segmented Genome of Influenza A Virus. Nat. Commun. 2020, 11, 1–13. [Google Scholar] [CrossRef] [PubMed]
  134. Noda, T.; Sagara, H.; Yen, A.; Takada, A.; Kida, H.; Cheng, R.H.; Kawaoka, Y. Architecture of Ribonucleoprotein Complexes in Influenza A Virus Particles. Nature 2006, 439, 490–492. [Google Scholar] [CrossRef]
  135. Noda, T.; Murakami, S.; Nakatsu, S.; Imai, H.; Muramoto, Y.; Shindo, K.; Sagara, H.; Kawaoka, Y. Importance of the 1 + 7 Configuration of Ribonucleoprotein Complexes for Influenza A Virus Genome Packaging. Nat. Commun. 2018, 9, 54. [Google Scholar] [CrossRef]
  136. Fournier, E.; Moules, V.; Essere, B.; Paillart, J.-C.; Sirbat, J.-D.; Isel, C.; Cavalier, A.; Rolland, J.-P.; Thomas, D.; Lina, B.; et al. A Supramolecular Assembly Formed by Influenza A Virus Genomic RNA Segments. Nucleic Acids Res. 2012, 40, 2197–2209. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  137. Duhaut, S.D.; McCauley, J.W. Defective RNAs Inhibit the Assembly of Influenza Virus Genome Segments in a Segment-Specific Manner. Virology 1996, 216, 326–337. [Google Scholar] [CrossRef]
  138. Duhaut, S.D.; Dimmock, N.J. Heterologous Protection of Mice from a Lethal Human H1N1 Influenza A Virus Infection by H3N8 Equine Defective Interfering Virus: Comparison of Defective RNA Sequences Isolated from the DI Inoculum and Mouse Lung. Virology 1998, 248, 241–253. [Google Scholar] [CrossRef] [Green Version]
  139. Jennings, P.A.; Finch, J.T.; Winter, G.; Robertson, J.S. Does the Higher Order Structure of the Influenza Virus Ribonucleoprotein Guide Sequence Rearrangements in Influenza Viral RNA? Cell 1983, 34, 619–627. [Google Scholar] [CrossRef]
  140. Noble, S.; Dimmock, N.J. Characterization of Putative Defective Interfering (DI) A/WSN RNAs Isolated from the Lungs of Mice Protected from an Otherwise Lethal Respiratory Infection with Influenza Virus A/WSN (H1N1): A Subset of the Inoculum DI RNAs. Virology 1995, 210, 9–19. [Google Scholar] [CrossRef] [Green Version]
  141. Dos Santos Afonso, E.; Escriou, N.; Leclercq, I.; van der Werf, S.; Naffakh, N. The Generation of Recombinant Influenza A Viruses Expressing a PB2 Fusion Protein Requires the Conservation of a Packaging Signal Overlapping the Coding and Noncoding Regions at the 5′ End of the PB2 Segment. Virology 2005, 341, 34–46. [Google Scholar] [CrossRef] [Green Version]
  142. Liang, Y.; Hong, Y.; Parslow, T.G. Cis-Acting Packaging Signals in the Influenza Virus PB1, PB2, and PA Genomic RNA Segments. J. Virol. 2005, 79, 10348–10355. [Google Scholar] [CrossRef] [Green Version]
  143. Liang, Y.; Huang, T.; Ly, H.; Parslow, T.G.; Liang, Y. Mutational Analyses of Packaging Signals in Influenza Virus PA, PB1, and PB2 Genomic RNA Segments. J. Virol. 2008, 82, 229–236. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Fujii, K.; Fujii, Y.; Noda, T.; Muramoto, Y.; Watanabe, T.; Takada, A.; Goto, H.; Horimoto, T.; Kawaoka, Y. Importance of Both the Coding and the Segment-Specific Noncoding Regions of the Influenza A Virus NS Segment for Its Efficient Incorporation into Virions. J. Virol. 2005, 79, 3766–3774. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Fujii, Y.; Goto, H.; Watanabe, T.; Yoshida, T.; Kawaoka, Y. Selective Incorporation of Influenza Virus RNA Segments into Virions. Proc. Natl. Acad. Sci. USA 2003, 100, 2002–2007. [Google Scholar] [CrossRef] [Green Version]
  146. Marsh, G.A.; Hatami, R.; Palese, P. Specific Residues of the Influenza A Virus Hemagglutinin Viral RNA Are Important for Efficient Packaging into Budding Virions. J. Virol. 2007, 81, 9727–9736. [Google Scholar] [CrossRef] [Green Version]
  147. Ozawa, M.; Maeda, J.; Iwatsuki-Horimoto, K.; Watanabe, S.; Goto, H.; Horimoto, T.; Kawaoka, Y. Nucleotide Sequence Requirements at the 5′ End of the Influenza A Virus M RNA Segment for Efficient Virus Replication. J. Virol. 2009, 83, 3384–3388. [Google Scholar] [CrossRef] [Green Version]
  148. Ozawa, M.; Fujii, K.; Muramoto, Y.; Yamada, S.; Yamayoshi, S.; Takada, A.; Goto, H.; Horimoto, T.; Kawaoka, Y. Contributions of Two Nuclear Localization Signals of Influenza A Virus Nucleoprotein to Viral Replication. J. Virol. 2007, 81, 30–41. [Google Scholar] [CrossRef] [Green Version]
  149. Hutchinson, E.C.; Wise, H.M.; Kudryavtseva, K.; Curran, M.D.; Digard, P. Characterisation of Influenza A Viruses with Mutations in Segment 5 Packaging Signals. Vaccine 2009, 27, 6270–6275. [Google Scholar] [CrossRef] [Green Version]
  150. Goto, H.; Muramoto, Y.; Noda, T.; Kawaoka, Y. The Genome-Packaging Signal of the Influenza A Virus Genome Comprises a Genome Incorporation Signal and a Genome-Bundling Signal. J. Virol. 2013, 87, 11316–11322. [Google Scholar] [CrossRef] [Green Version]
  151. Noda, T.; Sugita, Y.; Aoyama, K.; Hirase, A.; Kawakami, E.; Miyazawa, A.; Sagara, H.; Kawaoka, Y. Three-Dimensional Analysis of Ribonucleoprotein Complexes in Influenza A Virus. Nat. Commun. 2012, 3, 639. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  152. Fournier, E.; Moules, V.; Essere, B.; Paillart, J.C.; Sirbat, J.D.; Cavalier, A.; Rolland, J.P.; Thomas, D.; Lina, B.; Isel, C.; et al. Interaction Network Linking the Human H3N2 Influenza A Virus Genomic RNA Segments. Vaccine 2012, 30, 7359–7367. [Google Scholar] [CrossRef] [PubMed]
  153. Gavazzi, C.; Yver, M.; Isel, C.; Smyth, R.P.R.P.; Rosa-Calatrava, M.; Lina, B.; Moulès, V.; Marquet, R. A Functional Sequence-Specific Interaction between Influenza A Virus Genomic RNA Segments. Proc. Natl. Acad. Sci. USA 2013, 110, 16604–16609. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  154. Gavazzi, C.; Isel, C.; Fournier, E.; Moules, V.; Cavalier, A.; Thomas, D.; Lina, B.; Marquet, R. An in Vitro Network of Intermolecular Interactions between Viral RNA Segments of an Avian H5N2 Influenza A Virus: Comparison with a Human H3N2 Virus. Nucleic Acids Res. 2013, 41, 1241–1254. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  155. Hutchinson, E.C.; Curran, M.D.; Read, E.K.; Gog, J.R.; Digard, P. Mutational Analysis of Cis-Acting RNA Signals in Segment 7 of Influenza A Virus. J. Virol. 2008, 82, 11869–11879. [Google Scholar] [CrossRef] [Green Version]
  156. Gilbertson, B.; Zheng, T.; Gerber, M.; Printz-Schweigert, A.; Ong, C.; Marquet, R.; Isel, C.; Rockman, S.; Brown, L. Influenza NA and PB1 Gene Segments Interact during the Formation of Viral Progeny: Localization of the Binding Region within the PB1 Gene. Viruses 2016, 8, 238. [Google Scholar] [CrossRef]
  157. Dadonaite, B.; Gilbertson, B.; Knight, M.L.; Trifkovic, S.; Rockman, S.; Laederach, A.; Brown, L.E.; Fodor, E.; Bauer, D.L.V. The Structure of the Influenza A Virus Genome. Nat. Microbiol. 2019, 4, 1781–1789. [Google Scholar] [CrossRef]
  158. Le Sage, V.; Kanarek, J.P.; Snyder, D.J.; Cooper, V.S.; Lakdawala, S.S.; Lee, N. Mapping of Influenza Virus RNA-RNA Interactions Reveals a Flexible Network. Cell Rep. 2020, 31, 107823. [Google Scholar] [CrossRef]
  159. Lu, Z.; Chang, H.Y. The RNA Base-Pairing Problem and Base-Pairing Solutions. Cold Spring Harb. Perspect. Biol. 2018, 10, a034926. [Google Scholar] [CrossRef] [PubMed]
  160. Fuchs, R.T.; Sun, Z.; Zhuang, F.; Robb, G.B. Bias in Ligation-Based Small RNA Sequencing Library Construction Is Determined by Adaptor and RNA Structure. PLoS ONE 2015, 10, e0126049. [Google Scholar] [CrossRef] [Green Version]
  161. Bolte, H.; Rosu, M.E.; Hagelauer, E.; García-Sastre, A.; Schwemmle, M. Packaging of the Influenza A Virus Genome Is Governed by a Plastic Network of RNA/Protein Interactions. J. Virol. 2018, 93, e01861-18. [Google Scholar] [CrossRef] [Green Version]
  162. Williams, G.D.; Townsend, D.; Wylie, K.M.; Kim, P.J.; Amarasinghe, G.K.; Kutluay, S.B.; Boon, A.C.M. Nucleotide Resolution Mapping of Influenza A Virus Nucleoprotein-RNA Interactions Reveals RNA Features Required for Replication. Nat. Commun. 2018, 9, 465. [Google Scholar] [CrossRef] [Green Version]
  163. Moreira, É.A.; Weber, A.; Bolte, H.; Kolesnikova, L.; Giese, S.; Lakdawala, S.; Beer, M.; Zimmer, G.; García-Sastre, A.; Schwemmle, M.; et al. A Conserved Influenza A Virus Nucleoprotein Code Controls Specific Viral Genome Packaging. Nat. Commun. 2016, 7, 12861. [Google Scholar] [CrossRef] [PubMed]
  164. Jayaram, H.; Estes, M.K.; Prasad, B.V.V. Emerging Themes in Rotavirus Cell Entry, Genome Organization, Transcription and Replication. Virus Res. 2004, 101, 67–81. [Google Scholar] [CrossRef]
  165. Fajardo, T.; Sung, P.-Y.; Celma, C.C.; Roy, P. Rotavirus Genomic RNA Complex Forms via Specific RNA-RNA Interactions: Disruption of RNA Complex Inhibits Virus Infectivity. Viruses 2017, 9, 167. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  166. Fajardo, T.; Sung, P.-Y.; Roy, P. Disruption of Specific RNA-RNA Interactions in a Double-Stranded RNA Virus Inhibits Genome Packaging and Virus Infectivity. PLoS Pathog. 2015, 11, e1005321. [Google Scholar] [CrossRef]
  167. Sung, P.-Y.; Roy, P. Sequential Packaging of RNA Genomic Segments during the Assembly of Bluetongue Virus. Nucleic Acids Res. 2014, 42, 13824–13838. [Google Scholar] [CrossRef] [Green Version]
  168. AlShaikhahmed, K.; Leonov, G.; Sung, P.-Y.; Bingham, R.J.; Twarock, R.; Roy, P. Dynamic Network Approach for the Modelling of Genomic Sub-Complexes in Multi-Segmented Viruses. Nucleic Acids Res. 2018, 46, 12087–12098. [Google Scholar] [CrossRef] [PubMed]
  169. Lakdawala, S.S.; Wu, Y.; Wawrzusin, P.; Kabat, J.; Broadbent, A.J.; Lamirande, E.W.; Fodor, E.; Altan-Bonnet, N.; Shroff, H.; Subbarao, K. Influenza a Virus Assembly Intermediates Fuse in the Cytoplasm. PLoS Pathog. 2014, 10, e1003971. [Google Scholar] [CrossRef]
  170. Chou, Y.; Heaton, N.S.; Gao, Q.; Palese, P.; Singer, R.H.; Singer, R.; Lionnet, T. Colocalization of Different Influenza Viral RNA Segments in the Cytoplasm before Viral Budding as Shown by Single-Molecule Sensitivity FISH Analysis. PLoS Pathog. 2013, 9, e1003358. [Google Scholar] [CrossRef]
  171. Payne, J.L.; Wagner, A. The Causes of Evolvability and Their Evolution. Nat. Rev. Genet. 2019, 20, 24–38. [Google Scholar] [CrossRef]
  172. Smyth, R.P.; Negroni, M. A Step Forward Understanding HIV-1 Diversity. Retrovirology 2016, 13, 27. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  173. McDonald, S.M.; Nelson, M.I.; Turner, P.E.; Patton, J.T. Reassortment in Segmented RNA Viruses: Mechanisms and Outcomes. Nat. Rev. Microbiol. 2016, 14, 448–460. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  174. Smyth, R.P.; Schlub, T.E.; Grimm, A.J.; Waugh, C.; Ellenberg, P.; Chopra, A.; Mallal, S.; Cromer, D.; Mak, J.; Davenport, M.P. Identifying Recombination Hot Spots in the HIV-1 Genome. J. Virol. 2014, 88, 2891–2902. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  175. Schlub, T.E.; Smyth, R.P.; Grimm, A.J.; Mak, J.; Davenport, M.P. Accurately Measuring Recombination between Closely Related HIV-1 Genomes. PLoS Comput. Biol. 2010, 6, e1000766. [Google Scholar] [CrossRef]
  176. Cromer, D.; Schlub, T.E.; Smyth, R.P.; Grimm, A.J.; Chopra, A.; Mallal, S.; Davenport, M.P.; Mak, J. HIV-1 Mutation and Recombination Rates Are Different in Macrophages and T-Cells. Viruses 2016, 8, 118. [Google Scholar] [CrossRef] [Green Version]
  177. Schlub, T.E.; Grimm, A.J.; Smyth, R.P.; Cromer, D.; Chopra, A.; Mallal, S.; Venturi, V.; Waugh, C.; Mak, J.; Davenport, M.P. Fifteen to Twenty Percent of HIV Substitution Mutations Are Associated with Recombination. J. Virol. 2014, 88, 3837–3849. [Google Scholar] [CrossRef] [Green Version]
  178. Levy, D.N.; Aldrovandi, G.M.; Kutsch, O.; Shaw, G.M. Dynamics of HIV-1 Recombination in Its Natural Target Cells. Proc. Natl. Acad. Sci. USA 2004, 101, 4204–4209. [Google Scholar] [CrossRef] [Green Version]
  179. Chin, M.P.S.; Rhodes, T.D.; Chen, J.; Fu, W.; Hu, W.-S. Identification of a Major Restriction in HIV-1 Intersubtype Recombination. Proc. Natl. Acad. Sci. USA 2005, 102, 9002–9007. [Google Scholar] [CrossRef] [Green Version]
  180. Paillart, J.C.; Skripkin, E.; Ehresmann, B.; Ehresmann, C.; Marquet, R. A Loop-Loop “Kissing” Complex Is the Essential Part of the Dimer Linkage of Genomic HIV-1 RNA. Proc. Natl. Acad. Sci. USA 1996, 93, 5572–5577. [Google Scholar] [CrossRef] [Green Version]
  181. Hill, M.K.; Shehu-Xhilaga, M.; Campbell, S.M.; Poumbourios, P.; Crowe, S.M.; Mak, J. The Dimer Initiation Sequence Stem-Loop of Human Immunodeficiency Virus Type 1 Is Dispensable for Viral Replication in Peripheral Blood Mononuclear Cells. J. Virol. 2003, 77, 8329–8335. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  182. Jones, K.L.; Sonza, S.; Mak, J. Primary T-Lymphocytes Rescue the Replication of HIV-1 DIS RNA Mutants in Part by Facilitating Reverse Transcription. Nucleic Acids Res. 2008, 36, 1578–1588. [Google Scholar] [CrossRef] [Green Version]
  183. Cobbin, J.C.; Ong, C.; Verity, E.; Gilbertson, B.P.; Rockman, S.P.; Brown, L.E. Influenza Virus PB1 and Neuraminidase Gene Segments Can Cosegregate during Vaccine Reassortment Driven by Interactions in the PB1 Coding Region. J. Virol. 2014, 88, 8971–8980. [Google Scholar] [CrossRef] [Green Version]
  184. Essere, B.; Yver, M.; Gavazzi, C.; Terrier, O.; Isel, C.; Fournier, E.; Giroux, F.; Textoris, J.; Julien, T.; Socratous, C.; et al. Critical Role of Segment-Specific Packaging Signals in Genetic Reassortment of Influenza A Viruses. Proc. Natl. Acad. Sci. USA 2013, 110, E3840–E3848. [Google Scholar] [CrossRef] [Green Version]
  185. White, M.C.; Tao, H.; Steel, J.; Lowen, A.C. H5N8 and H7N9 Packaging Signals Constrain HA Reassortment with a Seasonal H3N2 Influenza A Virus. Proc. Natl. Acad. Sci. USA 2019. [Google Scholar] [CrossRef] [Green Version]
  186. Gao, Q.; Palese, P. Rewiring the RNAs of Influenza Virus to Prevent Reassortment. Proc. Natl. Acad. Sci. USA 2009, 106, 15891–15896. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  187. Nelson, M.I.; Holmes, E.C. The Evolution of Epidemic Influenza. Nat. Rev. Genet. 2007, 8, 196–205. [Google Scholar] [CrossRef] [PubMed]
  188. Rangan, R.; Watkins, A.M.; Chacon, J.; Kretsch, R.; Kladwang, W.; Zheludev, I.N.; Townley, J.; Rynge, M.; Thain, G.; Das, R. De Novo 3D Models of SARS-CoV-2 RNA Elements from Consensus Experimental Secondary Structures. Nucleic Acids Res. 2021, 49, 3092–3108. [Google Scholar] [CrossRef] [PubMed]
  189. Bhatt, P.R.; Scaiola, A.; Loughran, G.; Leibundgut, M.; Kratzel, A.; Meurs, R.; Dreos, R.; O’Connor, K.M.; McMillan, A.; Bode, J.W.; et al. Structural Basis of Ribosomal Frameshifting during Translation of the SARS-CoV-2 RNA Genome. Science 2021, 372, 1306–1313. [Google Scholar] [CrossRef]
  190. Kappel, K.; Zhang, K.; Su, Z.; Watkins, A.M.; Kladwang, W.; Li, S.; Pintilie, G.; Topkar, V.V.; Rangan, R.; Zheludev, I.N.; et al. Accelerated Cryo-EM-Guided Determination of Three-Dimensional RNA-Only Structures. Nat. Methods 2020, 17, 699–707. [Google Scholar] [CrossRef]
  191. Tihova, M.; Dryden, K.A.; Le, T.L.; Harvey, S.C.; Johnson, J.E.; Yeager, M.; Schneemann, A. Nodavirus Coat Protein Imposes Dodecahedral RNA Structure Independent of Nucleotide Sequence and Length. J. Virol. 2004, 78, 2897–2905. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  192. Hesketh, E.L.; Meshcheriakova, Y.; Dent, K.C.; Saxena, P.; Thompson, R.F.; Cockburn, J.J.; Lomonossoff, G.P.; Ranson, N.A. Mechanisms of Assembly and Genome Packaging in an RNA Virus Revealed by High-Resolution Cryo-EM. Nat. Commun. 2015, 6, 10113. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  193. Beren, C.; Cui, Y.; Chakravarty, A.; Yang, X.; Rao, A.L.N.; Knobler, C.M.; Zhou, Z.H.; Gelbart, W.M. Genome Organization and Interaction with Capsid Protein in a Multipartite RNA Virus. Proc. Natl. Acad. Sci. USA 2020, 117, 10673–10680. [Google Scholar] [CrossRef] [PubMed]
  194. Chandler-Bostock, R.; Mata, C.P.; Bingham, R.J.; Dykeman, E.C.; Meng, B.; Tuthill, T.J.; Rowlands, D.J.; Ranson, N.A.; Twarock, R.; Stockley, P.G. Assembly of Infectious Enteroviruses Depends on Multiple, Conserved Genomic RNA-Coat Protein Contacts. PLoS Pathog. 2020, 16, e1009146. [Google Scholar] [CrossRef]
  195. San Emeterio, J.; Pollack, L. Visualizing a Viral Genome with Contrast Variation Small Angle X-Ray Scattering. J. Biol. Chem. 2020, 295, 15923–15932. [Google Scholar] [CrossRef]
  196. Lin, T.; Cavarelli, J.; Johnson, J.E. Evidence for Assembly-Dependent Folding of Protein and RNA in an Icosahedral Virus. Virology 2003, 314, 26–33. [Google Scholar] [CrossRef] [Green Version]
  197. Morandi, E.; Manfredonia, I.; Simon, L.M.; Anselmi, F.; van Hemert, M.J.; Oliviero, S.; Incarnato, D. Genome-Scale Deconvolution of RNA Structure Ensembles. Nat. Methods 2021, 18, 249–252. [Google Scholar] [CrossRef] [PubMed]
  198. Tomezsko, P.J.; Corbin, V.D.A.; Gupta, P.; Swaminathan, H.; Glasgow, M.; Persad, S.; Edwards, M.D.; Mcintosh, L.; Papenfuss, A.T.; Emery, A.; et al. Determination of RNA Structural Diversity and Its Role in HIV-1 RNA Splicing. Nature 2020, 582, 438–442. [Google Scholar] [CrossRef]
  199. Schmidt, N.; Lareau, C.A.; Keshishian, H.; Ganskih, S.; Schneider, C.; Hennig, T.; Melanson, R.; Werner, S.; Wei, Y.; Zimmer, M.; et al. The SARS-CoV-2 RNA-Protein Interactome in Infected Human Cells. Nat. Microbiol. 2021, 6, 339–353. [Google Scholar] [CrossRef]
  200. Weidmann, C.A.; Mustoe, A.M.; Jariwala, P.B.; Calabrese, J.M.; Weeks, K.M. Analysis of RNA-Protein Networks with RNP-MaP Defines Functional Hubs on RNA. Nat. Biotechnol. 2020, 39, 347–356. [Google Scholar] [CrossRef] [PubMed]
  201. Gao, R.; Yu, C.-C.J.; Gao, L.; Piatkevich, K.D.; Neve, R.L.; Munro, J.B.; Upadhyayula, S.; Boyden, E.S. A Highly Homogeneous Polymer Composed of Tetrahedron-like Monomers for High-Isotropy Expansion Microscopy. Nat. Nanotechnol. 2021, 16, 698–707. [Google Scholar] [CrossRef]
  202. Ferrer, M.; Clerté, C.; Chamontin, C.; Basyuk, E.; Lainé, S.; Hottin, J.; Bertrand, E.; Margeat, E.; Mougel, M. Imaging HIV-1 RNA Dimerization in Cells by Multicolor Super-Resolution and Fluctuation Microscopies. Nucleic Acids Res. 2016, 44, 7922–7934. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  203. Vahey, M.D.; Fletcher, D.A. Low-Fidelity Assembly of Influenza A Virus Promotes Escape from Host Cells. Cell 2020, 180, 205. [Google Scholar] [CrossRef] [PubMed]
  204. Sardo, L.; Hatch, S.C.; Chen, J.; Nikolaitchik, O.; Burdick, R.C.; Chen, D.; Westlake, C.J.; Lockett, S.; Pathak, V.K.; Hu, W.-S. Dynamics of HIV-1 RNA Near the Plasma Membrane during Virus Assembly. J. Virol. 2015, 89, 10832–10840. [Google Scholar] [CrossRef] [Green Version]
  205. Jouvenet, N.; Bieniasz, P.D.; Simon, S.M. Imaging the Biogenesis of Individual HIV-1 Virions in Live Cells. Nature 2008, 454, 236–240. [Google Scholar] [CrossRef] [PubMed]
  206. Pénzes, Z.; Wroe, C.; Brown, T.D.; Britton, P.; Cavanagh, D. Replication and Packaging of Coronavirus Infectious Bronchitis Virus Defective RNAs Lacking a Long Open Reading Frame. J. Virol. 1996, 70, 8660–8668. [Google Scholar] [CrossRef] [Green Version]
  207. Li, Q.; Tong, Y.; Xu, Y.; Niu, J.; Zhong, J. Genetic Analysis of Serum-Derived Defective Hepatitis C Virus Genomes Revealed Novel Viral Cis Elements for Virus Replication and Assembly. J. Virol. 2018, 92. [Google Scholar] [CrossRef] [Green Version]
  208. Fosmire, J.A.; Hwang, K.; Makino, S. Identification and Characterization of a Coronavirus Packaging Signal. J. Virol. 1992, 66, 3522–3530. [Google Scholar] [CrossRef] [Green Version]
  209. van der Most, R.G.; Bredenbeek, P.J.; Spaan, W.J. A Domain at the 3′ End of the Polymerase Gene Is Essential for Encapsidation of CoronSavirus Defective Interfering RNAs. J. Virol. 1991, 65, 3219–3226. [Google Scholar] [CrossRef] [Green Version]
  210. Makino, S.; Yokomori, K.; Lai, M.M. Analysis of Efficiently Packaged Defective Interfering RNAs of Murine Coronavirus: Localization of a Possible RNA-Packaging Signal. J. Virol. 1990, 64, 6045–6053. [Google Scholar] [CrossRef] [Green Version]
Figure 1. RNA packaging signals (a) MS2 bacteriophage encodes a coat protein (CP) that binds to a 19-nucleotide stem loop structure, known as TR. Secondary structure model of the TR stem loop and three-dimensional structure of the coat protein-TR interaction; (PDB: 1AQ3) (b) MS2 bacteriophage genome is encapsidated through cooperative interactions between coat protein that are bound to a high affinity binding site (TR) and multiple low affinity binding sites encoded throughout the genome; (c) HIV-1 Gag recognizes the genomic RNA through a GU-rich high affinity binding site in the 5′ untranslated region (5′UTR). During assembly at the plasma membrane Gag switches its specificity from GU-rich sequences to A-rich sequences, which is proposed to favour assembly of Gag on its cognate genomic RNA.
Figure 1. RNA packaging signals (a) MS2 bacteriophage encodes a coat protein (CP) that binds to a 19-nucleotide stem loop structure, known as TR. Secondary structure model of the TR stem loop and three-dimensional structure of the coat protein-TR interaction; (PDB: 1AQ3) (b) MS2 bacteriophage genome is encapsidated through cooperative interactions between coat protein that are bound to a high affinity binding site (TR) and multiple low affinity binding sites encoded throughout the genome; (c) HIV-1 Gag recognizes the genomic RNA through a GU-rich high affinity binding site in the 5′ untranslated region (5′UTR). During assembly at the plasma membrane Gag switches its specificity from GU-rich sequences to A-rich sequences, which is proposed to favour assembly of Gag on its cognate genomic RNA.
Viruses 13 01788 g001
Figure 2. Selection of genomic RNA from non-genomic viral RNA. Dotted lines signify splicing or production of sub-genomic RNA. (a) Non-genomic RNA cannot be packaged because it does not contain the high affinity binding site for the (nucleo)capsid protein; (b) Non-genomic RNA contains the sequences for high affinity binding site, but it is not presented for binding. In HIV-1, negative regulatory elements (−ve) upstream of the splice donor (SD) conceal the Gag binding site unless counteracted by positive regulatory (+ve) sequences downstream of the SD.
Figure 2. Selection of genomic RNA from non-genomic viral RNA. Dotted lines signify splicing or production of sub-genomic RNA. (a) Non-genomic RNA cannot be packaged because it does not contain the high affinity binding site for the (nucleo)capsid protein; (b) Non-genomic RNA contains the sequences for high affinity binding site, but it is not presented for binding. In HIV-1, negative regulatory elements (−ve) upstream of the splice donor (SD) conceal the Gag binding site unless counteracted by positive regulatory (+ve) sequences downstream of the SD.
Viruses 13 01788 g002
Figure 3. Successful genome packaging requires RNA signals to direct the genome from sites of replication to sites of assembly. (a) Sites of assembly generated by liquid-liquid phase separation (LLPS). Specific interactions between the SARS Coronavirus-2 (SARS-CoV-2) nucleoprotein (N) and its genome induce LLPS. Sub-genomic RNA is excluded. (b) Influenza complementary RNA (cRNA) replication intermediates are not correctly exported from the nucleus and are therefore not packaged. (c) HIV-1 genomic RNA contains the Rev Response Element (RRE) which binds to the viral Rev protein needed to export the RNA from the nucleus. Rev binding is proposed to enhance packaging by transporting RNA to the correct sub-cellular location.
Figure 3. Successful genome packaging requires RNA signals to direct the genome from sites of replication to sites of assembly. (a) Sites of assembly generated by liquid-liquid phase separation (LLPS). Specific interactions between the SARS Coronavirus-2 (SARS-CoV-2) nucleoprotein (N) and its genome induce LLPS. Sub-genomic RNA is excluded. (b) Influenza complementary RNA (cRNA) replication intermediates are not correctly exported from the nucleus and are therefore not packaged. (c) HIV-1 genomic RNA contains the Rev Response Element (RRE) which binds to the viral Rev protein needed to export the RNA from the nucleus. Rev binding is proposed to enhance packaging by transporting RNA to the correct sub-cellular location.
Viruses 13 01788 g003
Figure 4. RNA structural switches and long-distance interactions regulate the balance between genome replication, packaging, and translation. (a) The HCV life cycle is regulated by a complex network of long-distance intra-molecular interactions and inter-molecular interactions. The packaging site which binds to the HCV core protein resides in the 3′ untranslated region (3′UTR). A long-distance base pairing between the cis-acting element (CRE) in the coding region and the X-tail in the 3′UTR regulates the balance between replication and packaging (light green dotted line). A long-distance interaction between the CRE and the internal ribosome entry site (IRES) regulates the balance between replication and translation (dark green dotted line). The 3′UTR is alternatively structured, leading to the formation of homodimers through an intermolecular interaction. The HCV 5′UTR binds the host microRNA miR-122 to regulate different aspects of HCV replication; (b) A structural switch in the HIV-1 5′UTR regulates the balance between genome translation and packaging. Transcripts beginning with three G residues fold into a monomer conformation and are preferentially translated. Transcripts beginning with one G residue fold into a dimer conformation. Structural switching is mediated by mutually exclusion interactions between regions U5 (pink), SL1 (blue), the AUG region (green), and a host tRNA (purple).
Figure 4. RNA structural switches and long-distance interactions regulate the balance between genome replication, packaging, and translation. (a) The HCV life cycle is regulated by a complex network of long-distance intra-molecular interactions and inter-molecular interactions. The packaging site which binds to the HCV core protein resides in the 3′ untranslated region (3′UTR). A long-distance base pairing between the cis-acting element (CRE) in the coding region and the X-tail in the 3′UTR regulates the balance between replication and packaging (light green dotted line). A long-distance interaction between the CRE and the internal ribosome entry site (IRES) regulates the balance between replication and translation (dark green dotted line). The 3′UTR is alternatively structured, leading to the formation of homodimers through an intermolecular interaction. The HCV 5′UTR binds the host microRNA miR-122 to regulate different aspects of HCV replication; (b) A structural switch in the HIV-1 5′UTR regulates the balance between genome translation and packaging. Transcripts beginning with three G residues fold into a monomer conformation and are preferentially translated. Transcripts beginning with one G residue fold into a dimer conformation. Structural switching is mediated by mutually exclusion interactions between regions U5 (pink), SL1 (blue), the AUG region (green), and a host tRNA (purple).
Viruses 13 01788 g004
Figure 5. Segmented viruses package their genomes randomly or by using segment specific packaging signals. (a) Bunyavirus randomly package three genome segments, small (S), medium (M) and large (L), such that many progeny virions are empty or incomplete; (b) Influenza A virus package 8 genome segments selectively. Genome segments within budding virions are organized into a 1 + 7 arrangement with one central segment surrounded by seven others; (c) Influenza vRNAs are packaged as viral ribonucleoproteins (vRNPs). vRNAs are bound by nucleoprotein (NP) and a heterotrimeric polymerase. vRNA is incompletely coated by NP allowing for inter-segment vRNA-vRNA interactions to occur as a possible mechanism underlying the selective packaging process.
Figure 5. Segmented viruses package their genomes randomly or by using segment specific packaging signals. (a) Bunyavirus randomly package three genome segments, small (S), medium (M) and large (L), such that many progeny virions are empty or incomplete; (b) Influenza A virus package 8 genome segments selectively. Genome segments within budding virions are organized into a 1 + 7 arrangement with one central segment surrounded by seven others; (c) Influenza vRNAs are packaged as viral ribonucleoproteins (vRNPs). vRNAs are bound by nucleoprotein (NP) and a heterotrimeric polymerase. vRNA is incompletely coated by NP allowing for inter-segment vRNA-vRNA interactions to occur as a possible mechanism underlying the selective packaging process.
Viruses 13 01788 g005
Figure 6. Viral RNA packaging influences viral evolution. (a) HIV-1 virions contain two copies of the genome. After co-infection, up to 50% of progeny co-package different genomes. Template switching during reverse transcription produce cDNA that is a chimera of the two genomes; (b) Influenza A virions contain eight different vRNA segments. After co-infection, reassortant progeny virions are produced containing a mixture of segments from each parental virus.
Figure 6. Viral RNA packaging influences viral evolution. (a) HIV-1 virions contain two copies of the genome. After co-infection, up to 50% of progeny co-package different genomes. Template switching during reverse transcription produce cDNA that is a chimera of the two genomes; (b) Influenza A virions contain eight different vRNA segments. After co-infection, reassortant progeny virions are produced containing a mixture of segments from each parental virus.
Viruses 13 01788 g006
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Ye, L.; Ambi, U.B.; Olguin-Nava, M.; Gribling-Burrer, A.-S.; Ahmad, S.; Bohn, P.; Weber, M.M.; Smyth, R.P. RNA Structures and Their Role in Selective Genome Packaging. Viruses 2021, 13, 1788. https://doi.org/10.3390/v13091788

AMA Style

Ye L, Ambi UB, Olguin-Nava M, Gribling-Burrer A-S, Ahmad S, Bohn P, Weber MM, Smyth RP. RNA Structures and Their Role in Selective Genome Packaging. Viruses. 2021; 13(9):1788. https://doi.org/10.3390/v13091788

Chicago/Turabian Style

Ye, Liqing, Uddhav B. Ambi, Marco Olguin-Nava, Anne-Sophie Gribling-Burrer, Shazeb Ahmad, Patrick Bohn, Melanie M. Weber, and Redmond P. Smyth. 2021. "RNA Structures and Their Role in Selective Genome Packaging" Viruses 13, no. 9: 1788. https://doi.org/10.3390/v13091788

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