Skip to main content
Microbiology and Molecular Biology Reviews : MMBR logoLink to Microbiology and Molecular Biology Reviews : MMBR
. 2018 Mar 14;82(2):e00067-17. doi: 10.1128/MMBR.00067-17

Emergency Services of Viral RNAs: Repair and Remodeling

Vadim I Agol a,b,, Anatoly P Gmyl a,c
PMCID: PMC5968460  PMID: 29540453

SUMMARY

Reproduction of RNA viruses is typically error-prone due to the infidelity of their replicative machinery and the usual lack of proofreading mechanisms. The error rates may be close to those that kill the virus. Consequently, populations of RNA viruses are represented by heterogeneous sets of genomes with various levels of fitness. This is especially consequential when viruses encounter various bottlenecks and new infections are initiated by a single or few deviating genomes. Nevertheless, RNA viruses are able to maintain their identity by conservation of major functional elements. This conservatism stems from genetic robustness or mutational tolerance, which is largely due to the functional degeneracy of many protein and RNA elements as well as to negative selection. Another relevant mechanism is the capacity to restore fitness after genetic damages, also based on replicative infidelity. Conversely, error-prone replication is a major tool that ensures viral evolvability. The potential for changes in debilitated genomes is much higher in small populations, because in the absence of stronger competitors low-fit genomes have a choice of various trajectories to wander along fitness landscapes. Thus, low-fit populations are inherently unstable, and it may be said that to run ahead it is useful to stumble. In this report, focusing on picornaviruses and also considering data from other RNA viruses, we review the biological relevance and mechanisms of various alterations of viral RNA genomes as well as pathways and mechanisms of rehabilitation after loss of fitness. The relationships among mutational robustness, resilience, and evolvability of viral RNA genomes are discussed.

KEYWORDS: RNA viruses, evolvability, fitness, mutational tolerance, repair, replication fidelity, robustness

INTRODUCTION

Replicative systems of RNA viruses are typical error-prone, primarily due to the low fidelity of their RNA-dependent RNA polymerases (RdRP) and the lack (in the overwhelming majority of cases) of proofreading mechanisms. Although the extents of replicative infidelity vary among different RNA viruses, the purified picornaviral RdRPs generally demonstrate high levels of nucleotide misincorporation, especially in the case of transitions, which sometimes have a rate as high as ∼10−4 per site per template copy (16). These values obviously depend on the experimental conditions and do not necessarily accurately correspond to the real situations in infected cells (which also vary). Nevertheless, taking into account that, for example, picornaviral genomes consist of ∼104 nucleotides (nt), such a level of enzyme infidelity is consistent with the acquisition, on average, of one mutation by each newly synthesized molecule of RNA of these viruses. Indeed, a recent study directly demonstrated that genomes of the progeny of a clone of poliovirus had, on average, more than two mutations each after merely two passages in tissue culture (7).

Interestingly, such replicative negligence is not an obligatory feature of RdRPs. Indeed, even a point mutation may increase the accuracy of these enzymes severalfold (812; also see the other references in this paragraph). However, a more faithful replication may result in a fitness loss by preventing accumulation of adaptive mutations (1322), although the relationships between fidelity and fitness may differ between virus-host systems, being influenced in part by the conditions under which the fitness assay is performed (2329). Moreover, a decrease in replicative fidelity may also lead to fitness loss, suggesting that the accuracy of replication of RNA viruses is naturally fine-tuned (29).

Due to replicative infidelity, populations of these viruses are highly heterogeneous, i.e., they exist as quasispecies (3034). Importantly, the level of heterogeneity depends not only on the intrinsic fidelity of the viral RdRPs but also on the mode of genome replication (35). Two distinct mechanisms, the “stamping machine” and “geometric replication” mechanisms, may be operating. The former implies that all the progeny genomes within a given cell are templated just by the infecting genome(s), whereas the latter posits that the newly synthesized genomes serve as templates as well, forming their own distinct “lineages.” Obviously, the heterogeneity of the final harvest of the infected cell should be much larger in the case of geometric replication and should depend on the number of genome “generations” in this cell. Specifically, it was estimated that, on average, 5 generations of poliovirus RNA were produced in a single HeLa cell under the experimental conditions used (35). The mode of replication and number of genome generations vary in different virus-host systems, thereby affecting the size of the space of genome diversity.

In any case, heterogeneous populations are expected to include numerous low-fit or even dead genomes. Indeed, a severalfold increase in the error rate (leading to the appearance of a few additional mutations in newly synthesized viral genomes) may result in a total extinction—“error catastrophe”—of the relevant population (3641). Replicative infidelity may be especially relevant to fitness if it concerns small viral populations and various kinds of bottlenecks which viruses may encounter. In particular, such bottlenecks occur when the virus has to overcome various barriers, whether they are intrahost (e.g., blood-brain barrier and others) or interhost (4245) barriers. On the one hand, overcoming various barriers may require a significant level of infidelity, leading in particular to the generation of adaptive mutations. On the other hand, if after overcoming such barriers infection is initiated by only a few or even a single low-fit infectious genome, the newly generated viral population may be more or less seriously invalidated or deadlocked altogether, a phenomenon known as the Muller ratchet (4649).

Notably, bottlenecking effects have an important but very poorly understood aspect. It is known that a single infectious dose in tissue cultures, e.g., a PFU of poliovirus, may contain as many as hundreds of virions. Although some of these virions may carry dead genomes, there is no reason to believe that such genomes constitute a majority, let alone the overwhelming one. Hence, only a minute minority of the potentially infectious genomes invading target cells appears to have the chance to initiate viral reproduction. Very little is known about the nature and factors affecting this phenomenon, except that the specific infectivity of viruses or viral RNA (number of virions or RNA molecules able to start productive infection) may vary by several orders of magnitude, depending on the primary structure of the genome (5053), and also varies for different host cells. The number of physical particles needed to avoid a fitness decrease due to a natural bottleneck is generally unknown, although these values may be quite relevant.

Notwithstanding these circumstances, RNA viruses demonstrate a remarkable potential for genetic and structural conservation under constant conditions. For example, sequences of portions of the genomes of several poliovirus strains studied in our lab (54, 55) turned out to be nearly identical to the sequences of these strains grown independently for nearly 3 decades in other countries (5660). Of course, such identity concerned the consensus (i.e., averaged) sequences. Under natural conditions, the primary structure of the genomes of RNA viruses undergoes more or less marked alterations, both neutral and adaptive, but the identity of the virus as belonging to a distinct type or species is usually retained. Much more rarely, sharp qualitative changes and generation of new viral taxa may also occur.

Somewhat paradoxically, both genetic conservatism and evolvability are based, to a significant extent, on replicative infidelity. When, for any reason, new infection is initiated by a debilitated viral variant, there is often the possibility of regaining the original fitness through acquisition of reversions or compensatory mutations. On the other hand, population heterogeneity includes viral variants that, regardless of their fitness in a constant environment, exhibit enhanced fitness if the population meets unfavorable conditions, e.g., innate or adaptive immunity, antiviral drugs, unfamiliar hosts, etc. For example, polioviruses are highly sensitive to the inhibitory effect of millimolar concentrations of guanidine hydrochloride. However, populations of this very sensitive virus always contain tiny proportions (depending on the drug concentration tested) of guanidine-resistant (gr) mutants (61), or even mutants whose growth requires the presence of this drug (62; also see below). The presence of drug-resistant and drug-dependent variants in largely sensitive populations has also been reported for other picornaviruses and other inhibitors (6371). Different variants within heterogeneous viral populations may complement each other in performing some functions, as appears to be the case with poliovirus (10, 14, 22, 32), although the converse situation, negative trans-dominance of debilitating mutations, has also been documented (72, 73).

Here we consider numerous ways in which RNA viruses overcome or mitigate negative effects of their replicative infidelity (and of genome-damaging environmental factors) and discuss why RNA viruses are remarkably robust despite their replicative infidelity and why they exhibit a remarkable evolvability despite their robustness. We focus predominantly on picornaviruses, one of the most extensively characterized viral families, but some additional lessons derived from the behavior of certain other RNA viruses are also considered. The regularities emerging from this analysis seem to be applicable to RNA viruses in general.

PICORNAVIRUSES AND THEIR GENOMES

Picornaviruses (representatives of the Picornaviridae family) are small (roughly 30 nm in diameter), nonenveloped, icosahedral animal viruses which are classified into >35 genera, ∼80 species, and hundreds of types and include pathogens causing important human and animal diseases, such as poliomyelitis, hepatitis A, common cold, myocarditis, encephalitis, foot-and-mouth disease, and many others (www.ictv.global/report/picornaviridae). They have a single-stranded RNA genome of positive polarity (i.e., translatable) containing ∼6,700 to 10,100 nt. This genome (Fig. 1A) carries a single open reading frame (ORF) that encodes a polyprotein, which is eventually processed into about a dozen “mature” functional proteins. The polyprotein is usually considered a modular structure composed of the following three parts: P1, containing capsid proteins VP1 to -4 and, in some viruses, also the leader protein L; P2, containing nonstructural proteins 2A, 2B, and 2C; and P3, containing nonstructural proteins 3A, 3B (or VPg, the RNA-priming protein), 3Cpro (protease), and 3Dpol (RdRP) (74, 75). As far as picornaviral proteomics is concerned, the most striking variability is exhibited by the nonstructural proteins L and 2A, which are involved mainly (though not exclusively) in the interaction with cellular innate immunity and generally are not essential for viral viability. They may be regarded as accessory or “security” proteins (76). Picornaviruses differ from one another not only with respect to the structure and function of these proteins but also by having various numbers of distinct molecular species: 0 to 2 for L (if L*, encoded by another ORF of the genomes of some cardioviruses, is also counted) and up to 5 for 2A (76, 77).

FIG 1.

FIG 1

Schematic representation of the genomes of picornaviruses. (A) Genome of monocistronic (most prevalent) picornaviruses. The leader L and 2A proteins (shown in red and not to scale) are highly variable in structure and function and may be dispensable for viral viability. They may be represented by several distinct copies or be absent in a given genome. The various positions of the replicative cre (oriI) cis-element (hairpins) are indicated. The VPg protein covalently linked to the 5′ end of the viral RNA is shown as a red circle. (B) Genome of dicistronic picornaviruses. The position of cre is unknown.

The genomes of some picornaviruses also have additional, relatively small overlapping ORFs (7880). In certain cases, the relevant proteins were shown to perform “security” functions, and it may be supposed that still-uncharacterized such proteins are mainly dedicated to the same profession. It may also be noted that picornaviruses have been described in which the proteins characteristic of this viral group are encoded in two separate ORFs, corresponding to P1 and P2-P3, respectively (81, 82) (Fig. 1B). Furthermore, enteroviruses of the G species which contain a new gene at the 2C/3A junction, encoding a predicted papain-like protease similar to that of coronaviruses, were recently isolated (83, 84). Also, the existence of a potential second ORF downstream of the entire main polyprotein-encoding sequence was reported (85). Still another possible pathway for evolutionary diversification of picornaviruses, genome segmentation, was demonstrated experimentally (see below). These examples illustrate the evolutionary potential of picornaviruses.

The polyprotein-encoding part of picornavirus genomes is flanked by 5′ and 3′ untranslated regions (5′ UTR and 3′ UTR) containing key cis-elements that vary markedly in structure in different picornaviral genera but perform similar functions, i.e., replicative (oriL and oriR; in the 5′ UTR and 3′ UTR, respectively) and translational (internal ribosome entry site [IRES]; predominantly in the 5′ UTR but also in the intercistronic region of bicistronic representatives) functions. The functional and structural features of these elements vary tremendously among different picornavirus genera (86, 87) (Fig. 2), and it is reasonable to briefly characterize them, especially those that are considered often.

FIG 2.

FIG 2

Examples of structural variability of the key cis-elements of picornavirus genomes. (A) oriLs of enteroviruses (as exemplified by this element in poliovirus type I), hepatoviruses (hepatitis A virus), cardioviruses (mengovirus), and kobuviruses (aichivirus). (B) IRES elements of type I (poliovirus type 1), type II (Saffold virus), type III (hepatitis A virus), type IV (HCV-like; bat hepatovirus), and type V (aichivirus). (C) Examples of different oriRs. Thin lines in all panels are used to link distinct structural elements and are not depicted to scale.

The ∼90-nt enterovirus oriL has a cloverleaf-like structure (88, 89) (Fig. 2A). This element plays multiple roles in viral reproduction, mainly through promoting formation of a complex ribonucleoprotein (RNP) structure involving several viral and host proteins (8993). An essential component of this complex is the viral RdRP (3Dpol), which is recruited there in the form of its precursor, 3CD, i.e., covalently linked to the viral protease (3Cpro) (91). This oriL-3CD interaction involves the hairpin domain d of oriL and the 3Cpro moiety of 3CD, as originally demonstrated by Andino et al. (90, 94). The interaction between these ligands is important for the initiation of the synthesis of both the viral (positive) and complementary (negative) RNA strands (89, 93, 9598). It was recently shown that (at least in the case of coxsackievirus B3 [CVB3]) oriL, which is required for efficient genome replication, is not indispensable for viral viability: its removal does not kill the virus but rather decreases the efficiency of its RNA replication ∼105-fold (99). It should be noticed, however, that the nonessentiality of oriL has so far been shown only for this particular virus (see also a reservation at the end of this section). oriLs of picornaviruses other than enteroviruses exhibit great variability (some examples are presented in Fig. 2A), but their detailed and functional characterization has yet to be performed.

Initiation of translation of picornaviral RNA is accomplished via a cap-independent, IRES-dependent mechanism (100103). Structures of IRES elements of different picornaviruses exhibit remarkable variations (86, 87, 104110) and are represented by at least five structural types specific for this viral family or related to the IRES elements of certain flaviviruses (111, 112) (Fig. 2B). These translational cis-elements serve to bind ribosomes in order to create conditions for the initiation of translation (87, 107, 113). This process requires participation of both canonical initiation factors and IRES-specific (and cell-specific) trans-acting factors (ITAFs) (109, 114).

The structures of picornavirus oriRs are rather variable (86) (Fig. 2C). Even within a single genus, Enterovirus, there are several distinct structural classes of this cis-element. In many enteroviruses, it is represented by a multidomain quasi-globular RNA structure maintained by a tertiary kissing interaction. Polioviruses and other representatives of enterovirus C species have two stem-loops, X and Y, while enteroviruses of species B have an additional stem-loop, Z, and human rhinoviruses have a single stem-loop (115121). Picornavirus oriRs are polyfunctional (122), being involved in negative RNA strand initiation (116, 119, 123), polyadenylation of the genomic RNA (124), and control of translation through (primarily but not only protein-mediated) interaction with the 5′ UTR, which results in noncovalent circularization of the genome (96, 125130). The peculiarities of the oriR structures of various enteroviruses are in part responsible for their cell-specific fitness differences, which are especially prominent in neural cells, a fact which has direct relevance to viral pathogenicity (131133).

It is noteworthy that enterovirus oriR, which is conserved and pivotal for efficient genome replication, is nevertheless not essential, as demonstrated by the viability (though with a markedly low fitness) of viruses lacking or having a severely damaged element (118, 123, 124, 134). On the other hand, deletion of a stem-loop of oriR of mengovirus (a cardiovirus) was reported to be lethal (135).

As mentioned above, in addition to their individual functions, the enteroviral 5′ UTR and 3′ UTR appear to interact with one another through the mutual affinity between proteins to which they bind, which results in noncovalent circularization of the picornavirus genome playing an important part in the control of viral translation.

The picornavirus genomes also contain an additional important replicative cis-element, cre (cis-acting replicative element) (136138), also known as oriI. It folds into hairpin-containing structures (139, 140) that are located in different regions of the polyprotein ORF or in the 5′ UTR in different picornaviruses (86, 141) (Fig. 1A). The loop of its hairpin contains an oligoadenylate sequence which serves as the template for the uridylylation of the protein VPg (3B) by the viral RdRP (142, 143). The uridylylated form of VPg, VPgpUpUOH, is the primer for initiation of the synthesis of viral RNA molecules (144148). Again, mutational inactivation of cre may not kill the virus but rather decreases the efficiency of replication of its RNA ∼105-fold, as reported for CVB3 (149). Note that oriL and cre elements of enterovirus genomes appear to work in cooperation, with both responsible for the correct VPg-pUpU-dependent initiation of the positive RNA strands. The loss of functional cre resulted in deletion of large 5′-terminal portions of CBV3 oriL and, reciprocally, deletion of oriL (which may occur upon chronic infection both in cultured cells and in diseased organisms [150153]) was reported to make functional cre dispensable (149).

Picornavirus RNAs also have some additional conserved functional structures, such as the hairpin inhibiting the host antiviral enzyme RNase L (154, 155) as well as some other potentially biologically relevant elements with poorly characterized functions (156158).

A separate question concerns the role and mechanism of generation of the 3′-terminal poly(A) tract. This tract is believed to play a role in the preservation of the genomic end, in particular by association with the poly(A)-binding protein (125, 159), which promotes the above-mentioned circularization of the viral RNA through protein-protein interaction. The poly(A) tract is synthesized by using the 5′-terminal poly(U) sequence in the negative viral RNA strands as the template, which in turn is templated by the poly(A) segment in the parental positive strand (160). It was noted, however, that the poly(A) stretches may be markedly longer than the poly(U) ones (161, 162).

Admittedly, great care should be taken in interpreting the observed effects of various detrimental alterations of viral genomes. Such effects may be modified by the presence in the investigated quasispecies populations of minute, hardly detectable (by common sequencing techniques) amounts of different genome variants, which may serve as complementation or recombination partners. For example, the population of oriL-truncated CVB3 RNA in human endomyocardial tissue was reported to contain 0.9% molecules with an apparently intact 5′ end (153).

CONSERVATION VERSUS EVOLVABILITY

Evolutionary mechanisms serve two opposite goals: to maintain the stability of viral genomes (to retain their structural and functional identity), on the one hand, and to allow their evolvability (to ensure their capacity to change), on the other. These operate mostly under constant and changed environmental conditions, respectively.

So far, no ancient picornaviral RNA genomes from paleontological or archeological samples are available, and the “ages” of the relevant viral species can be roughly evaluated only on the basis of comparison of nucleotide sequences of the most distant representatives of their genomes. Due to inherent limitations of bioinformatics tools owing to the saturation of the level of nucleotide substitutions and the effects of purifying selection, they may provide more or less reliable estimates only for relatively recent time intervals (up to several hundred or thousand years) (163). Given this limitation, the estimates indicate that picornaviral species may retain their recognizability for at least centuries (164). On the other hand, taking into account the conservation of the key structural features of their virions and some essential nonstructural proteins, there is little doubt that all picornaviruses had a common ancestor in the very far past, i.e., are a monophyletic group whose members diverged through various qualitative jumps (165).

Mechanisms involved in genome conservation and evolvability are considered below.

TYPES AND MECHANISMS OF GENETIC MODIFICATIONS

During their life cycle and evolution, viral RNA genomes should cope with various intrinsic and extrinsic factors potentially capable of disturbing their functions. Adverse phenotypic effects of a plethora of natural or engineered genetic alterations of viral genomes are reported in the literature. Only some illuminating examples are considered here.

Point Mutations

A major source of genetic modifications experienced by the genomes of RNA viruses is the inaccuracy of their replication machinery, caused by the above-mentioned error-prone performance of the viral RdRP and the lack (with the exception of a small set of viruses [see below]) of proofreading capacity. Also, eukaryotic cells possess two families of nucleotide deaminases: double-stranded RNA-specific adenosine deaminases (ADARs), which convert adenosine into inosine (166, 167), and single-stranded RNA-specific cytidine deaminases (APOBECs), which convert cytidine to uridine (168). Although the effects of these multifunctional enzymes on RNA viruses are thus far focused mostly on different aspects of virus-host interactions (169, 170), their potential mutagenic effects should not be underappreciated. Such effects of ionizing radiation and other environmental mutagenic factors should not be ignored either.

Insertions, Deletions, and Replacements

An important class of modifications of RNA genomes is represented by their rearrangements, i.e., acquisition of deletions, insertions (including duplications), and replacements of genomic parts by related or unrelated sequences. The rate of generation of insertions/deletions (indels) during a cycle of picornavirus reproduction appears to be comparable to that of the appearance of point mutations, as suggested, for example, by studies on the mechanisms of recovery after adverse alterations of the poliovirus genome (171, 172). Also, ready accumulation of deletion-containing, so-called defective interfering (DI) genomes (see below) under certain conditions demonstrates that such rearrangements are relatively common products of normal reproduction of RNA viruses. Nevertheless, indels are less common in natural viral populations due to their removal by negative selection because of generally strong fitness-impairing effects. These effects may be due to various mechanisms, the most important being interruption (or shifting) of ORFs, destruction of important protein structures and RNA cis-acting elements as well as (for long inserts) their A/U content (173), and limitation in the RNA genome length (174176).

Genomic rearrangements may be caused by intra- or intermolecular recombination, and the latter may involve genomes of different viruses with various levels of relatedness and even cellular RNAs, resulting in such cases in replacements of portions of the viral genomes by foreign sequences. The key role of such replacements in the evolution of RNA viruses is most clearly illustrated by the above-mentioned qualitative differences in the structure of the replicative and translational cis-elements (Fig. 2), shared sometimes by viruses of different families (as is the case, for example, with IRES elements of some picorna- and flaviviruses).

There are two fundamentally distinct mechanisms of RNA recombination: replicative and nonreplicative (Fig. 3). The possible existence of these mechanisms was considered long ago, when appropriate tools for their investigation were not yet available (177179). Further studies demonstrated that RNA viruses exploit both mechanisms (180).

FIG 3.

FIG 3

Models of replicative and nonreplicative recombination. See the text for details.

Replicative recombination occurs via template switches, whereby a working molecule of RdRP prematurely terminates its elongation; the newly synthesized, uncompleted chain departs from its template and lands on a new template to serve as the primer for continuation of the synthesis. This is a generally accepted model (54, 181185), and most recently, it was supported by demonstration of the dependence of recombination frequency on the fidelity of the viral RdRP (12, 186188).

Nonreplicative RNA recombination implicates joining of fragments originating from distinct viral (or cellular) RNA molecules without involvement of the viral RdRP, though this enzyme is of course required for further copying of chimeric molecules generated thusly. This kind of recombination was first demonstrated to occur in experiments with pairs of segments originating from poliovirus genomes disrupted in either the 5′ UTR or the RdRP-encoding region (189, 190). Subsequently, it was observed in other picornaviruses (191, 192) as well as in flaviviruses (193196). As discussed elsewhere (197), this mechanism may also be responsible for certain earlier observed cases of recombination between alphaviruses (198) and rubiviruses (199) (both belonging to the Togaviridae family) and hence appears to be a general phenomenon.

It should be admitted that the mechanistic aspects of both types of RNA recombination are as yet purely understood. Both replicative and nonreplicative recombination can generate homologous (precise) and nonhomologous (imprecise) recombinants, but the nonreplicative mechanism is expected to produce predominantly nonhomologous ones. Unfortunately, no tools are currently available to learn which of these two mechanisms is operative during a given case of natural recombination.

Genome Truncation and Disruption

Both termini of viral RNA genomes (and their complementary sequences) are known to be vulnerable targets for the host cell 5′- and 3′-exonucleases (200). Genome truncation is usually accompanied by a more or less marked loss of fitness and is a subject of numerous studies (see below). On the other hand, situations involving disruptions of viral genomes are very rarely investigated directly due to significant experimental difficulties. Yet losses of genome integrity are expected to occur often enough due to several mechanisms, such as activities of defensive antiviral nucleases (e.g., RNase L) and endoribonucleases involved in RNA interference as well as various mechanisms of host mRNA decay, including the nonsense-mediated one (200202). The involvement of the latter is expected due to the likely presence of stop codons in the viral quasispecies caused by the infidelity of replication. In addition, environmental factors, such as ionizing radiation and alkylating agents, may also lead to RNA genome disruption. UV-induced viral RNA self-cleavage was recently reported (203).

Randomization

Although the randomization of portions of viral RNA genomes is an artificial intervention, the results obtained with the aid of this approach can be quite relevant to the topic of this review. There are different reasons for experiments with selection of viable viruses from pools of viral RNA with randomized segments. If the segments are relatively short, such experiments can reveal structural features required to ensure viability of the virus and affecting its fitness. In other words, such experiments may provide valuable information about the phenotypic effects of a set of point mutations. On the other hand, randomization of larger genome segments can be regarded as just the replacement of a functional part of the genome with irrelevant (but not necessarily inactive [204]) sequences.

(RELATIVE) NEUTRALITY OF VARIOUS MUTATIONS

Although a significant proportion of point mutations in the genomes of picornaviruses (7) as well as various other RNA viruses (205208) are detrimental, many nucleotide alterations are fitness neutral or exhibit relatively mild fitness defects. This ability to more or less tolerate genetic alterations may stem from different roots. One of the major ones is the degeneracy of the genetic code. Many nucleotide point mutations in the protein-encoding regions are synonymous, and synonymous mutations are often neutral. Remarkably, simultaneous introduction of 1,297 synonymous nucleotide changes into the poliovirus genome did not appreciably change the viral phenotype (53). However, the biological equality of synonymous codons has several limitations.

First, synonymous mutations are not necessarily neutral because of the phenotypic significance of codon (and codon pair) biases, as shown for picornaviruses (50, 51, 53, 209212), other positive-strand RNA viruses (213217), and negative-strand RNA viruses (217223). These biases may be due mainly to the effects on kinetics of translation (e.g., owing to uneven representation of tRNA species corresponding to different synonymous codons and to some features of the ribosome machinery [224]). In particular, the different rates of reading of synonymous codons affect the dynamics of folding of the generated proteins (225). In addition, synonymous substitutions in the cis-acting RNA elements located within the protein-encoding sequences may also exhibit phenotypic effects (86, 138, 147, 157, 226). The possible involvement of some other mechanisms, such as the existence of alternative functional ORFs and the relevant frameshifting signals (78, 227231) as well as the abundance of CpG and UpA dinucleotides (232236), should also be taken into account.

Mutational alterations of amino acids (i.e., nonsynonymous nucleotide substitutions) do not necessarily result in changed fitness. Even substantial alterations of the chemical nature of mutated amino acids do not obligatorily cause appreciable phenotypic changes, as exemplified by the outcome for replacements of certain charged amino acids by alanine in the 2C protein of poliovirus (237) or for a Val-to-Arg replacement in the RdRP of mengovirus (238). Similarly, some mutations in the TGK peptide of the oriL-interacting motif of 3Cpro do not exhibit any appreciable phenotypic alterations, at least in in vitro experiments (M. A. Prostova, E. I. Smertina, D. V. Bakhmutov, A. A. Gasparyan, E. V. Khitrina, M. S. Kolesnikova, A. P. Gmyl, and V. I. Agol, unpublished data). Even prevention of viral polyprotein cleavage at a canonical site may be relatively well tolerated, as demonstrated with engineered foot-and-mouth disease virus (FMDV) mutants having amino acid substitutions preventing disruption of the bond between the VP1 and 2A moieties normally cleaved by the viral 3Cpro protein (239, 240). The 2A protein of this virus is composed of merely 18 amino acids, and capsids of the viable progeny of the mutants contained the VP1-2A fusions instead of VP1.

Another important factor contributing to mutational neutrality is the degeneracy of the spatial RNA structure. Regulatory RNA cis-elements usually are composed of active parts ensuring specific RNA-RNA or RNA-protein interactions and scaffolding parts involved in retaining these regulatory parts in the appropriate conformation. These scaffolding functions can be fulfilled by oligonucleotide elements with different primary structures. Thus, the secondary structure of the 5′ UTR of the circulating polioviruses may be conserved despite the acquisition of various point mutations. Moreover, even active, ligand-interacting moieties of RNA cis-elements can also exhibit some degree of degeneracy, allowing replacement of a nucleotide by another one without an appreciable loss of function. This is true, for example, for the 5′-end-adjacent cis-element oriL of the poliovirus RNA, which, as mentioned above, interacts with the viral nonstructural protein 3CD. Quite illuminating (and unexpected) results in this respect were obtained upon the randomization of the apical tetraloop and two adjacent base pairs of domain d of poliovirus oriL (52). These experiments demonstrated that each position in the 39 unique sequences of this octanucleotide in 62 investigated viable plaque-forming viruses could be occupied by any nucleotide (with the exception of one position, which lacked U), though with certain sequence preferences. A closer look indicated that the tetranucleotide corresponding to the loop in nearly half of the isolates fitted the YNMG (Y = U/C, N = any nucleotide, and M = A/C) consensus, and the spatial structures of the relevant tetraloops are known (241). Certain tetranucleotide loops of the genomes of several other isolates had sequences compatible with either a YNUG or GNUA consensus. Some tetraloops with such sequences are known to be able to adopt YNMG-like conformations as well (242, 243). When genomes with various YNMG or YNUG tetraloops were engineered, they (and some genomes with GNUA tetraloops) exhibited wild-type-like phenotypes. Thus, tetraloops able to fold stably into a YNMG-like spatial structure appeared to be well-accepted partners for their protein ligands, clearly illustrating that functional degeneracy of spatial RNA structures may contribute to viral mutational robustness (also see below).

Many experimentally introduced deletions in other parts of the untranslated regions of picornaviral genomes did not result in appreciable functional deficiencies (244, 245). Even the entire deletion of two conserved secondary structure domains of the IRES still did not kill the virus (246).

Various indels in the protein-encoding part of the genome may not be accompanied by marked fitness losses. Thus, engineered insertions of various foreign functional elements, e.g., antigens (e.g., see references 173, 175, and 247 to 258), tags (e.g., see references 176 and 259 to 264), large structures such as IRES elements creating bicistronic genomes (265267), or even IRES elements together with sequences encoding additional polypeptides (268270), may be relatively well tolerated and not infrequently proved to be genetically stable (see below, however). Relatively long deletions in the C-terminal region of the FMDV nonstructural 3A protein were not accompanied by any significant phenotypic alterations, at least in vitro (271). Replacements of octapeptides in an antigenically dominant loop of the FMDV VP1 protein by unrelated sequences also produced viable and relatively stable viruses (272, 273).

The occurrence of indels that are not markedly harmful has been documented for noncoding regions of natural picornaviral genomes as well. For example, length differences of up to several dozen nucleotides were registered for the 5′ UTRs (60) and 3′ UTRs (274, 275) of closely related representatives of polioviruses and FMDV, respectively. The indels may not be strictly neutral, but if they are associated with some fitness cost, the detrimental phenotypic changes are likely to be suppressed by some second-site mutations (see below). In some cases, such indels appeared to even be advantageous, judging by their strong conservation in representatives of a given picornavirus species, as is the case, for example, with the triplication of the VPg gene in the FMDV RNA (276, 277) and the duplication of the 5′-terminal cis-element of the bovine enterovirus RNA (278, 279). Conserved duplications in untranslated and coding RNA regions of other viruses have been described as well (cf. references 280 to 284).

Notably, replacements of the functional genomic parts by functionally analogous parts of the genomes of not closely related viruses may sometimes also be relatively well tolerated, as exemplified by engineered poliovirus genomes with the IRES of encephalomyocarditis virus (EMCV; a cardiovirus) (117, 268) or human hepatitis C virus (HCV) (285) in place of their own structurally different IRES.

It should again be noted that the attribution of neutrality to specific mutations has to be done cautiously. The true neutrality of a given natural mutation and quantitative estimations of relative fitness are rarely investigated in rigorous experiments. A mutation can be neutral under certain conditions but may be linked to an altered phenotype under others; host-dependent or cell-dependent mutations are good examples of this. Thus, a point mutation (A637C) within a double-stranded stem of the type II IRES of Theiler's murine encephalomyelitis virus (TMEV; a cardiovirus), while not markedly affecting the capacity of the virus to grow in nonneural cells, resulted in a several-orders-of-magnitude decrease in its neurovirulence (286). This attenuating effect was due to a decreased affinity of the mutated IRES for the neural form of polypyrimidine tract-binding protein (nPTB) required for efficient initiation of translation of the viral RNA in these cells. FMDV with an extended (>150 nt) deletion in the replication-controlling 5′-terminal hairpin of its RNA did not demonstrate any appreciable phenotypic changes upon infection of cells with deficient innate immunity but proved to be highly attenuated in a mouse model (287). Furthermore, a mutation can be truly neutral within a certain genetic context but may confer an altered fitness within another context owing to epistasis (288). Also, it was demonstrated recently that certain synonymous substitutions, while not immediately changing the viral phenotype in vitro, might markedly diminish the mutational tolerance. Thus, the replacement of Ser and Leu codons in some genes of CVB3 and influenza A viruses by synonymous ones more prone to be converted into stop codons by point mutations rendered these viruses more vulnerable to natural and drug-induced mutagenesis (289).

On the other hand, it is not always correct to consider the mutational loss of the capacity to trigger an obvious cytopathic effect (CPE) (e.g., a lack of plaque-forming activity) sufficient evidence for killing of the virus, as usually done. For example, as already mentioned, complete destruction of the CVB3 replicative element oriL (99) or cre (149), resulting in a loss of cytopathic activity, is nevertheless compatible with viral viability: such viruses are able to grow, though with a greatly diminished efficiency, causing persistent, noncytopathic infections. An FMDV variant that accumulated various debilitating mutations during multiple plaque-to-plaque (bottlenecking) passages represented another example of conversion of a lytic virus into a noncytocidal one (290).

NATURAL TOOLS FOR CURING DAMAGED RNA GENOMES

Viruses have evolved a number of efficient tools to cope with the potential genetic damages caused by the infidelity of their replicative machinery (as well as by other adverse factors [see references 291 and 292]). One such tool, obviously, is negative selection resulting in the elimination of less-fit variants. However, this mechanism cannot ensure the maintenance of genetic stability upon overcoming various bottlenecks. In these cases, the key roles are played by the rehabilitation tools that are detailed below, but the most general principle is to “fight fire with fire.” In other words, the impairments caused by nucleotide substitutions can be counteracted by either reversions or compensatory second-site mutations, owing again to the infidelity of viral RdRPs. More extended genetic alterations, such as indels, can be restored or compensated by intra- or intermolecular rearrangements based on RNA recombination. Of course, recombination may also cure some defects caused by point mutations.

Injured RNA genomes, even if they are dead, as such, may survive for at least some generations due to complementation, i.e., help provided in trans by proteins (or RNA cis-elements) encoded by their coinfecting viruses. This mechanism of cooperative interaction was described long ago for the case of drug-sensitive and drug-resistant (or -dependent) picornaviruses (293297), but its wider biological relevance became especially appreciated after the realization that viral populations are represented by quasispecies, i.e., swarms of closely related but distinct individuals (10, 14, 32, 73, 298301). The prolonged survival of impaired genomes owing to intrapopulation complementation may provide time for their adaptive remodeling, resulting in the restoration of their capacity for independent existence.

REHABILITATION AFTER ADVERSE CHANGES IN THE UNTRANSLATED REGIONS OF THE GENOME

5′ UTR

As mentioned above, the picornaviral 5′ UTRs contain at least two important functional elements, involved in genome replication (oriL) and translation (IRES), which exhibit marked structural differences in different representatives of this family. They may be separated from each other as well as from the downstream ORF by spacers with various structures and lengths. Although these spacers usually exhibit a marked level of conservation, their specific functions are so far defined rather poorly. The effects of adverse modifications of distinct 5′ UTR parts are considered separately.

oriL.

Generally, enterovirus genomes begin with two unpaired uridines. However, RNAs of CVB3 and poliovirus lacking these 5′-terminal nucleotides were found to be infectious and able to restore their wild-type structure (302, 303). The deleted residues were most likely provided by the VPg-pU-pU primer. However, such a mechanism did not appear to work with similarly deleted RNA of hepatitis A virus (303).

The poliovirus oriL can sustain various internal alterations without significant functional impairments. This circumstance endows it with a substantial mutational robustness and creates numerous possibilities for rehabilitation through compensatory second-site mutations should some point mutations inflict an appreciable fitness loss (52). The rehabilitation is, however, not always complete. Although different apical tetraloops of domain d potentially able to acquire a YNMG-like conformation are compatible with viability, they endow different levels of fitness. A likely explanation for this phenomenon is the dynamic nature of RNA folding, specifically the instability of a given conformation (caused by thermal motions of constituents) and its existence in equilibrium with others (304, 305). For a given RNA spatial structure, the time of existence in a distinct conformation depends on specific nucleotide sequences. Thus, not all YNMG-like conformations are functionally equal with regard to recognition by their ligands. A factor underlying this inequality is the proportion of time during which the element really adopts the conformation recognizable by its ligand (in this case, the 3CD protein). The fitness of viruses with relatively “poor” YNMG-like folding was enhanced by mutational “improvements” increasing the probability of adopting the necessary conformation (52). Moreover, some tetraloops with known non-YNMG folding (e.g., of the GNRA class) were not lethal but rather quasi-infectious, i.e., no viable viruses with exactly the engineered inappropriate structures could be recovered upon the transfection of susceptible cells, but plaques caused by pseudorevertants did appear upon prolonged incubations. An explanation may consist of the assumption that even these “incorrect” tetraloops (or, rather, their minute proportions) may acquire an acceptable conformation (52, 305, 306). A strong, though not fatal, debilitation of poliovirus caused by inversion of the sequence of stem-loop b of oriL could be partially ameliorated by spontaneous acquisition of nucleotide substitutions in the loop (117).

Deleterious effects of mutations in the RNA cis-elements may also be restored by compensatory mutations in their ligands. As originally demonstrated by Andino et al. (90, 94) and studied in more detail by Prostova et al. (52; unpublished data), harmful alterations of domain d of oriL may lead to fitness-restoring amino acid replacements in an RNA-binding motif of protein 3Cpro, in particular by changes of the conserved tripeptide TGK into IGK, VGK, and some others.

However, if the original debilitating mutation completely prevents the replicative ability, then there is obviously no possibility for rehabilitation.

IRES.

The readiness of reversion of adverse point mutations in the picornavirus IRES was well illustrated by early studies on the instability of the attenuating mutations in this element of the Sabin oral polio vaccine (OPV) strains. In the guts of vaccinees, such reversions occur very often and quite rapidly at positions 480, 481, and 472 in the RNAs of OPV serotypes 1, 2, and 3, respectively (307311; reviewed in references 312 to 314), leading to restoration of their somewhat impaired secondary (for serotypes 1 and 3) or tertiary (for serotype 2) structures (104, 311, 315) (Fig. 4) and, consequently, translational activity (316319) and neurovirulence (308, 315, 320, 321).

FIG 4.

FIG 4

Some attenuating mutations in the IRES elements of Sabin strains and their deattenuation/reversion. (A) Portion of the secondary structure of the IRES of poliovirus type 1. In the parental Mahoney strain, the structure is partly stabilized by pairing between nt A480 and U525 (marked in green), while it is partially destabilized in Sabin-1 by the A480G substitution (red), contributing to the attenuation of neurovirulence. In the organisms of vaccinees or their contacts, either reversion to A480 or the compensatory mutation (pseudoreversion) G525C often takes place, resulting in partial deattenuation. (B) Portion of the proposed tertiary structure of the IRES of poliovirus type 2. The C398-G481 interaction (green) in the genome of P712, the wild-type predecessor of Sabin-2, was replaced with U398-A481 (red) in the vaccine virus. In the organisms of recipients of the vaccine, the deattenuating replacement A481G usually takes place, accompanied by the reversion at position 398, thereby enhancing the potential for tertiary interaction.

Similar deattenuating, fitness-increasing effects could be achieved by second-site mutations (pseudoreversions) resulting in the restoration of the impaired secondary or tertiary IRES structure. Wild polioviruses of serotype 1 contain an A480-U525 base pair that strengthens a double-stranded element (104) (Fig. 4A). As noted above, the Sabin strain of this serotype has a destabilizing mutation at position 480 (A to G) contributing to its attenuated phenotype, but this destabilization is not usually conserved in vaccinees and their contacts. The strengthening of this base pairing might be accomplished not only by the reversion (G480A) but also by a compensating pseudoreversion (U525C) (311, 322), which also resulted in a similar increase in virulence (323) and in restoration of the in vitro translation efficiency (311). Structural modeling suggested that G481 of the predecessor of the Sabin-2 strain, strain P712, could potentially participate in a long-range interaction with C398, supporting the generation of a tertiary structure element (311) (Fig. 4B). On the other hand, the Sabin-2 strain possesses A481 and U398, which also can pair with each other, suggesting the functional relevance of this interaction. Nucleotide A481 contributes to the attenuation phenotype of this vaccine virus, but it readily reverted to G in organisms of vaccinees, which is the event expected to disrupt the tertiary interaction. However, the potential for formation a stable C398-G481 pair was regained in most isolates due to the reversion at position 398 as well (234, 311, 322) (Fig. 4B).

In addition, adverse phenotypic effects (e.g., a temperature-sensitive [ts] phenotype) of attenuating mutations in the poliovirus IRES could be partially (and in a cell-specific manner) alleviated by amino acid replacements in the nonstructural 2A protein (324), although the molecular mechanism underlying this improvement is not defined.

Poliovirus IRES (structural type I) (Fig. 2B) also exhibits remarkable plasticity in the response to debilitating indels. Among its several functional elements, there is a tandem of an oligopyrimidine (Yn) and a cryptic AUG (located at positions 559 to 563 and 586 to 588, respectively, in the Mahoney strain of serotype 1) separated by the 22-nt spacer (171, 325327). Engineered poliovirus genomes with 23-nt or 39-nt inserts or 8-nt deletions in this spacer proved to be quasi-infectious (171, 172). In the case of insertions, full or partial fitness restoration was spontaneously achieved in the transfected cells by either deletions resulting in the shortening of the Yn-AUG spacer to a length close to its natural value or the appearance of a new, noninitiator AUG upstream of the natural and defunctionalized cryptic AUG, again with a more or less “normal” distance from Yn (Fig. 5). The fitness recovery of genomes with shortened spacers may be accomplished in three different ways: acquisition of an insert (e.g., 9 nt); generation of a new, functional cryptic AUG downstream of the original cryptic AUG and at a comfortable distance from Yn; or deletion of a large (∼150 nt) sequence, resulting in the appearance of a new, comfortably distanced partner for Yn, the initiator A743UG (Fig. 5). Interestingly, engineered mutant polioviruses with similar extended deletions proved to be markedly attenuated in the monkey neurovirulence assay (328).

FIG 5.

FIG 5

Pathways of rehabilitation after damaging indels in the poliovirus IRES. The wild-type poliovirus IRES (partly presented in the black frame) has a functionally important tandem of an oligopyrimidine (Yn; green rectangle) and a cryptic (noninitiator) A586UG (green circle) with a spacer of 22 nt. The initiator A743UG is marked as a red circle. Engineering of debilitated constructs and selection of well-fit pseudorevertants are indicated by black and green arrows, respectively. Engineered insertions and deletions are depicted by red and dashed black lines, respectively. Deactivated cryptic AUG and spontaneously acquired functional cryptic AUGs are shown as yellow and purple circles, respectively. See the text for details.

One approach to investigating effects of heterologous replacements in the IRES sequence consists of scanning mutagenesis, whereby different adjoining oligonucleotides are replaced by a more or less random oligonucleotide of the same length. Scanning mutagenesis was used to generate 14 octanucleotide replacements in different loci of the poliovirus IRES (246). Several of them did not kill the virus, though they markedly decreased its fitness. Some mutants exhibited a ts phenotype, while some others, though initially considered noninfectious, were in fact quasi-infectious and generated viable pseudorevertants upon further passaging. Their “revival” was due to one or two second-site point mutations in the case of one such mutant and to an ∼150-nt large deletion in the downstream region of the 5′ UTR in the case of another (246). The reversions appeared to be host dependent, since the quasi-infectious genomes generated different sets of pseudorevertants in HeLa and neuroblastoma cells (329). The latter observation was likely due to distinctive advantages of certain IRES structures in HeLa and neuroblastoma cells (330332). Note that the replacement at positions 585 to 592, destroying an RNA helical structure and the cryptic AUG codon, generated a large deletion resulting in creation of a tandem of Yn and the initiator AUG, similar to that illustrated in Fig. 5.

The rehabilitative capacity of the IRES elements of structural type II was studied less extensively, but the general conclusions were in line with those just described. Destabilization of one of the helices of the IRES of EMCV by a point mutation could be compensated at least partially by acquisition of the true reversion or second-site mutations, particularly (but not only) those that restored the stability of the helix (333, 334). The increased fitness of some pseudorevertants of this mutant could be ascribed to alterations outside the IRES. An apparently compensatory mutation outside the IRES (in the leader protein) was also discovered upon transfection of a genome containing a point mutation in an unpaired segment of the IRES, but explanations for the effects of such second-site mutations are lacking.

Another kind of compensatory mutation concerns the above-mentioned A637C replacement in the TMEV IRES (286). Although this mutation resulted in a strong attenuation of neurovirulence of the virus, intracerebral injections of large doses triggered encephalitis in some mice. The virus isolated from the brains of such animals contained a mutation, U649C, in a loop of the same module of the IRES. This mutation increased the IRES-nPTB affinity and restored neurovirulence of the virus (286).

Large insertions between Yn and the initiator AUG of the TMEV RNA resulted in a significant attenuation of virus neurovirulence for mice (335). The viruses isolated from the brains of animals that received large doses of mutated viruses and did succumb to the disease invariably had acquired either deletions or a new AUG-generating mutation, both adjusting the Yn-AUG distance to a more comfortable (closer to the wild-type) value, resembling the above-described results with the poliovirus IRES of structural type I.

5′ UTR spacer.

The oriL and IRES in various picornaviruses may be separated by conserved spacer sequences with poorly defined functions. Their alterations might also lead to adverse fitness effects, which may be ameliorated by second-site mutations. Thus, a 4-nt insert at position 220 of the 5′ UTR of Sabin-1 poliovirus resulted in a small-plaque phenotype, but different large-plaque variants were selected after passages (244). The fitness gain was accompanied in each case by the acquisition of two second-site point mutations in different regions of the 5′ UTR, obligatorily including one of the two above-mentioned deattenuation mutations, at position 480 or 525. A strong ts phenotype caused by a 4-nt deletion at the same locus could be ameliorated spontaneously by enlarging this deletion to 41 nt (336).

3′ UTR

oriR.

Complete or partial deletions of the picornaviral oriR may result in a marked suppression of genome replication at normal or supra-optimal temperatures. These adverse effects may be ameliorated, at least partially, upon passage of the mutants. Thus, an 8-nt insert in the poliovirus 3′ UTR resulted in ts viruses, from which ts+ revertants were selected (337, 338). The phenotypic improvement was accompanied by deletions of 7 or 8 partly original, partly inserted nucleotides (in different isolates, this resulted in four distinct but closely related structures) and by an additional point mutation in some of them. A 14-nt deletion of a stem-loop of the structurally different oriR of mengovirus led to a quasi-infectious genome (135). A partial compensatory effect was achieved by an amino acid substitution in the viral RdRP, perhaps through increasing its affinity for oriR. If so, this pseudoreversion gives another example of compensation of a mutation in an RNA cis-element by alteration of its protein ligand. An even stronger fitness-increasing effect resulted from natural acquisition of a combination of this RdRP mutation with a mutation at another 3′ UTR position (135).

Debilitating effects of various disruptions of the tertiary kissing interactions between stem-loops X and Y of the coxsackievirus A-9 (CAV9) oriR (Fig. 2C) could be compensated, at least partially, by the natural nucleotide substitutions restoring this interaction, including a variant in which the kissing interaction was shifted by 1 nt (120). Interestingly, disruption of certain single base pairs in the kissing interaction of the CVB3 oriR may kill the virus (119), whereas more extensive destabilization of this interaction led to quasi-infectious genomes which, after a relatively long period of marginal replication, could increase (though far from completely) their fitness through either a single-nucleotide insertion, allowing an alternative, sufficiently strong kissing, or, unexpectedly, the complete loss of the X and Y domains (124) (Fig. 6).

FIG 6.

FIG 6

Example of rehabilitation of a picornavirus by the loss of an important RNA cis-element. Functional replicative activity of oriR of the coxsackievirus B3 genome requires a tertiary (kissing) interaction between its X and Y domains. Mutational alteration of 4 nt in the loop of domain X (shown in red) destroyed this interaction and rendered the viral genome quasi-infectious. A significant fitness gain could be achieved by the spontaneous destruction (rather than repair) of oriR through the deletion of domains X and Y. The termination codon of the polyprotein ORF, UAG, is underlined.

Poly(A).

A set of poly(A)-lacking CVB3 genomes with either plain poly(A) deletion or such a deletion together with removal or randomization of the entire oriR, or with the replacement of poly(A) with other homopolymeric sequences, proved to be viable (124). The recovered viruses contained a variety of 3′-terminal sequences, all ending with the regenerated poly(A) sequence. The genome of a virus recovered after transfection with the 3′ UTR-lacking RNA terminated with UAGUCGAn, where the doubly underlined triplet is the translation termination codon of the polyprotein ORF and the singly underlined one is from engineering. Normally, poly(A) is templated by the 5′-terminal poly(U) sequence of the viral negative RNA strand, which in turn is synthesized by copying the poly(A) sequence of the viral genome, but in this case the latter was absent. It is possible that the polyadenylation was accomplished by the terminal adenylyltransferase activity of the viral RdRP, which was demonstrated to exist in the RdRP of a related poliovirus (339). It also cannot be ruled out that poly(A) was supplied (before the onset of replication?) by a still-undefined cellular mechanism. In any case, the virus was able to survive even such a severe trauma.

An interesting evolutionary trajectory exhibited another engineered genome terminating in the poly(A)-lacking ORF-N111-UCGA sequence (where N111 is a randomized RNA segment) (124). The genome of the virus recovered after transfection was terminated with ORF-N111-UCGAGAAU13AAUAAAAn. Thus, the virus acquired, in addition to poly(A), an AU-rich segment (shown in italics) which contained the AAUAAA cellular polyadenylation signal (underlined) and could be involved in the initial polyadenylation of the engineered genome. The origin of this segment is unknown, but it possibly came from a cellular RNA, e.g., the casein kinase II mRNA, by recombination (replicative or nonreplicative). After further passages, the additional fitness gain was accompanied by further transformation of the genome, which, after 10 passages, acquired a 30-nt segment of the cellular hnRNP U mRNA. Again, recombination was likely responsible for this acquisition. Reappearance of the poly(A) sequence in the engineered tailless genome of hepatitis A virus (having an entirely different oriR structure) was also reported (340).

The results described above again illustrate the diverse tools allowing even severely damaged RNA viruses to significantly or completely regain their fitness.

REHABILITATION AFTER ADVERSE CHANGES IN THE INTERNAL REPLICATIVE CIS-ELEMENT cre (oriI)

A marked inhibition of poliovirus and FMDV genome replication caused by disruption of one or two base pairs in the stems of their cre elements could be restored by one or two naturally occurring second-site mutations, respectively, resulting in the reestablishment of the helical structure (138, 341, 342).

An exceptional and surprising case of true reversions of a set of 16 point mutations in the cre of the engineered RNA of CVB3 was recently published (343). This set was previously reported to be lethal, judging by the inability of the modified genome to induce detectable CPE in transfected cells (147). However, these mutations did not appear to kill the virus, which was still able to replicate in HeLa cells and in some organs of mice, though ∼105-fold less efficiently, and was able to trigger persistent (noncytopathic) infection (149). After 8 days of morphologically unapparent reproduction, reversion of all 16 mutations was detected. Strikingly, no intermediates with only some of the reversions could be detected on previous days, implying that only complete reversion endowed the virus with the ability to win the competition with the mutated variants. The genome of the revertant lacked a number of 5′-terminal nucleotides (i.e., part of oriL), which were lost before the complete set of reversions had been acquired. The mechanism of the structural recovery in this case is unknown, but the possibility that the reversion occurred through recombination with a cryptic, independently noninfectious but intact-cre-element-carrying genome which was present in the quasispecies population should be considered.

REHABILITATION AFTER ADVERSE CHANGES IN VIRAL PROTEINS

The Sabin OPV strains contain attenuation mutations not only in their IRES elements but also in the encoded proteins (321, 344347). Some of them are known to readily revert in vaccinees or during subsequent transmission, resulting in the restoration of viral fitness (reviewed in references 312 and 313). Since such attenuating mutations are differently located in the genomes of the three OPV serotypes, the possibility exists to replace genomic segments containing such mutations with fitness-enhancing homologous segments from another serotype present in the trivalent OPV, endowing recombinants with selective advantages. Indeed, recombination between OPV serotypes is quite common (310, 348352). The pattern of distribution of the crossovers in a large set of such intertypic vaccines/vaccine recombinants allowed us to propose the existence of serotype-specific “weak” (fitness-decreasing) regions which are strongly selected against (353). Intriguingly, the locations of these regions may not necessarily correlate with the locations of the known attenuating mutations, raising questions about the nature of their apparent weakness. It may be added that the Sabin strains can recombine with wild polioviruses and other viruses belonging to the enterovirus C species (354358), although the resulting fitness alterations are not yet adequately characterized.

An unusual case of reversion of a debilitating mutation in the poliovirus 3AB protein was described by de la Torre et al. (359). The engineered C-to-U transition resulting in the Thr67Ile substitution in this protein was accompanied by acquisition of a strong ts phenotype. The reversion of both the nucleotide and the phenotype was accomplished by passaging the mutant at the permissive temperature (33°C). Unexpectedly, three other synonymous mutations (introduced into the investigated genome as markers) also reverted to the wild type in the majority of the revertants. No intermediate genomes possessing only some of these synonymous mutations were detected, to some extent mimicking the above-described collective reversion in the cre element. Again, the involvement of recombination with a cryptic wild-type sequence cannot be ruled out.

Structural and functional impairments of a viral protein may be suppressed by second-site mutations. For example, the progeny of a CVB3 genome with an Asp24Ala mutation in the 3A protein acquired either the true reversion or second-site mutations at position 41 of this protein, restoring its dimerization capacity (360). Invalidation of a viral protein can also be compensated by changes in another viral protein involved in the interaction with the affected one. Thus, certain engineered replacements in protein 2C severely suppressed the production of infectious poliovirus due to impairment of the encapsidation mechanism. However, the fitness was spontaneously restored by either second-site mutations in 2C itself or compensatory mutations in its presumptive ligand, the capsid protein VP3 (237, 361, 362). Fittingly, some FMDV mutants with an alteration of the capsid maturation pathway caused by impaired cleavage at the VP1-2A border were reported to acquire mutations in 2C (240).

Still other rehabilitation variants are exemplified by a poliovirus with a 3-nt insert close to the 3′ terminus of the 3Cpro coding sequence, resulting in apparently complete suppression of viral polyprotein processing at a single cleavage site and rendering the viral genome quasi-infectious (363). Two types of pseudoreversions were detected. The RNA of one recovered virus contained point mutations in both the 3Cpro and 3Dpol coding sequences and also lacked the inserted trinucleotide, whereas this insert in the genome of the other revertant was, surprisingly, replaced by a 15-nt fragment of the rRNA.

It was also demonstrated that relatively short fitness-decreasing inserts generated by nonhomologous replicative and nonreplicative recombination could subsequently be removed by homologous recombination (190, 191, 364, 365).

Postdamage rehabilitation may involve one or more intermediates exhibiting different levels of fitness. Thus, a 12-nt insert in the 2C coding region of poliovirus RNA resulted in the acquisition of a ts phenotype (366). A pseudorevertant able to grow efficiently at the supra-optimal temperature was shown to have two second-site substitutions in 2C. However, this alteration was accompanied by another phenotypic change: reproduction of the virus became cold sensitive. Remarkably, the harvest of the original ts mutant also contained still another pseudorevertant exhibiting temperature dependence similar to that of the wild-type virus. Its 2C coding sequence contained the two mutations present in the cold-sensitive variant and an additional second-site amino acid change, strongly suggesting that it originated from the intermediate cold-sensitive virus (367).

Thus, many adverse amino acid alterations and short indels in poliovirus proteins (and, by implication, those of other picornaviruses) appear to be readily curable by using different rehabilitation mechanisms.

REHABILITATION AFTER LARGE INDELS AND REPLACEMENTS

True spontaneous reversions of large indels may be expected to occur relatively rarely. Nevertheless, an impressive case of very rapid (one cycle of reproduction) and precise deletion of a genomic insertion was observed upon transfection with poliovirus RNA possessing a tandemly duplicated VPg (3B) gene (368). The specific infectivity of the two-VPg RNA was several orders of magnitude lower than that of its wild-type counterpart, indicating that the insertion was nearly lethal. However, progeny of this debilitated genome exhibited the wild-type (i.e., single-VPg) RNA structure, with the 3′-proximal copy of the VPg gene precisely eliminated, perhaps by homologous intra- or intermolecular recombination. Why just the 3′-proximal copy was deleted is unknown. Interestingly, deletion of “additional” copies of the naturally triplicated FMDV VPg gene resulted, in contrast, in a significant fitness loss or death (369).

Though, as noted above, some engineered insertions encoding various useful experimental tools were reasonably well tolerated and proved to be relatively stable, genetic instability and loss of fitness of viruses with some similar inserts were also demonstrated. For example, insertion of several-hundred-nucleotide sequences encoding green fluorescent protein (GFP) or the Gag protein of human immunodeficiency virus between the 5′ UTR and the polyprotein ORF of poliovirus generated either quasi-infectious or low-fit progeny (174). Similar results were obtained after insertion of the luciferase gene at the same position of the RNA (370). These and some other useful engineered inserts were sometimes lost even upon the first passage. The loss was usually imprecise, resulting in distinct partial deletions suggestive of nonhomologous recombination events, and in certain cases, more than one such event may have been involved. The fitness of the partially repaired populations was markedly increased (though not to the wild-type level) after several passages due to the selection of less-damaged variants. More or less similar results were obtained in other studies as well (257). It should be kept in mind that the expression of a foreign sequence may not necessarily be due to the genetic stability of constructs but may also be due to the intrapopulation complementation of defective genomes (371).

A class of genomes with naturally occurring extended deletions, so-called defective interfering (DI) genomes, has long been known for many RNA viruses (372374; for some recent references, see references 375 to 385). For poliovirus, DI genomes are present as a minor component in regular laboratory stocks, but their abundance can be increased significantly upon passage at a high multiplicity of infection (MOI). The deletions usually map to the region encoding capsid proteins and may affect a significant portion of this region (386, 387). Although DI genomes are unable to produce viral particles due to a deficit of capsid proteins, they may be endowed with efficient replicative capacity and may not only ensure self-multiplication but also both interfere with viral growth (as their name implies) and provide some replicative proteins in trans, thereby assisting reproduction of mutant genomes with impaired replicative functions (371). It is reasonable to assume that similar deletions may spontaneously occur in other parts of the viral genome at comparable rates, but since the resultant viruses are replication deficient, they may escape detection. At least some of them may retain replicative capacity owing to complementation by their coreplicating full-length or defective quasispecies members and thus may participate in genetic exchanges. If so, they may provide an additional source of novel genetic material because they may undergo independent, less-constrained evolution. A remarkable illustration of their evolutionary potential was provided by experimental conversion of FMDV into a virus with a bipartite genome. Multiple (>200) tissue culture passages of this virus at a high MOI resulted in the generation of a complex population containing DI genomes with deletions in different genes (388). Due to their mutual complementation, a combination of DI genomes may be propagated at a high MOI in the absence of nondefective helpers, producing lytic infections. Notably, the population with the bipartite FMDV genome demonstrated a higher fitness than that of the parental virus with unsegmented RNA, apparently due to a higher stability of virions encapsidating smaller RNA molecules (389), and certain point mutations accumulated because of the infidelity of RNA replication (390). These observations demonstrated that genome segmentation may represent a mechanism for fitness recovery after genome damages and suggested a model for the origin of picorna-like viruses with a segmented genome (such viruses do indeed exist [391]).

Rehabilitation of fitness impairments caused by deletions in a nonstructural protein may sometimes be accomplished by the acquisition of compensatory mutations in other nonstructural proteins, as exemplified by the restoration of the replicative efficiency of an FMDV mutant lacking a significant part of its 3A protein by a point mutation in 2C (392).

Replacements of genomic regions, which may occur through recombination, are usually regarded as an evolutionary tool for adapting to new or unfavorable conditions or eliminating fitness-decreasing genetic changes (186, 280). However, they may also result in debilitation, and in certain cases such debilitation can be the goal of experimenters. An illuminating example is a poliovirus with the entire IRES exchanged for its counterpart from human rhinovirus type 2 (HRV2) (393). Such viruses, while retaining their poliovirus-like capacity to grow in nonneural human cells, exhibited a strong ts phenotype, rendering them highly inefficient at 37°C in neural (and murine) cells (393, 394). As a result, the chimera proved to be highly attenuated with respect to neurovirulence. The marked interest in such viruses is due to their excellent oncolytic properties, making them quite promising tools for treatment of human tumors of glial origin (395, 396). However, passages of these viruses in neural or murine cells under restrictive conditions resulted in partial recovery of their ability to grow at 37°C due to different sets of mutations, typically including 12- or 13-nt deletions just preceding the IRES as well as certain point mutations within the IRES (394). It was suggested that these modifications enhanced the capacity of the IRES to interact with cell-specific ITAFs, thereby optimizing the efficiency of translation.

In contrast to the poliovirus/HRV2 chimera, replacement of the CVB3 IRES with its counterpart from HRV2 did not result in a marked growth deficiency in neural cells (132). This phenotypic difference appeared to depend on the structural dissimilarity of the poliovirus (enterovirus C) and CVB3 (enterovirus B) oriR regions (Fig. 2C). When stem-loop Z of the CVB3/HRV2 recombinant was deleted (i.e., when the CVB3 oriR was converted into a poliovirus-like one), the capacity of the mutant to grow in neural cells was severely impaired, indicating a functional interdependence of the IRES and oriR. However, genetic determination of the phenotypic properties of these chimeric viruses was more complex. When CVB3 lacking stem-loop Z but possessing the HRV2 IRES was passaged in neural cells, its deficiency for growth in these cells was partially ameliorated due to either mutations in the viral nonstructural proteins 3A/3AB or these mutations in combination with mutations in 3Cpro/3CD (397). This observation indicates the existence of a complex network of functional interactions (epistasis) between different parts of the viral genome, i.e., the IRES, oriR, and several nonstructural proteins, and the possibility to exploit this network for the rehabilitation of genetic injuries.

A different path of rehabilitation after the damaging effect of IRES exchanges was observed when the CVB3 IRES was replaced by a homologous (and structurally similar) region of echovirus 12. This chimera also exhibited a severe host-specific ts phenotype (398). A pseudorevertant of this virus could be selected which regained its fitness through 3 mutations in the IRES, which were suggested to alter its secondary structure.

The above examples illustrate phenotypic improvements of genomes in which exchanges were done between IRES elements of the same structural type (type I). The replacement of the poliovirus IRES by the structurally unrelated IRES of HCV resulted in viable but small-plaque-forming chimeras (285, 399). Passaging one of these viruses resulted in a marked gain of fitness due to selection of mutants with a point substitution or a deletion in the foreign IRES, ensuring its better compatibility with the poliovirus oriL (399).

A poliovirus genome in which an extended part of the 5′ UTR (nt 220 to 627) was replaced with the relevant sequence from CVB3 RNA exhibited a ts phenotype. However, the wild-type level of fitness was regained through spontaneous deletion of 4 nt (positions 231 to 234, i.e., preceding the IRES) (400). For another chimera, in which a 220-nt 5′-terminal segment of CVB3 RNA was replaced by its poliovirus counterpart, reproduction in HeLa cells was suppressed but was restored after the deletion of a tetranucleotide in the same locus (positions 232 to 235) (401). This genome also demonstrated a low fitness in simian cells which, however, was improved upon passaging by the acquisition of two additional second-site mutations. The mechanism(s) underlying these impairments/rehabilitations is unknown.

Important results were obtained with engineered mosaic genomes encoding proteins derived from different viruses. A chimera in which the capsid-encoding part (P1) of CAV20 was replaced by its poliovirus analog proved to be quasi-infectious, and its pseudorevertant small-plaque-forming genomes had acquired single point mutations in either the capsid VP3 or nonstructural 2C protein (370). The combination of these two mutations restored the fitness to a nearly wild-type level. The defect in the original chimera was traced to impaired encapsidation of the viral RNA, strongly supporting the role of the VP3-2A interaction in this process (402).

The above examples again demonstrate the multitude of potential trajectories leading to the improvement of the decreased fitness of injured RNA genomes. Although most of these results were obtained by genetic engineering, natural generation of such chimeras is also quite likely.

REHABILITATION AFTER GENOME DISRUPTION

It seems likely enough that fragmentation of viral RNA resulting from either nucleolytic cleavage or incomplete copying is not infrequent during viral reproduction, but there are no reliable tools to detect, let alone investigate, the fate of the relevant fragments. In any case, RNA fragments that are large enough may serve as recombination partners not only to help the recovery of the disrupted viral genomes but also to fuse with sequences coming from different viruses, or even from viral and cellular RNAs (363, 403405), thereby contributing to viral evolvability.

DEVELOPMENT OF RESISTANCE TO MUTAGENIC AND SOME OTHER INHIBITORS

Although development of viral resistance to inhibitors and rehabilitation after debilitating mutations are formally different topics, the molecular mechanisms underlying these two processes have many common features, since the escape from inhibitory effects of antivirals in some respects mimics restoration of viral functions inflicted by mutations. This is especially true for viral mutagenic inhibitors. Recently, much attention was paid to ribavirin, a purine nucleoside analog which efficiently suppresses a variety of RNA (and DNA) viruses and is widely used in clinical practices (406). A key but not sole mechanism of its activity is its incorporation (after phosphorylation of the respective triphosphate by host enzymes) into viral RNAs by the viral RdRP, mainly in the place of guanosine.

During the synthesis of positive and negative viral RNA strands, ribavirin pairs nearly equally well with cytosine and uracil, resulting in G-to-A and C-to-U mutations, respectively, in viral RNA genomes. Due to accumulation of multiple mutations, incubation of virus-infected cells with this inhibitor may result in complete inactivation of the newly synthesized viral genomes (“error catastrophe”) (41, 407409). However, after multiple passages of poliovirus in the presence of the drug, ribavirin-resistant mutants were isolated, and the mutation responsible for the resistance was traced to 3Dpol replacements, originally Gly64Ser in the case of poliovirus (8). The resistance was due to a significant increase in the fidelity of the mutated polymerase, which became more reluctant to use the inhibitor as a substrate.

Passaging other picornaviruses, such as CVB3 (410), enterovirus 71 (18), and FMDV (26, 411), in the presence of ribavirin or other mutagenic inhibitors (e.g., 5-fluorouracil [FU] or azacytidine) also resulted in the selection of mutants with altered properties (primarily increased fidelity) of the viral RdRP owing to mutations in different loci of the enzyme. Cross-resistance to these inhibitors was also demonstrated (also see a relevant study with HCV [412]). Thus, the rehabilitation of the functionally inefficient genome in the case of mutagenic inhibitors was facilitated just by their mutagenic effect.

RdRP-dependent resistance to a mutagenic inhibitor can be achieved not only through an increase in the general fidelity of this enzyme but also through more fine-tuned changes in its properties (413). Viral replication in the presence of the pyrimidine analog FU is usually accompanied by the accumulation of A-to-G and U-to-C transitions. However, a particular FU-resistant (or rather FU-dependent) FMDV mutant encoded an RdRP with a point mutation able to specifically counteract the acquisition of just these transitions without markedly changing the mutant spectrum complexity in the viral progeny.

The acquisition of resistance to a mutagenic inhibitor may also be accompanied by resistance to inhibitors with different modes of action. Thus, a ribavirin-resistant mutant of CVB3 with a point mutation in the RdRP was also resistant to amiloride (410). The latter drug is not mutagenic but rather affects the intracellular ionic environment. It can be concluded that mutations altering RdRP fidelity may also modify some other properties of this enzyme.

Let us briefly consider the acquisition of resistance to some nonmutagenic inhibitors, which is also due to a single or a few point mutations. The above-mentioned guanidine-resistant (gr) poliovirus mutants are a typical example. The drug targets the viral multifunctional protein 2C (414418), which possesses an RNA-dependent NTPase activity (419, 420) and, in particular, is hypothesized (though not proved) to function as an RNA helicase (421423). Both gr and guanidine-dependent (gd) classes of mutants display a wide range of phenotypes with different levels of dependence of reproductive efficiency on the guanidine concentration (416, 424). The mutants differ with respect to not only the tolerable concentration of the drug but also whether the virus is truly gr (i.e., able to grow equally well in the presence and absence of the drug) or gd (obligatorily requiring the drug for efficient reproduction). As expected, gr and gd mutants display some genetic variability within each class. A comparison of our results (418) with those reported by others revealed certain regularities.

The overwhelming majority of mutants with altered guanidine sensitivity possessed one of two amino acid replacements in the 2C protein (never both): either Asn179 was changed to Gly or Ala, or Met187 was replaced by Leu (for brevity, these mutants belonged to the N or M class, respectively). These mutations are located in the 2C regions thought to be involved in interactions with ATP. Some gr mutants of the N class and at least one gd mutant of the M class did not have any other 2C mutations, but the majority of mutants had additional replacements in other regions of this protein, likely associated with phenotype modulations. Some viruses also contained mutations outside 2C. The acquisition by wild genomes of even a single mutation of the N or M class might sometimes require alterations of two nucleotides (due to the properties of the genetic code). Nevertheless, the presence of more than one mutation does not necessarily mean that multiple consecutive steps were required for the acquisition of altered drug sensitivity, because mutants may well originate from representatives of the quasispecies population with sequences different from the master (prevalent) one. In any case, this set of data clearly demonstrates a multiplicity of trajectories which may allow inefficient genomes to reach fitness peaks.

Interestingly, the same point mutation in FMDV 2C (I268T) confers resistance not only to guanidine (68) but also to ribavirin, ameliorating the mutagenic effect of the latter (425).

It goes without saying that there are a great variety of viral inhibitors with other mechanisms of action and resistance (for example, see reference 426), but a more detailed consideration of the problem of viral drug resistance is outside the scope of this review.

RECOVERY AFTER DEBILITATING BOTTLENECKING

The quasispecies nature of viral populations predicts that various bottlenecking events, which regularly occur during viral replication within organisms and during interhost transmission (43), may often result in more or less significant fitness losses owing to the probability of the presence of adverse mutations in the transmitted genomes (the Muller ratchet). The rehabilitation of RNA viruses after damaging bottlenecking may exploit all the above-discussed mechanisms of regaining fitness, but specifically, this phenomenon in picornaviruses was first studied experimentally by multiple consecutive plaque-to-plaque passages of FMDV (427). This procedure resulted in various levels of fitness losses due to different mutations in different lineages. A low fitness of one of the victims of this ratchet was associated with several point mutations scattered over the genome as well as with a significant extension (heterogeneous in length but, on average, 28 nt long) of a penta-adenylate in the N-terminal region of the polyprotein ORF (428). When this invalidated virus and its four plaque-purified subclones were subjected to serial passages in susceptible cells at different MOIs, they regained their fitness to different extents but, remarkably, exploited different pathways to achieve this. The internal An stretch of the genome was invariably corrected either by true reversion to the original A5, by its shortening, or by an extended deletion of 69 nt that included it. In addition, separate lineages exhibited different sets of true reversions and novel point mutations (including single nucleotide deletions). These observations again illustrate how debilitated viruses can reach different fitness peaks by wandering along individual trajectories.

Notably, even more prolonged plaque-to-plaque transmission did not lead to complete extinction of the FMDV populations (290, 429), suggesting that the natural level of replicative infidelity of this virus is finely tuned to ensure prevention of error catastrophe under such conditions.

SOME ADDITIONAL LESSONS FROM OTHER RNA VIRUSES

Positive-Strand RNA Viruses

As we have seen above, a key factor influencing the rehabilitative capacity of damaged picornavirus genomes is the infidelity of the viral replicative machinery. As far as other viruses with positive-strand RNA are concerned, relatively high (∼10−4) and low (∼10−6) error rates have been reported for bacteriophage Qβ and coronaviruses, respectively (430432). These levels may vary even among different representatives of a particular species (433). The variability depends not only on the peculiarities of the RdRPs but also on the properties of other nonstructural proteins. Thus, the fidelity of alphavirus replication is also modulated by mutations in the helicase/protease (nsP2) (434). Remarkably, the Nidovirales (with the exception of arteriviruses) exhibit a capacity for correcting replicative nucleotide misincorporation that is apparently unique among RNA viruses and is due to the possession of a 3′-to-5′ exoribonuclease, the nonstructural protein nsp14 (433, 435438), which functions in cooperation with another nonstructural protein, nsp10 (439, 440). A decrease in severe acute respiratory syndrome (SARS) coronavirus fidelity due to inactivation of this exoribonuclease was accompanied by a significant loss of viral fitness (441).

The recombination frequency varies as well, for example, being relatively high in nidoviruses (442445) and rather low in flaviviruses (185, 446). Like the situation with picornaviruses, this frequency in other RNA viruses depends on the fidelity of the RdRP (381). On the one hand, low error and recombination rates endow viral genomes with a greater stability, but on the other hand, they diminish their capacity for recovery in the case of damage.

Numerous studies have demonstrated the importance of recombination for the rehabilitation of debilitated positive-strand RNA plant viruses (184), and only some examples are considered here. A segment (RNA3) of the tripartite RNA genome of the cowpea chlorotic mottle virus (a bromovirus) encodes two proteins. Two variants of this segment were constructed, with deletions inactivating one or the other of these proteins. When plants were coinoculated with both deletion variants and intact RNAs 1 and 2, the altered RNA3 was regenerated by recombination (447). The genome of a tobacco etch virus (a potyvirus) with engineered insertions of either its own genes (duplication) or foreign genes (pseudogenization) demonstrated either severe debilitation or even apparent death. However, passages of these low-fit mutants resulted in a more or less rapid enhancement of their reproductive (and competitive) capacity due to the removal of the inserts (448, 449). Inactivating indels in the replicase (i.e., RdRP) gene of phage Qβ could be repaired by homologous recombination, with the intact gene provided in trans by a resident plasmid (450), implying that rehabilitation of even such severe disturbances may occur under natural conditions as well.

In RNA viruses with non-IRES-dependent initiation mechanisms, recovery from significant translational defects may be based on different paths. For example, an engineered 19-nt deletion affecting the hairpins controlling translation of the MS2 phage capsid protein could be ameliorated, upon passages in Escherichia coli, by either deletion of 6 more nucleotides, acquisition of an unrelated 18-nt insert, or both modifications, creating novel functional regulatory structures which included the Shine-Dalgarno sequence (451). A special case of mutational robustness was recently discovered in nidoviruses (452). Proteins of these viruses are largely translated from subgenomic mRNAs (sgRNAs), and the synthesis of each sgRNA is controlled by a distinct cis-element (TRS). By using deep sequencing, numerous overlapping sgRNAs controlled by different TRS were found to encode a given viral protein. This redundancy ensures continued protein synthesis in the case of mutational inactivation of a TRS.

Various molecular mechanisms are operative in nonpicornaviral RNA viruses to repair or compensate for debilitating alterations of their genomic ends. Different pathways for rehabilitation after adverse modifications of the 5′ UTR were demonstrated for Venezuelan equine encephalitis virus (VEEV) or, more precisely, a chimera of VEEV with Sindbis virus (453). The 3′ end of the viral negative RNA strand (serving as the promoter for the synthesis of the positive strand) has a terminal unpaired dinucleotide and adjoining hairpin with a short C/G-rich stem. Mutations introduced into the similarly folded 5′ end of the positive RNA (into either the unpaired AU end or the hairpin's stem) had significant adverse effects on genome replication, but three classes of pseudorevertants were selected on passaging, containing various single-stranded 5′-terminal extensions rich in AUG or AU repeats, new heterologous stem-loops, or mutations in several nonstructural proteins. Interestingly, compensatory mutations in two nonstructural VEEV proteins were also observed after a detrimental (not fatal) extended deletion of two stem-loops of another cis-acting replicative element located not far from the beginning of the ORF (454). Truncation of the 5′-terminal nucleotides of certain plant viruses with positive-strand RNA genomes is also not lethal and can be repaired by various mechanisms, including recombination and nontemplated nucleotide additions by the viral RdRP (455). The RNA of plum pox virus (a potyvirus) is started with A4. This A4 sequence is regained after infection with engineered viral RNAs having either one additional A residue or deletion of one or two A residues at this location (456). It was suggested that the correction could occur during the synthesis of the 3′ end of the negative RNA rather than that of the 5′ end of the positive strand, and thus may involve one of the mechanisms of repair of 3′-terminal sequences considered just below.

Various deletions and insertions in the 3′ UTR as well as complete deletion of the poly(A) tail may not kill alphaviruses (457459), flaviviruses (460462), coronaviruses (463465), and various plant viruses (466469). The repair of the damaged genomes may again involve different mechanisms, including RNA recombination, the use of the viral RdRP- or host-dependent polyadenylation activities, and perhaps some others.

Circularization of flavivirus genomes through interactions between several complementary motifs located at both termini is important for viral reproduction (see reference 470 and references therein). Disruption of some such interactions resulted in a fitness loss, which could be restored at least partially by spontaneous reversions or second-site mutations leading to the restoration/rebuilding of the circularization potential (470, 471).

In positive-strand RNA plant viruses that possess 3′-terminal tRNA-like structures serving as important multifunctional cis-elements (472, 473), aminoacylation of this element is involved in translation and/or replication of the genome, and this reaction requires the presence of the 3′-CCAOH end. However, newly synthesized RNA molecules of these viruses often terminate with 3′-CCOH. Repair of the functional structure is likely accomplished by host enzymes, e.g., [CTP, ATP]:tRNA nucleotidyltransferases (474). Various terminal and internal substitutions and deletions in this element may also be corrected through recombination between distinct components of the multipartite viral genomes, and perhaps by other mechanisms (475478).

Some plant viruses with positive-strand RNA genomes having non-tRNA-like ends are terminated with the 3′-CCCOH sequence. This trinucleotide and adjacent sequences are important for efficient genome replication, but if modified or even absent, they may be spontaneously restored (479482). The rehabilitation can be achieved by implementation of different mechanisms (483). In viruses with multipartite RNA genomes (which sometimes have an additional, so-called satellite RNA), the damaged 3′ terminus of a genomic segment may be repaired by making use of another segment either as the recombination partner or as the template for synthesis of a primer for the initiation of the complement of the defective segment. Also, these viruses possess a functionally important stem-loop structure close to the genomic 3′ end. When the sequence of this element was randomized and the resulting RNA was inoculated into susceptible plants, viable viruses with a marked variety of structures of the relevant stem-loop were selected (484). Truncation of 3′-terminal sequences might result in various rearrangements at this end, such as additional deletions or acquisition of some foreign heterogeneous oligonucleotides, apparently arising through nonhomologous recombination with internal parts of the positive or negative viral RNA strand (485, 486).

Negative-Strand and Double-Stranded RNA Viruses

The fidelity of replication of viruses with negative-strand RNA genomes appears to be comparable to that of positive-strand viruses, varying among different representatives of both groups (487), but the former recombine much less often, and for some of them recombinogenic activity was not demonstrated at all (185, 488, 489). This may be explained at least in part by the use of RNP rather than naked RNA as a transcriptional template (490), which is expected to hamper the template switch and some other mechanisms of formation of chimeric genomes. Nevertheless, true intermolecular recombination in these viruses has continuously been demonstrated (491499). The error rate (487) and frequency of intermolecular recombination of viruses with double-stranded RNA genomes are relatively low due to peculiarities of their replicative machinery, but again, instances of natural recombination between them have been reported (500503).

An interesting example of regaining fitness after an extended in-frame deletion in the coding region of the genome was documented for a spontaneous mutant of influenza A virus (504). This mutant was missing 36 nt in the RNA segment coding for the NS1 protein and exhibited ts and small-plaque phenotypes. Passages of this virus in vitro or in vivo resulted in the generation of pseudorevertants with wild-type properties, the restoration of which was traced to a single amino acid substitution in the same protein without regaining the lost sequence. It was hypothesized that this substitution permitted generation of an alpha-helix element in the NS1 structure, compensating for the deletion-caused loss of the original alpha-helix.

The deficiency caused by a mutation in one viral protein in some cases can be more or less compensated by alterations in other viral proteins. For example, the matrix (M) protein of vesicular stomatitis virus (VSV) has an anti-interferon activity. Deletion of Met51 in this protein rendered the virus interferon sensitive and markedly suppressed its reproduction (505). The fitness was partially restored upon serial passages, and this improvement was attributed to two factors: mutations in another viral protein, phosphoprotein (known to exhibit anti-interferon activity in the rabies virus but not in VSV), and interferon-sensitive members of the quasispecies, which were complemented by their interferon-insensitive counterparts (506). Introduction of 1,378 synonymous mutations in the ORF of the L (RdRP) gene of human respiratory syncytial virus (a pneumovirus) rendered the virus highly temperature sensitive (507). However, passages of its different lineages at stepwise increasing temperatures resulted in the selection of multiple, less-debilitated variants with mutations in various other viral proteins (in 9 of the 11 ORFs) and also in the intergenic regions. The most significant compensatory effects were traced to each of the two alternative amino acid changes in the viral M2-1 antiterminator protein, but the level of rehabilitation could be increased further by combinations of these mutations with alteration in other proteins.

In certain families of viruses with negative-strand and double-stranded RNA (as well as positive-strand RNA), the genomes are composed of several distinct molecules. Such viruses possess an additional rehabilitation tool, reassortment, i.e., exchanges of individual genomic segments between coinfecting viruses (508).

Additional examples of the capacity of RNA viruses to cope with deleterious alterations of their genomes can be found in the review of Barr and Fearns (292).

TO RUN AHEAD, IT IS SOMETIMES USEFUL TO STUMBLE

As we have repeatedly described above, the genomes of RNA viruses are able to maintain their identity under constant conditions despite a significant infidelity of their replicative machinery. In contrast, debilitated genomes are quite unstable. The reason is quite obvious: replication errors are unlikely to increase the fitness of well-adapted viruses but have a much higher chance of being advantageous for weak ones, and the weaker the viruses are, the more different possibilities they have for becoming stronger. To allow them to exploit these possibilities and to fix even slightly advantageous genetic changes, it is necessary not to have more efficient competitors. The invalid viruses very attentively investigate a broad repertoire of options generated by error-prone replication. Point mutations and recombination create a rich swarm of variants, which can follow various evolutionary trajectories. In other words, low-fit viral populations are inherently metastable.

On the one hand, instability opens an array of opportunities for a debilitated virus to regain a wild-type-like genome and thus retain its identity. On the other hand, an important corollary of metastability is a significant potential for the acquisition of qualitatively novel genetic elements, ensuring new phenotypic properties. This constitutes an important basis for viral macroevolution, i.e., generation of novel taxa. Essentially the same events occur if low fitness is caused by changed environmental conditions, e.g., host changes. Indeed, cross-species transmission appears to be a major factor of evolution of RNA viruses (509). A schematic model of these processes is presented in Fig. 7.

FIG 7.

FIG 7

Model of macroevolution of RNA viruses.

Thus, to run ahead, it is sometimes useful to stumble.

ROBUSTNESS, RESILIENCE, AND EVOLVABILITY OF VIRAL RNA GENOMES

The data discussed above illustrate that the maintenance of the genetic and phenotypic identity of RNA viruses, despite the infidelity of their replicative machinery, is principally based on two fundamental properties: robustness and resilience. Robustness can be defined as “the invariance of phenotypes in the face of perturbation” (510). In the coding regions of the genomes, it is based primarily on the codon degeneracy and neutral character of many amino acid substitutions. The robustness of RNA cis-elements is due largely to the degeneracy of RNA spatial structures, i.e., the ability of diverse sequences to maintain, stably or temporarily, similar mutual orientations, as well as the phenotypic neutrality of alterations of certain nucleotides involved in the interactions with specific ligands.

On the other hand, a large proportion of mutations in both coding and noncoding regions of the viral RNA have more or less adverse effects. One may define robustness more broadly and loosely as the capacity to survive in the face of perturbations. As we have seen, certain conserved elements of the IRES and viral proteins can be damaged without a loss of viability. Even all the three key replicative RNA cis-elements of picornaviruses (oriL, cre, and oriR) and the conserved poly(A) tail may not be indispensable for viability. However, in apparent contradiction with this fact, various alterations in these elements were reported to be lethal (52, 119, 121, 136, 341, 511, 512). This discrepancy may represent a manifestation of a significant rule formulated by the Russian author Ilya Il'f: “A watchman was known to check passes very carefully, but those who had no passes at all were allowed to go unquestioned and freely.” Thus, it may sometimes be better to have no pass (e.g., a cis-element) at all than to have a wrong one (513). An explanation of this paradox may consist of the supposition that a complete lack of presumably essential genetic elements may not fully and unconditionally prohibit the relevant wild-type reaction but rather may dramatically decrease its efficiency, directing the process along a different pathway, for example, requiring another protein cofactor or not requiring some normally important protein participants at all. As a result, debilitated viruses can survive. For those with “wrong passes,” such pathways may be prohibited.

Numerous tools are at viral disposal for the restoration of diminished fitness, and the relevant capacity may be called resilience (52) or reparability. Rehabilitation can be achieved by either repair of the injured elements or (“plan B”) the compensatory modification of another viral element. In both instances, the same mechanisms as those that caused debilitation are implemented, i.e., infidelity of the replicative machinery (mutations and recombination). The fixation of the “cured” genome is achieved by selection.

The resilience mechanisms may serve two opposite evolutionary trends. On the one hand, they may result in the full or nearly full restoration (conservation) of the original genome structure and/or phenotype. On the other hand, they are very powerful factors contributing to viral evolvability. Indeed, stepwise acquisition of small improvements upon wandering along rugged fitness landscapes (514), especially in the absence of stronger competitors, may produce a remarkably broad menu of more-fit mutants able to satisfy various viral “tastes” and hence be a major tool for gaining qualitative novelties. The fixation of the achieved results is done by selection, which in both its forms (negative and positive) is one of the key factors contributing to the conservation and evolvability of RNA genomes.

The relationships between mutation rates, quasispecies, robustness, and evolvability of RNA viruses are discussed in detail in numerous publications (487, 515526). This problem is closely related to the emergence/reemergence of pathogenic viruses, which has recently become a hot topic (527532).

Obviously, the regularities discussed in this review are important not only for a deeper understanding of certain fundamental aspects of the lifestyle of RNA viruses but also for numerous applied problems, such as the efficiency of antiviral tools and the development of drug resistance (533537).

ACKNOWLEDGMENTS

Current research by A.P.G. is supported by Russian Science Foundation grant N15-15-00147.

We thank the reviewers for constructive criticism and valuable suggestions.

Biographies

graphic file with name zmr0021824800008.jpg

Vadim I. Agol, M.D., Ph.D., D.Sc., graduated from the 1st Moscow Medical Institute in 1951. In 1956, he joined the Laboratory of Biochemistry of the Institute of Poliomyelitis and Viral Encephalitides (now the M. P. Chumakov Center for Research and Development), where he has worked until now (as Head of the Laboratory from 1961 to 2009). In 1963, he participated in the organization of the Department of Virology of the Moscow State University, where he was Full Professor from 1969 to 2012. He also founded, in 1965, the Department of Virus-Cell Interactions at what is now the A. N. Belozersky Institute of Physical-Chemical Biology of the same university, and he has been its Head until the present. His research interests are in the molecular and cellular biology, evolution, and pathogenicity of RNA viruses, focusing on the translation, replication, recombination, and plasticity of the genomes of picornaviruses as well as on the interactions of these viruses with host defenses and their molecular epidemiology.

graphic file with name zmr0021824800009.jpg

Anatoly P. Gmyl joined the Laboratory of Biochemistry of the M. P. Chumakov Institute of Poliomyelitis and Viral Enchephalitides, headed by Professor Vadim Agol, in 1986, as an undergraduate student of the Moscow Institute of Physics and Technology. He earned M.S. and Ph.D. degrees in 1989 and 1996, respectively. For the first several years of his career, he participated in studies of various aspects of the cap-independent translation of picornavirus RNAs. His major interests then focused on the replication and evolution of RNA-containing viruses. In particular, he was the leading author of the papers describing the discovery and first characterization of nonreplicative RNA recombination. Since 2009, he has been Head of the Laboratory.

REFERENCES

  • 1.Ward CD, Stokes MAM, Flanegan JB. 1988. Direct measurement of the poliovirus RNA-polymerase error frequency in vitro. J Virol 62:558–562. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 2.Arnold JJ, Cameron CE. 2004. Poliovirus RNA-dependent RNA polymerase (3Dpol): pre-steady-state kinetic analysis of ribonucleotide incorporation in the presence of Mg2+. Biochemistry 43:5126–5137. doi: 10.1021/bi035212y. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Arnold JJ, Gohara DW, Cameron CE. 2004. Poliovirus RNA-dependent RNA polymerase (3Dpol): pre-steady-state kinetic analysis of ribonucleotide incorporation in the presence of Mn2+. Biochemistry 43:5138–5148. doi: 10.1021/bi035213q. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Castro C, Arnold JJ, Cameron CE. 2005. Incorporation fidelity of the viral RNA-dependent RNA polymerase: a kinetic, thermodynamic and structural perspective. Virus Res 107:141–149. doi: 10.1016/j.virusres.2004.11.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Freistadt MS, Vaccaro JA, Eberle KE. 2007. Biochemical characterization of the fidelity of poliovirus RNA-dependent RNA polymerase. Virol J 4:44. doi: 10.1186/1743-422X-4-44. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Arias A, Arnold JJ, Sierra M, Smidansky ED, Domingo E, Cameron CE. 2008. Determinants of RNA-dependent RNA polymerase (in)fidelity revealed by kinetic analysis of the polymerase encoded by a foot-and-mouth disease virus mutant with reduced sensitivity to ribavirin. J Virol 82:12346–12355. doi: 10.1128/JVI.01297-08. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Acevedo A, Brodsky L, Andino R. 2014. Mutational and fitness landscapes of an RNA virus revealed through population sequencing. Nature 505:686–690. doi: 10.1038/nature12861. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Pfeiffer JK, Kirkegaard K. 2003. A single mutation in poliovirus RNA-dependent RNA polymerase confers resistance to mutagenic nucleotide analogs via increased fidelity. Proc Natl Acad Sci U S A 100:7289–7294. doi: 10.1073/pnas.1232294100. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Cameron CE, Moustafa IM, Arnold JJ. 2016. Fidelity of nucleotide incorporation by the RNA-dependent RNA polymerase from poliovirus. Enzymes 39:293–323. doi: 10.1016/bs.enz.2016.02.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Borderia AV, Rozen-Gagnon K, Vignuzzi M. 2016. Fidelity variants and RNA quasispecies. Curr Top Microbiol Immunol 392:303–322. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Yang XR, Liu XR, Musser DM, Moustafa IM, Arnold JJ, Cameron CE, Boehr DD. 2017. Triphosphate reorientation of the incoming nucleotide as a fidelity checkpoint in viral RNA-dependent RNA polymerases. J Biol Chem 292:3810–3826. doi: 10.1074/jbc.M116.750638. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Peersen OB. 2017. Picornaviral polymerase structure, function, and fidelity modulation. Virus Res 234:4–20. doi: 10.1016/j.virusres.2017.01.026. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Pfeiffer JK, Kirkegaard K. 2005. Increased fidelity reduces poliovirus fitness and virulence under selective pressure in mice. PLoS Pathog 1:102–110. doi: 10.1371/journal.ppat.0010011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Vignuzzi M, Stone JK, Arnold JJ, Cameron CE, Andino R. 2006. Quasispecies diversity determines pathogenesis through cooperative interactions in a viral population. Nature 439:344–348. doi: 10.1038/nature04388. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Vignuzzi M, Wendt E, Andino R. 2008. Engineering attenuated virus vaccines by controlling replication fidelity. Nat Med 14:154–161. doi: 10.1038/nm1726. [DOI] [PubMed] [Google Scholar]
  • 16.Coffey LL, Beeharry Y, Borderia AV, Blanc H, Vignuzzi M. 2011. Arbovirus high fidelity variant loses fitness in mosquitoes and mice. Proc Natl Acad Sci U S A 108:16038–16043. doi: 10.1073/pnas.1111650108. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Cheung PPH, Watson SJ, Choy KT, Sia SF, Wong DDY, Poon LLM, Kellam P, Guan Y, Peiris JSM, Yen HL. 2014. Generation and characterization of influenza A viruses with altered polymerase fidelity. Nat Commun 5:4794. doi: 10.1038/ncomms5794. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Meng T, Kwang J. 2014. Attenuation of human enterovirus 71 high-replication-fidelity variants in AG129 mice. J Virol 88:5803–5815. doi: 10.1128/JVI.00289-14. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Zeng JX, Wang HW, Xie XC, Li C, Zhou GH, Yang DC, Yu L. 2014. Ribavirin-resistant variants of foot-and-mouth disease virus: the effect of restricted quasispecies diversity on viral virulence. J Virol 88:4008–4020. doi: 10.1128/JVI.03594-13. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Rozen-Gagnon K, Stapleford KA, Mongelli V, Blanc H, Failloux AB, Saleh MC, Vignuzzi M. 2014. Alphavirus mutator variants present host-specific defects and attenuation in mammalian and insect models. PLoS Pathog 10:e1003877. doi: 10.1371/journal.ppat.1003877. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.McDonald S, Block A, Beaucourt S, Moratorio G, Vignuzzi M, Peersen OB. 2016. Design of a genetically stable high fidelity coxsackievirus B3 polymerase that attenuates virus growth in vivo. J Biol Chem 291:13999–14011. doi: 10.1074/jbc.M116.726596. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Xiao YH, Dolan PT, Goldstein EF, Li M, Farkov M, Brodsky L, Andino R. 2017. Poliovirus intrahost evolution is required to overcome tissue-specific innate immune responses. Nat Commun 8:375. doi: 10.1038/s41467-017-00354-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.Gnadig NF, Beaucourt S, Campagnola G, Borderia AV, Sanz-Ramos M, Gong P, Blanc H, Peersen OB, Vignuzzi M. 2012. Coxsackievirus B3 mutator strains are attenuated in vivo. Proc Natl Acad Sci U S A 109:E2294–E2303. doi: 10.1073/pnas.1204022109. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Sadeghipour S, Bek EJ, McMinn PC. 2013. Ribavirin-resistant mutants of human enterovirus 71 express a high replication fidelity phenotype during growth in cell culture. J Virol 87:1759–1769. doi: 10.1128/JVI.02139-12. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Sadeghipour S, McMinn PC. 2013. A study of the virulence in mice of high copying fidelity variants of human enterovirus 71. Virus Res 176:265–272. doi: 10.1016/j.virusres.2013.06.019. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Zeng JX, Wang HW, Xie XC, Yang DC, Zhou GH, Yu L. 2013. An increased replication fidelity mutant of foot-and-mouth disease virus retains fitness in vitro and virulence in vivo. Antiviral Res 100:1–7. doi: 10.1016/j.antiviral.2013.07.008. [DOI] [PubMed] [Google Scholar]
  • 27.Xie XC, Wang HW, Zeng JX, Li C, Zhou GH, Yang DC, Yu L. 2014. Foot-and-mouth disease virus low-fidelity polymerase mutants are attenuated. Arch Virol 159:2641–2650. doi: 10.1007/s00705-014-2126-z. [DOI] [PubMed] [Google Scholar]
  • 28.Griesemer SB, Kramer LD, Van Slyke GA, Pata JD, Gohara DW, Cameron CE, Ciota AT. 2017. Mutagen resistance and mutation restriction of St. Louis encephalitis virus. J Gen Virol 98:201–211. doi: 10.1099/jgv.0.000682. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Rai DK, Diaz-San Segundo F, Campagnola G, Keith A, Schafer EA, Kloc A, de los Santos T, Peersen O, Rieder E. 2017. Attenuation of foot-and-mouth disease virus by engineering viral polymerase fidelity. J Virol 91:e00081-17. doi: 10.1128/JVI.00081-17. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Domingo E, Holland JJ. 1997. RNA virus mutations and fitness for survival. Annu Rev Microbiol 51:151–178. doi: 10.1146/annurev.micro.51.1.151. [DOI] [PubMed] [Google Scholar]
  • 31.Lauring AS, Andino R. 2010. Quasispecies theory and the behavior of RNA viruses. PLoS Pathog 6:e1001005. doi: 10.1371/journal.ppat.1001005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Andino R, Domingo E. 2015. Viral quasispecies. Virology 479:46–51. doi: 10.1016/j.virol.2015.03.022. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Eigen M. 2016. The concept of the quasispecies will soon be 50 years old. Introduction. Curr Top Microbiol Immunol 392:vii. [PubMed] [Google Scholar]
  • 34.Domingo E, Sheldon J, Perales C. 2012. Viral quasispecies evolution. Microbiol Mol Biol Rev 76:159–216. doi: 10.1128/MMBR.05023-11. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Schulte MB, Draghi JA, Plotkin JB, Andino R. 2015. Experimentally guided models reveal replication principles that shape the mutation distribution of RNA viruses. Elife 4:e03753. doi: 10.7554/eLife.03753. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Eigen M. 2002. Error catastrophe and antiviral strategy. Proc Natl Acad Sci U S A 99:13374–13376. doi: 10.1073/pnas.212514799. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Anderson JP, Daifuku R, Loeb LA. 2004. Viral error catastrophe by mutagenic nucleosides. Annu Rev Microbiol 58:183–205. doi: 10.1146/annurev.micro.58.030603.123649. [DOI] [PubMed] [Google Scholar]
  • 38.Domingo E, Escarmis C, Lazaro E, Manrubia SC. 2005. Quasispecies dynamics and RNA virus extinction. Virus Res 107:129–139. doi: 10.1016/j.virusres.2004.11.003. [DOI] [PubMed] [Google Scholar]
  • 39.Grande-Perez A, Lazaro E, Lowenstein P, Domingo E, Manrubia SC. 2005. Suppression of viral infectivity through lethal defection. Proc Natl Acad Sci U S A 102:4448–4452. doi: 10.1073/pnas.0408871102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Summers J, Litwin S. 2006. Examining the theory of error catastrophe. J Virol 80:20–26. doi: 10.1128/JVI.80.1.20-26.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Tejero H, Montero F, Nuno JC. 2016. Theories of lethal mutagenesis: from error catastrophe to lethal defection. Curr Top Microbiol Immunol 392:161–179. doi: 10.1007/82_2015_463. [DOI] [PubMed] [Google Scholar]
  • 42.Manrubia SC, Escarmis C, Domingo E, Lazaro E. 2005. High mutation rates, bottlenecks, and robustness of RNA viral quasispecies. Gene 347:273–282. doi: 10.1016/j.gene.2004.12.033. [DOI] [PubMed] [Google Scholar]
  • 43.Zwart MP, Elena SF. 2015. Matters of size: genetic bottlenecks in virus infection and their potential impact on evolution. Annu Rev Virol 2:161–179. doi: 10.1146/annurev-virology-100114-055135. [DOI] [PubMed] [Google Scholar]
  • 44.Poirier EZ, Vignuzzi M. 2017. Virus population dynamics during infection. Curr Opin Virol 23:82–87. doi: 10.1016/j.coviro.2017.03.013. [DOI] [PubMed] [Google Scholar]
  • 45.Huang S-W, Huang Y-H, Tsai H-P, Kuo P-H, Wang S-M, Liu C-C, Wang J-R. 2017. A selective bottleneck shapes the evolutionary mutant spectra of enterovirus A71 during viral dissemination in humans. J Virol 91:e01062-17. doi: 10.1128/JVI.01062-17. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Chao L. 1990. Fitness of RNA virus decreased by Muller ratchet. Nature 348:454–455. doi: 10.1038/348454a0. [DOI] [PubMed] [Google Scholar]
  • 47.Clarke DK, Duarte EA, Moya A, Elena SF, Domingo E, Holland J. 1993. Genetic bottlenecks and population passages cause profound fitness differences in RNA viruses. J Virol 67:222–228. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Escarmis C, Lazaro E, Manrubia SC. 2006. Population bottlenecks in quasispecies dynamics. Curr Top Microbiol Immunol 299:141–170. [DOI] [PubMed] [Google Scholar]
  • 49.Escarmis C, Perales C, Domingo E. 2009. Biological effect of Muller's ratchet: distant capsid site can affect picornavirus protein processing. J Virol 83:6748–6756. doi: 10.1128/JVI.00538-09. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Burns CC, Shaw J, Campagnoli R, Jorba J, Vincent A, Quay J, Kew O. 2006. Modulation of poliovirus replicative fitness in HeLa cells by deoptimization of synonymous codon usage in the capsid region. J Virol 80:3259–3272. doi: 10.1128/JVI.80.7.3259-3272.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Mueller S, Papamichail D, Coleman JR, Skiena S, Wimmer E. 2006. Reduction of the rate of poliovirus protein synthesis through large-scale codon deoptimization causes attenuation of viral virulence by lowering specific infectivity. J Virol 80:9687–9696. doi: 10.1128/JVI.00738-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Prostova MA, Gmyl AP, Bakhmutov DV, Shishova AA, Khitrina EV, Kolesnikova MS, Serebryakova MV, Isaeva OV, Agol VI. 2015. Mutational robustness and resilience of a replicative cis-element of RNA virus: promiscuity, limitations, relevance. RNA Biol 12:1338–1354. doi: 10.1080/15476286.2015.1100794. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53.Song YT, Gorbatsevych O, Liu Y, Mugavero J, Shen SH, Ward CB, Asare E, Jiang P, Paul AV, Mueller S, Wimmer E. 2017. Limits of variation, specific infectivity, and genome packaging of massively recoded poliovirus genomes. Proc Natl Acad Sci U S A 114:E8731–E8740. doi: 10.1073/pnas.1714385114. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Romanova LI, Blinov VM, Tolskaya EA, Viktorova EG, Kolesnikova MS, Guseva EA, Agol VI. 1986. The primary structure of crossover regions of intertypic poliovirus recombinants: a model of recombination between RNA genomes. Virology 155:202–213. doi: 10.1016/0042-6822(86)90180-7. [DOI] [PubMed] [Google Scholar]
  • 55.Tolskaya EA, Romanova LI, Blinov VM, Viktorova EG, Sinyakov AN, Kolesnikova MS, Agol VI. 1987. Studies on the recombination between RNA genomes of poliovirus: the primary structure and nonrandom distribution of crossover regions in the genomes of intertypic poliovirus recombinants. Virology 161:54–61. doi: 10.1016/0042-6822(87)90170-X. [DOI] [PubMed] [Google Scholar]
  • 56.Racaniello VR, Baltimore D. 1981. Molecular cloning of poliovirus cDNA and determination of the complete nucleotide sequence of the viral genome. Proc Natl Acad Sci U S A 78:4887–4891. doi: 10.1073/pnas.78.8.4887. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Kitamura N, Semler BL, Rothberg PG, Larsen GR, Adler CJ, Dorner AJ, Emini EA, Hanecak R, Lee JJ, Vanderwerf S, Anderson CW, Wimmer E. 1981. Primary structure, gene organization and polypeptide expression of poliovirus RNA. Nature 291:547–553. doi: 10.1038/291547a0. [DOI] [PubMed] [Google Scholar]
  • 58.Nomoto A, Omata T, Toyoda H, Kuge S, Horie H, Kataoka Y, Genba Y, Nakano Y, Imura N. 1982. Complete nucleotide sequence of the attenuated poliovirus Sabin 1 strain genome. Proc Natl Acad Sci U S A 79:5793–5797. doi: 10.1073/pnas.79.19.5793. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59.Stanway G, Cann AJ, Hauptmann R, Hughes P, Clarke LD, Mountford RC, Minor PD, Schild GC, Almond JW. 1983. The nucleotide sequence of poliovirus type 3 Leon 12 a1b: comparison with poliovirus type 1. Nucleic Acids Res 11:5629–5643. doi: 10.1093/nar/11.16.5629. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Toyoda H, Kohara M, Kataoka Y, Suganuma T, Omata T, Imura N, Nomoto A. 1984. Complete nucleotide sequences of all three poliovirus serotype genomes: implication for genetic relationship, gene function and antigenic determinants. J Mol Biol 174:561–585. doi: 10.1016/0022-2836(84)90084-6. [DOI] [PubMed] [Google Scholar]
  • 61.Melnick JL, Crowther D, Barrera-Oro J. 1961. Rapid development of drug-resistant mutants of poliovirus. Science 134:557. doi: 10.1126/science.134.3478.557. [DOI] [PubMed] [Google Scholar]
  • 62.Loddo B, Ferrari W, Spanedda A, Brotzu G. 1962. In vitro guanidino-resistance and guanidino-dependence of poliovirus. Experientia 18:518–519. doi: 10.1007/BF02151608. [DOI] [PubMed] [Google Scholar]
  • 63.Eggers HJ, Tamm I. 1961. Spectrum and characteristics of the virus inhibitory action of 2-(α-hydroxybenzyl)-benzimidazole. J Exp Med 113:657–682. doi: 10.1084/jem.113.4.657. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.Eggers HJ, Tamm I. 1963. Drug dependence of enteroviruses: variants of coxsackie A9 and ECHO 13 viruses that require 2-(α-hydroxybenzyl)-benzimidazole for growth. Virology 20:62–74. doi: 10.1016/0042-6822(63)90141-7. [DOI] [Google Scholar]
  • 65.Loddo B. 1980. Development of drug resistance and dependence in viruses. Pharmacol Ther 10:431–460. doi: 10.1016/0163-7258(80)90026-1. [DOI] [PubMed] [Google Scholar]
  • 66.Heinz BA, Rueckert RR, Shepard DA, Dutko FJ, McKinlay MA, Fancher M, Rossmann MG, Badger J, Smith TJ. 1989. Genetic and molecular analyses of spontaneous mutants of human rhinovirus 14 that are resistant to an antiviral compound. J Virol 63:2476–2485. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67.Wang W, Lee WM, Mosser AG, Rueckert RR. 1998. WIN 52035-dependent human rhinovirus 16: assembly deficiency caused by mutations near the canyon surface. J Virol 72:1210–1218. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68.Pariente N, Airaksinen A, Domingo E. 2003. Mutagenesis versus inhibition in the efficiency of extinction of foot-and-mouth disease virus. J Virol 77:7131–7138. doi: 10.1128/JVI.77.12.7131-7138.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69.Liu H-M, Roberts JA, Moore D, Anderson B, Pallansch MA, Pevear DC, Collett MS, Oberste MS. 2012. Characterization of poliovirus variants selected for resistance to the antiviral compound V-073. Antimicrob Agents Chemother 56:5568–5574. doi: 10.1128/AAC.00539-12. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70.Stoyanova A, Nikolova I, Purstinger G, Dobrikov G, Dimitrov V, Philipov S, Galabov AS. 2015. Anti-enteroviral triple combination of viral replication inhibitors: activity against coxsackievirus B1 neuroinfection in mice. Antivir Chem Chemother 24:136–147. doi: 10.1177/2040206616671571. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.Collett MS, Hincks JR, Benschop K, Duizer E, van der Avoort H, Rhoden E, Liu HM, Oberste MS, McKinlay MA, Hartford M. 2017. Antiviral activity of pocapavir in a randomized, blinded, placebo-controlled human oral poliovirus vaccine challenge model. J Infect Dis 215:335–343. doi: 10.1093/infdis/jiw542. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72.Crowder S, Kirkegaard K. 2005. Trans-dominant inhibition of RNA viral replication can slow growth of drug-resistant viruses. Nat Genet 37:701–709. doi: 10.1038/ng1583. [DOI] [PubMed] [Google Scholar]
  • 73.Perales C, Mateo R, Mateu MG, Domingo E. 2007. Insights into RNA virus mutant spectrum and lethal mutagenesis events: replicative interference and complementation by multiple point mutants. J Mol Biol 369:985–1000. doi: 10.1016/j.jmb.2007.03.074. [DOI] [PubMed] [Google Scholar]
  • 74.Agol VI. 2002. Picornavirus genome: an overview, p 127–148. In Semler BL, Wimmer E (ed), Molecular biology of picornaviruses. ASM Press, Washington, DC. [Google Scholar]
  • 75.Palmenberg A, Neubauer D, Skern T. 2010. Genome organization and encoded proteins, p 3–17. In Ehrenfeld E, Domingo E, Roos RP (ed), The picornaviruses. ASM Press, Washington, DC. [Google Scholar]
  • 76.Agol VI, Gmyl AP. 2010. Viral security proteins: counteracting host defences. Nat Rev Microbiol 8:867–878. doi: 10.1038/nrmicro2452. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77.Boros A, Pankovics P, Reuter G. 2014. Avian picornaviruses: molecular evolution, genome diversity and unusual genome features of a rapidly expanding group of viruses in birds. Infect Genet Evol 28:151–166. doi: 10.1016/j.meegid.2014.09.027. [DOI] [PubMed] [Google Scholar]
  • 78.Kong WP, Roos RP. 1991. Alternative translation initiation site in the DA strain of Theiler's murine encephalomyelitis virus. J Virol 65:3395–3399. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 79.Loughran G, Firth AE, Atkins JF. 2011. Ribosomal frameshifting into an overlapping gene in the 2B-encoding region of the cardiovirus genome. Proc Natl Acad Sci U S A 108:E1111–E1119. doi: 10.1073/pnas.1102932108. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 80.Finch LK, Ling R, Napthine S, Olspert A, Michiels T, Lardinois C, Bell S, Loughran G, Brierley I, Firth AE. 2015. Characterization of ribosomal frameshifting in Theiler's murine encephalomyelitis virus. J Virol 89:8580–8589. doi: 10.1128/JVI.01043-15. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 81.Woo PC, Lau SK, Choi GK, Huang Y, Teng JL, Tsoi HW, Tse H, Yeung ML, Chan KH, Jin DY, Yuen KY. 2012. Natural occurrence and characterization of two internal ribosome entry site elements in a novel virus, canine picodicistrovirus, in the picornavirus-like superfamily. J Virol 86:2797–2808. doi: 10.1128/JVI.05481-11. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 82.Reuter G, Boros A, Foldvari G, Szekeres S, Matics R, Kapusinszky B, Delwart E, Pankovics P. 2018. Dicipivirus (family Picornaviridae) in wild Northern white-breasted hedgehog (Erinaceus roumanicus). Arch Virol 163:175–181. doi: 10.1007/s00705-017-3565-0. [DOI] [PubMed] [Google Scholar]
  • 83.Shang PC, Misra S, Hause B, Fang Y. 2017. A naturally occurring recombinant enterovirus expresses a torovirus deubiquitinase. J Virol 91:e00450-17. doi: 10.1128/JVI.00450-17. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 84.Knutson TP, Velayudhan BT, Marthaler DG. 2017. A porcine enterovirus G associated with enteric disease contains a novel papain-like cysteine protease. J Gen Virol 98:1305–1310. doi: 10.1099/jgv.0.000799. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 85.Boros A, Pankovics P, Knowles NJ, Nemes C, Delwart E, Reuter G. 2014. Comparative complete genome analysis of chicken and turkey megriviruses (family Picornaviridae): long 3′ untranslated regions with a potential second open reading frame and evidence for possible recombination. J Virol 88:6434–6443. doi: 10.1128/JVI.03807-13. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 86.Liu Y, Wimmer E, Paul AV. 2009. Cis-acting RNA elements in human and animal plus-strand RNA viruses. Biochim Biophys Acta 1789:495–517. doi: 10.1016/j.bbagrm.2009.09.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 87.Martinez-Salas E, Francisco-Velilla R, Fernandez-Chamorro J, Lozano G, Diaz-Toledano R. 2015. Picornavirus IRES elements: RNA structure and host protein interactions. Virus Res 206:62–73. doi: 10.1016/j.virusres.2015.01.012. [DOI] [PubMed] [Google Scholar]
  • 88.Rivera VM, Welsh JD, Maizel JV. 1988. Comparative sequence analysis of the 5′ noncoding region of the enteroviruses and rhinoviruses. Virology 165:42–50. doi: 10.1016/0042-6822(88)90656-3. [DOI] [PubMed] [Google Scholar]
  • 89.Andino R, Rieckhof GE, Baltimore D. 1990. A functional ribonucleoprotein complex forms around the 5′ end of poliovirus RNA. Cell 63:369–380. doi: 10.1016/0092-8674(90)90170-J. [DOI] [PubMed] [Google Scholar]
  • 90.Andino R, Rieckhof GE, Achacoso PL, Baltimore D. 1993. Poliovirus RNA synthesis utilizes an RNP complex formed around the 5′-end of viral RNA. EMBO J 12:3587–3598. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 91.Harris KS, Xiang WK, Alexander L, Lane WS, Paul AV, Wimmer E. 1994. Interaction of poliovirus polypeptide 3CDpro with the 5′-termini and 3′-termini of the poliovirus genome—identification of viral and cellular cofactors needed for efficient binding. J Biol Chem 269:27004–27014. [PubMed] [Google Scholar]
  • 92.Ogram SA, Flanegan JB. 2011. Non-template functions of viral RNA in picornavirus replication. Curr Opin Virol 1:339–346. doi: 10.1016/j.coviro.2011.09.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 93.Leveque N, Garcia M, Bouin A, Nguyen JHC, Tran GP, Andreoletti L, Semler BL. 2017. Functional consequences of RNA 5′-terminal deletions on coxsackievirus B3 RNA replication and ribonucleoprotein complex formation. J Virol 91:e00423-17. doi: 10.1128/JVI.00423-17. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 94.Andino R, Rieckhof GE, Trono D, Baltimore D. 1990. Substitutions in the protease (3Cpro) gene of poliovirus can suppress a mutation in the 5′ noncoding region. J Virol 64:607–612. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 95.Xiang WK, Harris KS, Alexander L, Wimmer E. 1995. Interaction between the 5′-terminal cloverleaf and 3AB/3CDpro of poliovirus is essential for RNA replication. J Virol 69:3658–3667. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 96.Barton DJ, O'Donnell BJ, Flanegan JB. 2001. 5′ cloverleaf in poliovirus RNA is a cis-acting replication element required for negative-strand synthesis. EMBO J 20:1439–1448. doi: 10.1093/emboj/20.6.1439. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 97.Lyons T, Murray KE, Roberts AW, Barton DJ. 2001. Poliovirus 5′-terminal cloverleaf RNA is required in cis for VPg uridylylation and the initiation of negative-strand RNA synthesis. J Virol 75:10696–10708. doi: 10.1128/JVI.75.22.10696-10708.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 98.Vogt DA, Andino R. 2010. An RNA element at the 5′-end of the poliovirus genome functions as a general promoter for RNA synthesis. PLoS Pathog 6:e1000936. doi: 10.1371/journal.ppat.1000936. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 99.Jaramillo L, Smithee S, Tracy S, Chapman NM. 2016. Domain I of the 5′ non-translated genomic region in coxsackievirus B3 RNA is not required for productive replication. Virology 496:127–130. doi: 10.1016/j.virol.2016.05.021. [DOI] [PubMed] [Google Scholar]
  • 100.Jang SK, Krausslich HG, Nicklin MJH, Duke GM, Palmenberg AC, Wimmer E. 1988. A segment of the 5′ nontranslated region of encephalomyocarditis virus RNA directs internal entry of ribosomes during in vitro translation. J Virol 62:2636–2643. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 101.Trono D, Pelletier J, Sonenberg N, Baltimore D. 1988. Translation in mammalian cells of a gene linked to the poliovirus 5′ noncoding region. Science 241:445–448. doi: 10.1126/science.2839901. [DOI] [PubMed] [Google Scholar]
  • 102.Pelletier J, Sonenberg N. 1988. Internal initiation of translation of eukaryotic mRNA directed by a sequence derived from poliovirus RNA. Nature 334:320–325. doi: 10.1038/334320a0. [DOI] [PubMed] [Google Scholar]
  • 103.Pelletier J, Sonenberg N. 1989. Internal binding of eukaryotic ribosomes on poliovirus RNA: translation in HeLa cell extracts. J Virol 63:441–444. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 104.Pilipenko EV, Blinov VM, Romanova LI, Sinyakov AN, Maslova SV, Agol VI. 1989. Conserved structural domains in the 5′-untranslated region of picornaviral genomes: an analysis of the segment controlling translation and neurovirulence. Virology 168:201–209. doi: 10.1016/0042-6822(89)90259-6. [DOI] [PubMed] [Google Scholar]
  • 105.Pilipenko EV, Blinov VM, Chernov BK, Dmitrieva TM, Agol VI. 1989. Conservation of the secondary structure elements of the 5′-untranslated region of cardiovirus and aphthovirus RNAs. Nucleic Acids Res 17:5701–5711. doi: 10.1093/nar/17.14.5701. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 106.Jang SK. 2006. Internal initiation: IRES elements of picornaviruses and hepatitis C virus. Virus Res 119:2–15. doi: 10.1016/j.virusres.2005.11.003. [DOI] [PubMed] [Google Scholar]
  • 107.Niepmann M. 2009. Internal translation initiation of picornaviruses and hepatitis C virus. Biochim Biophys Acta 1789:529–541. doi: 10.1016/j.bbagrm.2009.05.002. [DOI] [PubMed] [Google Scholar]
  • 108.Sweeney TR, Dhote V, Yu YP, Hellen CUT. 2012. A distinct class of internal ribosomal entry site in members of the Kobuvirus and proposed Salivirus and Paraturdivirus genera of the Picornaviridae. J Virol 86:1468–1486. doi: 10.1128/JVI.05862-11. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 109.Lee KM, Chen CJ, Shih SR. 2017. Regulation mechanisms of viral IRES-driven translation. Trends Microbiol 25:546–561. doi: 10.1016/j.tim.2017.01.010. [DOI] [PubMed] [Google Scholar]
  • 110.Yamamoto H, Unbehaun A, Spahn CMT. 2017. Ribosomal chamber music: toward an understanding of IRES mechanisms. Trends Biochem Sci 42:655–668. doi: 10.1016/j.tibs.2017.06.002. [DOI] [PubMed] [Google Scholar]
  • 111.Hellen CUT, de Breyne S. 2007. A distinct group of hepacivirus/pestivirus-like internal ribosomal entry sites in members of diverse picornavirus genera: evidence for modular exchange of functional noncoding RNA elements by recombination. J Virol 81:5850–5863. doi: 10.1128/JVI.02403-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 112.Lozano G, Martinez-Salas E. 2015. Structural insights into viral IRES-dependent translation mechanisms. Curr Opin Virol 12:113–120. doi: 10.1016/j.coviro.2015.04.008. [DOI] [PubMed] [Google Scholar]
  • 113.Pestova TV, Kolupaeva VG, Lomakin IB, Pilipenko EV, Shatsky IN, Agol VI, Hellen CUT. 2001. Molecular mechanisms of translation initiation in eukaryotes. Proc Natl Acad Sci U S A 98:7029–7036. doi: 10.1073/pnas.111145798. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 114.Pilipenko EV, Pestova TV, Kolupaeva VG, Khitrina EV, Poperechnaya AN, Agol VI, Hellen CUT. 2000. A cell cycle-dependent protein serves as a template-specific translation initiation factor. Genes Dev 14:2028–2045. [PMC free article] [PubMed] [Google Scholar]
  • 115.Pilipenko EV, Maslova SV, Sinyakov AN, Agol VI. 1992. Towards identification of cis-acting elements involved in the replication of enterovirus and rhinovirus RNAs: a proposal for the existence of tRNA-like terminal structures. Nucleic Acids Res 20:1739–1745. doi: 10.1093/nar/20.7.1739. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 116.Pilipenko EV, Poperechny KV, Maslova SV, Melchers WJG, Bruins Slot HJ, Agol VI. 1996. Cis-element, oriR, involved in the initiation of (−) strand poliovirus RNA: a quasi-globular multi-domain RNA structure maintained by tertiary (‘kissing') interactions. EMBO J 15:5428–5436. [PMC free article] [PubMed] [Google Scholar]
  • 117.Rohll JB, Percy N, Ley R, Evans DJ, Almond JW, Barclay WS. 1994. The 5′-untranslated regions of picornavirus RNAs contain independent functional domains essential for RNA replication and translation. J Virol 68:4384–4391. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 118.Todd S, Semler BL. 1996. Structure-infectivity analysis of the human rhinovirus genomic RNA 3′ non-coding region. Nucleic Acids Res 24:2133–2142. doi: 10.1093/nar/24.11.2133. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 119.Melchers WJG, Hoenderop JGJ, Bruins Slot HJ, Pleij CWA, Pilipenko EV, Agol VI, Galama JMD. 1997. Kissing of the two predominant hairpin loops in the coxsackie B virus 3′ untranslated region is the essential structural feature of the origin of replication required for negative-strand RNA synthesis. J Virol 71:686–696. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 120.Mirmomeni MH, Hughes PJ, Stanway G. 1997. An RNA tertiary structure in the 3′ untranslated region of enteroviruses is necessary for efficient replication. J Virol 71:2363–2370. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 121.Wang JH, Bakkers JM, Galama JM, Bruins Slot HJ, Pilipenko EV, Agol VI, Melchers WJG. 1999. Structural requirements of the higher order RNA kissing element in the enteroviral 3′ UTR. Nucleic Acids Res 27:485–490. doi: 10.1093/nar/27.2.485. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 122.Zoll J, Heus HA, van Kuppeveld FJM, Melchers WJG. 2009. The structure-function relationship of the enterovirus 3′-UTR. Virus Res 139:209–216. doi: 10.1016/j.virusres.2008.07.014. [DOI] [PubMed] [Google Scholar]
  • 123.Brown DM, Cornell CT, Tran GP, Nguyen JHC, Semler BL. 2005. An authentic 3′ noncoding region is necessary for efficient poliovirus replication. J Virol 79:11962–11973. doi: 10.1128/JVI.79.18.11962-11973.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 124.van Ooij MJM, Polacek C, Glaudemans DHRF, Kuijpers J, van Kuppeveld FJM, Andino R, Agol VI, Melchers WJG. 2006. Polyadenylation of genomic RNA and initiation of antigenomic RNA in a positive-strand RNA virus are controlled by the same cis-element. Nucleic Acids Res 34:2953–2965. doi: 10.1093/nar/gkl349. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 125.Herold J, Andino R. 2001. Poliovirus RNA replication requires genome circularization through a protein-protein bridge. Mol Cell 7:581–591. doi: 10.1016/S1097-2765(01)00205-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 126.Svitkin YV, Imataka H, Khaleghpour K, Kahvejian A, Liebig HD, Sonenberg N. 2001. Poly(A)-binding protein interactions with eIF4G stimulates picornavirus IRES-dependent translations. RNA 7:1743–1752. doi: 10.1017/S135583820100108X. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 127.Serrano P, Pulido MR, Saiz M, Martinez-Salas E. 2006. The 3′ end of the foot-and-mouth disease virus genome establishes two distinct long-range RNA-RNA interactions with the 5′ end region. J Gen Virol 87:3013–3022. doi: 10.1099/vir.0.82059-0. [DOI] [PubMed] [Google Scholar]
  • 128.Ogram SA, Spear A, Sharma N, Flanegan JB. 2010. The 5′ CL-PCBP RNP complex, 3′ poly(A) tail and 2Apro are required for optimal translation of poliovirus RNA. Virology 397:14–22. doi: 10.1016/j.virol.2009.11.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 129.Verma B, Bhattacharyya S, Das S. 2010. Polypyrimidine tract-binding protein interacts with coxsackievirus B3 RNA and influences its translation. J Gen Virol 91:1245–1255. doi: 10.1099/vir.0.018507-0. [DOI] [PubMed] [Google Scholar]
  • 130.Diaz-Toledano R, Lozano G, Martinez-Salas E. 2017. In-cell SHAPE uncovers dynamic interactions between the untranslated regions of the foot-and-mouth disease virus RNA. Nucleic Acids Res 45:1416–1432. doi: 10.1093/nar/gkw795. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 131.Merkle I, van Ooij MJM, van Kuppeveld FJM, Glaudemans DHRF, Galama JMD, Henke A, Zell R, Melchers WJG. 2002. Biological significance of a human enterovirus B-specific RNA element in the 3′ nontranslated region. J Virol 76:9900–9909. doi: 10.1128/JVI.76.19.9900-9909.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 132.Dobrikova E, Florez P, Bradrick S, Gromeier M. 2003. Activity of a type 1 picornavirus internal ribosomal entry site is determined by sequences within the 3′ nontranslated region. Proc Natl Acad Sci U S A 100:15125–15130. doi: 10.1073/pnas.2436464100. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 133.Brown DM, Kauder SE, Cornell CT, Jang GM, Racaniello VR, Semler BL. 2004. Cell-dependent role for the poliovirus 3′ noncoding region in positive-strand RNA synthesis. J Virol 78:1344–1351. doi: 10.1128/JVI.78.3.1344-1351.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 134.Todd S, Towner JS, Brown DM, Semler BL. 1997. Replication-competent picornaviruses with complete genomic RNA 3′ noncoding region deletions. J Virol 71:8868–8874. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 135.Duque H, Palmenberg AC. 2001. Phenotypic characterization of three phylogenetically conserved stem-loop motifs in the mengovirus 3′ untranslated region. J Virol 75:3111–3120. doi: 10.1128/JVI.75.7.3111-3120.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 136.McKnight KL, Lemon SM. 1998. The rhinovirus type 14 genome contains an internally located RNA structure that is required for viral replication. RNA 4:1569–1584. doi: 10.1017/S1355838298981006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 137.Lobert PE, Escriou N, Ruelle J, Michiels T. 1999. A coding RNA sequence acts as a replication signal in cardioviruses. Proc Natl Acad Sci U S A 96:11560–11565. doi: 10.1073/pnas.96.20.11560. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 138.Goodfellow I, Chaudhry Y, Richardson A, Meredith J, Almond JW, Barclay W, Evans DJ. 2000. Identification of a cis-acting replication element within the poliovirus coding region. J Virol 74:4590–4600. doi: 10.1128/JVI.74.10.4590-4600.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 139.Goodfellow IG, Kerrigan D, Evans DJ. 2003. Structure and function analysis of the poliovirus cis-acting replication element (CRE). RNA 9:124–137. doi: 10.1261/rna.2950603. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 140.Thiviyanathan V, Yang Y, Kaluarachchi K, Rijnbrand R, Gorenstein DG, Lemon SM. 2004. High-resolution structure of a picornaviral internal cis-acting RNA replication element (cre). Proc Natl Acad Sci U S A 101:12688–12693. doi: 10.1073/pnas.0403079101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 141.Steil BP, Barton DJ. 2009. Cis-active RNA elements (CREs) and picornavirus RNA replication. Virus Res 139:240–252. doi: 10.1016/j.virusres.2008.07.027. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 142.Paul AV, Rieder E, Kim DW, van Boom JH, Wimmer E. 2000. Identification of an RNA hairpin in poliovirus RNA that serves as the primary template in the in vitro uridylylation of VPg. J Virol 74:10359–10370. doi: 10.1128/JVI.74.22.10359-10370.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 143.Yang Y, Rijnbrand R, McKnight KL, Wimmer E, Paul A, Martin A, Lemon SM. 2002. Sequence requirements for viral RNA replication and VPg uridylylation directed by the internal cis-acting replication element (cre) of human rhinovirus type 14. J Virol 76:7485–7494. doi: 10.1128/JVI.76.15.7485-7494.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 144.Murray KE, Barton DJ. 2003. Poliovirus CRE-dependent VPg uridylylation is required for positive-strand RNA synthesis but not for negative-strand RNA synthesis. J Virol 77:4739–4750. doi: 10.1128/JVI.77.8.4739-4750.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 145.Morasco BJ, Sharma N, Parilla J, Flanegan JB. 2003. Poliovirus cre(2C)-dependent synthesis of VPgpUpU is required for positive- but not negative-strand RNA synthesis. J Virol 77:5136–5144. doi: 10.1128/JVI.77.9.5136-5144.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 146.Goodfellow IG, Polacek C, Andino R, Evans DJ. 2003. The poliovirus 2C cis-acting replication element-mediated uridylylation of VPg is not required for synthesis of negative-sense genomes. J Gen Virol 84:2359–2363. doi: 10.1099/vir.0.19132-0. [DOI] [PubMed] [Google Scholar]
  • 147.van Ooij MJM, Vogt DA, Paul A, Castro C, Kuijpers J, van Kuppeveld FJM, Cameron CE, Wimmer E, Andino R, Melchers WJG. 2006. Structural and functional characterization of the coxsackievirus B3 CRE(2C): role of CRE(2C) in negative- and positive-strand RNA synthesis. J Gen Virol 87:103–113. doi: 10.1099/vir.0.81297-0. [DOI] [PubMed] [Google Scholar]
  • 148.Paul AV, Wimmer E. 2015. Initiation of protein-primed picornavirus RNA synthesis. Virus Res 206:12–26. doi: 10.1016/j.virusres.2014.12.028. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 149.Smithee S, Tracy S, Chapman NM. 2015. Mutational disruption of cis-acting replication element 2C in coxsackievirus B3 leads to 5′-terminal genomic deletions. J Virol 89:11761–11772. doi: 10.1128/JVI.01308-15. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 150.Kim KS, Tracy S, Tapprich W, Bailey J, Lee CK, Kim K, Barry WH, Chapman NM. 2005. 5′-terminal deletions occur in coxsackievirus B3 during replication in murine hearts and cardiac myocyte cultures and correlate with encapsidation of negative-strand viral RNA. J Virol 79:7024–7041. doi: 10.1128/JVI.79.11.7024-7041.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 151.Kim KS, Chapman NM, Tracy S. 2008. Replication of coxsackievirus B3 in primary cell cultures generates novel viral genome deletions. J Virol 82:2033–2037. doi: 10.1128/JVI.01774-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 152.Chapman NM, Kim KS, Drescher KM, Oka K, Tracy S. 2008. 5′ terminal deletions in the genome of a coxsackievirus B2 strain occurred naturally in human heart. Virology 375:480–491. doi: 10.1016/j.virol.2008.02.030. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 153.Bouin A, Nguyen Y, Wehbe M, Renois F, Fornes P, Bani-Sadr F, Metz D, Andreoletti L. 2016. Major persistent 5′ terminally deleted coxsackievirus B3 populations in human endomyocardial tissues. Emerg Infect Dis 22:1488–1490. doi: 10.3201/eid2208.160186. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 154.Han JQ, Townsend HL, Jha BK, Paranjape JM, Silverman RH, Barton DJ. 2007. A phylogenetically conserved RNA structure in the poliovirus open reading frame inhibits the antiviral endoribonuclease RNase L. J Virol 81:5561–5572. doi: 10.1128/JVI.01857-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 155.Keel AY, Jha BK, Kieft JS. 2012. Structural architecture of an RNA that competitively inhibits RNase L. RNA 18:88–99. doi: 10.1261/rna.030007.111. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 156.Witwer C, Rauscher S, Hofacker IL, Stadler PF. 2001. Conserved RNA secondary structures in Picornaviridae genomes. Nucleic Acids Res 29:5079–5089. doi: 10.1093/nar/29.24.5079. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 157.Song YT, Liu Y, Ward CB, Mueller S, Futcher B, Skiena S, Paul AV, Wimmer E. 2012. Identification of two functionally redundant RNA elements in the coding sequence of poliovirus using computer-generated design. Proc Natl Acad Sci U S A 109:14301–14307. doi: 10.1073/pnas.1211484109. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 158.Burrill CP, Westesson O, Schulte MB, Strings VR, Segal M, Andino R. 2013. Global RNA structure analysis of poliovirus identifies a conserved RNA structure involved in viral replication and infectivity. J Virol 87:11670–11683. doi: 10.1128/JVI.01560-13. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 159.Kempf BJ, Barton DJ. 2015. Picornavirus RNA polyadenylation by 3Dpol, the viral RNA-dependent RNA polymerase. Virus Res 206:3–11. doi: 10.1016/j.virusres.2014.12.030. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 160.Steil BP, Kempf BJ, Barton DJ. 2010. Poly(A) at the 3′ end of positive-strand RNA and VPg-linked poly(U) at the 5′ end of negative-strand RNA are reciprocal templates during replication of poliovirus RNA. J Virol 84:2843–2858. doi: 10.1128/JVI.02620-08. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 161.Spector DH, Baltimore D. 1975. Polyadenylic-acid on poliovirus RNA. IV. Poly(U) in replicative intermediate and double-stranded RNA. Virology 67:498–505. [DOI] [PubMed] [Google Scholar]
  • 162.Larsen GR, Dorner AJ, Harris TJR, Wimmer E. 1980. The structure of poliovirus replicative form. Nucleic Acids Res 8:1217–1229. doi: 10.1093/nar/8.6.1217. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 163.Duchene S, Holmes EC, Ho SYW. 2014. Analyses of evolutionary dynamics in viruses are hindered by a time-dependent bias in rate estimates. Proc Biol Sci 281:20140732. doi: 10.1098/rspb.2014.0732. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 164.Jorba J, Campagnoli R, De L, Kew O. 2008. Calibration of multiple poliovirus molecular clocks covering an extended evolutionary range. J Virol 82:4429–4440. doi: 10.1128/JVI.02354-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 165.Koonin EV, Wolf YI, Nagasaki K, Dolja VV. 2008. The Big Bang of picorna-like virus evolution antedates the radiation of eukaryotic supergroups. Nat Rev Microbiol 6:925–939. doi: 10.1038/nrmicro2030. [DOI] [PubMed] [Google Scholar]
  • 166.Goodman RA, Macbeth MR, Beal PA. 2012. ADAR proteins: structure and catalytic mechanism. Curr Top Microbiol Immunol 353:1–33. doi: 10.1007/82_2011_144. [DOI] [PubMed] [Google Scholar]
  • 167.Deffit SN, Hundley HA. 2016. To edit or not to edit: regulation of ADAR editing specificity and efficiency. Wiley Interdiscip Rev RNA 7:113–127. doi: 10.1002/wrna.1319. [DOI] [PubMed] [Google Scholar]
  • 168.Salter JD, Bennett RP, Smith HC. 2016. The APOBEC protein family: united by structure, divergent in function. Trends Biochem Sci 41:578–594. doi: 10.1016/j.tibs.2016.05.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 169.Moris A, Murray S, Cardinaud S. 2014. AID and APOBECs span the gap between innate and adaptive immunity. Front Microbiol 5:534. doi: 10.3389/fmicb.2014.00534. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 170.Tomaselli S, Galeano F, Locatelli F, Gallo A. 2015. ADARs and the balance game between virus infection and innate immune cell response. Curr Issues Mol Biol 17:37–51. [PubMed] [Google Scholar]
  • 171.Pilipenko EV, Gmyl AP, Maslova SV, Svitkin YV, Sinyakov AN, Agol VI. 1992. Prokaryotic-like cis elements in the cap-independent internal initiation of translation on picornavirus RNA. Cell 68:119–131. doi: 10.1016/0092-8674(92)90211-T. [DOI] [PubMed] [Google Scholar]
  • 172.Gmyl AP, Pilipenko EV, Maslova SV, Belov GA, Agol VI. 1993. Functional and genetic plasticities of the poliovirus genome: quasi-infectious RNAs modified in the 5′-untranslated region yield a variety of pseudorevertants. J Virol 67:6309–6316. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 173.Lee SG, Kim DY, Hyun BH, Bae YS. 2002. Novel design architecture for genetic stability of recombinant poliovirus: the manipulation of G/C contents and their distribution patterns increases the genetic stability of inserts in a poliovirus-based RPS-Vax vector system. J Virol 76:1649–1662. doi: 10.1128/JVI.76.4.1649-1662.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 174.Mueller S, Wimmer E. 1998. Expression of foreign proteins by poliovirus polyprotein fusion: analysis of genetic stability reveals rapid deletions and formation of cardioviruslike open reading frames. J Virol 72:20–31. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 175.Miller JP, Geng Y, Ng HL, Yang OO, Krogstad P. 2009. Packaging limits and stability of HIV-1 sequences in a coxsackievirus B vector. Vaccine 27:3992–4000. doi: 10.1016/j.vaccine.2009.04.035. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 176.Seago J, Juleff N, Moffat K, Berryman S, Christie JM, Charleston B, Jackson T. 2013. An infectious recombinant foot-and-mouth disease virus expressing a fluorescent marker protein. J Gen Virol 94:1517–1527. doi: 10.1099/vir.0.052308-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 177.Cooper PD. 1968. A genetic map of poliovirus temperature-sensitive mutants. Virology 35:584–596. doi: 10.1016/0042-6822(68)90287-0. [DOI] [PubMed] [Google Scholar]
  • 178.Cooper PD. 1977. Genetics of picornaviruses, p 133–207. In Fraenkel-Conrat H, Wagner RR (ed), Comprehensive virology, vol 9 Springer US, Boston, MA. [Google Scholar]
  • 179.Copper PD, Steiner-Pryor A, Scotti PD, Delong D. 1974. On the nature of poliovirus genetic recombinants. J Gen Virol 23:41–49. doi: 10.1099/0022-1317-23-1-41. [DOI] [PubMed] [Google Scholar]
  • 180.Agol VI. 2010. Picornaviruses as a model for studying the nature of RNA recombination, p 239–252. In Domingo E, Ehrenfeld E, Roos RP (ed), Picornaviruses: molecular biology, evolution, and pathogenesis. ASM Press, Washington, DC. [Google Scholar]
  • 181.Kirkegaard K, Baltimore D. 1986. The mechanism of RNA recombination in poliovirus. Cell 47:433–443. doi: 10.1016/0092-8674(86)90600-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 182.Nagy PD, Simon AE. 1997. New insights into the mechanisms of RNA recombination. Virology 235:1–9. doi: 10.1006/viro.1997.8681. [DOI] [PubMed] [Google Scholar]
  • 183.Arnold JJ, Cameron CE. 1999. Poliovirus RNA-dependent RNA polymerase (3Dpol) is sufficient for template switching in vitro. J Biol Chem 274:2706–2716. doi: 10.1074/jbc.274.5.2706. [DOI] [PubMed] [Google Scholar]
  • 184.Sztuba-Solinska J, Urbanowicz A, Figlerowicz M, Bujarski JJ. 2011. RNA-RNA recombination in plant virus replication and evolution. Annu Rev Phytopathol 49:415–443. doi: 10.1146/annurev-phyto-072910-095351. [DOI] [PubMed] [Google Scholar]
  • 185.Simon-Loriere E, Holmes EC. 2011. Why do RNA viruses recombine? Nat Rev Microbiol 9:617–626. doi: 10.1038/nrmicro2614. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 186.Xiao YH, Rouzine IM, Bianco S, Acevedo A, Goldstein EF, Farkov M, Brodsky L, Andino R. 2016. RNA recombination enhances adaptability and is required for virus spread and virulence. Cell Host Microbe 19:493–503. doi: 10.1016/j.chom.2016.03.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 187.Woodman A, Arnold JJ, Cameron CE, Evans DJ. 2016. Biochemical and genetic analysis of the role of the viral polymerase in enterovirus recombination. Nucleic Acids Res 44:6883–6895. doi: 10.1093/nar/gkw567. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 188.Kempf BJ, Peersen OB, Barton DJ. 2016. Poliovirus polymerase Leu420 facilitates RNA recombination and ribavirin resistance. J Virol 90:8410–8421. doi: 10.1128/JVI.00078-16. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 189.Gmyl AP, Belousov EV, Maslova SV, Khitrina EV, Chetverin AB, Agol VI. 1999. Nonreplicative RNA recombination in poliovirus. J Virol 73:8958–8965. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 190.Gmyl AP, Korshenko SA, Belousov EV, Khitrina EV, Agol VI. 2003. Nonreplicative homologous RNA recombination: promiscuous joining of RNA pieces? RNA 9:1221–1231. doi: 10.1261/rna.5111803. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 191.Holmblat B, Jegouic S, Muslin C, Blondel B, Joffret ML, Delpeyroux F. 2014. Nonhomologous recombination between defective poliovirus and coxsackievirus genomes suggests a new model of genetic plasticity for picornaviruses. mBio 5:e01119-14. doi: 10.1128/mBio.01119-14. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 192.Schibler M, Piuz I, Hao WD, Tapparel C. 2015. Chimeric rhinoviruses obtained via genetic engineering or artificially induced recombination are viable only if the polyprotein coding sequence derives from the same species. J Virol 89:4470–4480. doi: 10.1128/JVI.03668-14. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 193.Gallei A, Pankraz A, Thiel HJ, Becher P. 2004. RNA recombination in vivo in the absence of viral replication. J Virol 78:6271–6281. doi: 10.1128/JVI.78.12.6271-6281.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 194.Austermann-Busch S, Becher P. 2012. RNA structural elements determine frequency and sites of nonhomologous recombination in an animal plus-strand RNA virus. J Virol 86:7393–7402. doi: 10.1128/JVI.00864-12. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 195.Scheel TKH, Galli A, Li YP, Mikkelsen LS, Gottwein JM, Bukh J. 2013. Productive homologous and non-homologous recombination of hepatitis C virus in cell culture. PLoS Pathog 9:e1003228. doi: 10.1371/journal.ppat.1003228. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 196.Kleine Buning M, Meyer D, Austermann-Busch S, Roman-Sosa G, Rumenapf T, Becher P. 2017. Nonreplicative RNA recombination of an animal plus-strand RNA virus in the absence of efficient translation of viral proteins. Genome Biol Evol 9:817–829. doi: 10.1093/gbe/evx046. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 197.Gmyl AP, Agol VI. 2005. Diverse mechanisms of RNA recombination. Mol Biol 39:529–542. doi: 10.1007/s11008-005-0069-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 198.Raju R, Subramaniam SV, Hajjou M. 1995. Genesis of Sindbis virus by in vivo recombination of nonreplicative RNA precursors. J Virol 69:7391–7401. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 199.Adams SD, Tzeng WP, Chen MH, Frey TK. 2003. Analysis of intermolecular RNA-RNA recombination by rubella virus. Virology 309:258–271. doi: 10.1016/S0042-6822(03)00064-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 200.Ullmer W, Semler BL. 2016. Diverse strategies used by picornaviruses to escape host RNA decay pathways. Viruses 8:335. doi: 10.3390/v8120335. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 201.Perez-Ortin JE, Alepuz P, Chavez S, Choder M. 2013. Eukaryotic mRNA decay: methodologies, pathways, and links to other stages of gene expression. J Mol Biol 425:3750–3775. doi: 10.1016/j.jmb.2013.02.029. [DOI] [PubMed] [Google Scholar]
  • 202.Drappier M, Michiels T. 2015. Inhibition of the OAS/RNase L pathway by viruses. Curr Opin Virol 15:19–26. doi: 10.1016/j.coviro.2015.07.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 203.Ariza-Mateos A, Prieto-Vega S, Diaz-Toledano R, Birk A, Szeto H, Mena I, Berzal-Herranz A, Gomez J. 2012. RNA self-cleavage activated by ultraviolet light-induced oxidation. Nucleic Acids Res 40:1748–1766. doi: 10.1093/nar/gkr822. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 204.Neme R, Amador C, Yildirim B, McConnell E, Tautz D. 2017. Random sequences are an abundant source of bioactive RNAs or peptides. Nat Ecol Evol 1:0217. doi: 10.1038/s41559-017-0127. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 205.Sanjuan R, Moya A, Elena SF. 2004. The distribution of fitness effects caused by single-nucleotide substitutions in an RNA virus. Proc Natl Acad Sci U S A 101:8396–8401. doi: 10.1073/pnas.0400146101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 206.Carrasco P, de la Iglesia F, Elena SF. 2007. Distribution of fitness and virulence effects caused by single-nucleotide substitutions in tobacco etch virus. J Virol 81:12979–12984. doi: 10.1128/JVI.00524-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 207.Domingo-Calap P, Cuevas JM, Sanjuan R. 2009. The fitness effects of random mutations in single-stranded DNA and RNA bacteriophages. PLoS Genet 5:e1000742. doi: 10.1371/journal.pgen.1000742. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 208.Visher E, Whitefield SE, McCrone JT, Fitzsimmons W, Lauring AS. 2016. The mutational robustness of influenza A virus. PLoS Pathog 12:e1005856. doi: 10.1371/journal.ppat.1005856. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 209.Coleman JR, Papamichail D, Skiena S, Futcher B, Wimmer E, Mueller S. 2008. Virus attenuation by genome-scale changes in codon pair bias. Science 320:1784–1787. doi: 10.1126/science.1155761. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 210.Lauring AS, Acevedo A, Cooper SB, Andino R. 2012. Codon usage determines the mutational robustness, evolutionary capacity, and virulence of an RNA virus. Cell Host Microbe 12:623–632. doi: 10.1016/j.chom.2012.10.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 211.Diaz-San Segundo F, Medina GN, Ramirez-Medina E, Velazquez-Salinas L, Koster M, Grubman MJ, de los Santos T. 2016. Synonymous deoptimization of foot-and-mouth disease virus causes attenuation in vivo while inducing a strong neutralizing antibody response. J Virol 90:1298–1310. doi: 10.1128/JVI.02167-15. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 212.Kjar J, Belsham GJ. 2018. Selection of functional 2A sequences within foot-and-mouth disease virus; requirements for the NPGP motif with a distinct codon bias. RNA 24:12–17. doi: 10.1261/rna.063339.117. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 213.Cuevas JM, Domingo-Calap P, Sanjuan R. 2012. The fitness effects of synonymous mutations in DNA and RNA viruses. Mol Biol Evol 29:17–20. doi: 10.1093/molbev/msr179. [DOI] [PubMed] [Google Scholar]
  • 214.Usami A, Mochizuki T, Tsuda S, Ohki ST. 2013. Large-scale codon de-optimisation of the p29 replicase gene by synonymous substitutions causes a loss of infectivity of melon necrotic spot virus. Arch Virol 158:1979–1985. doi: 10.1007/s00705-013-1683-x. [DOI] [PubMed] [Google Scholar]
  • 215.Shen SH, Stauft CB, Gorbatsevych O, Song YT, Ward CB, Yurovsky A, Mueller S, Futcher B, Wimmer E. 2015. Large-scale recoding of an arbovirus genome to rebalance its insect versus mammalian preference. Proc Natl Acad Sci U S A 112:4749–4754. doi: 10.1073/pnas.1502864112. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 216.Gao L, Wang LH, Huang C, Yang LL, Guo XK, Yu ZB, Liu YH, Yang P, Feng WH. 2015. HP-PRRSV is attenuated by de-optimization of codon pair bias in its RNA-dependent RNA polymerase nsp9 gene. Virology 485:135–144. doi: 10.1016/j.virol.2015.07.012. [DOI] [PubMed] [Google Scholar]
  • 217.Martinez MA, Jordan-Paiz A, Franco S, Nevot M. 2016. Synonymous virus genome recoding as a tool to impact viral fitness. Trends Microbiol 24:134–147. doi: 10.1016/j.tim.2015.11.002. [DOI] [PubMed] [Google Scholar]
  • 218.Mueller S, Coleman JR, Papamichail D, Ward CB, Nimnual A, Futcher B, Skiena S, Wimmer E. 2010. Live attenuated influenza virus vaccines by computer-aided rational design. Nat Biotechnol 28:723–726. doi: 10.1038/nbt.1636. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 219.Meng J, Lee S, Hotard AL, Moore ML. 2014. Refining the balance of attenuation and immunogenicity of respiratory syncytial virus by targeted codon deoptimization of virulence genes. mBio 5:e01704-14. doi: 10.1128/mBio.01704-14. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 220.Le Nouën C, Brock LG, Luongo C, McCarty T, Yang L, Mehedi M, Wimmer E, Mueller S, Collins PL, Buchholz UJ, DiNapoli JM. 2014. Attenuation of human respiratory syncytial virus by genome-scale codon-pair deoptimization. Proc Natl Acad Sci U S A 111:13169–13174. doi: 10.1073/pnas.1411290111. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 221.Fan RLY, Valkenburg SA, Wong CKS, Li OTW, Nicholls JM, Rabadan R, Peiris JSM, Poon LLM. 2015. Generation of live attenuated influenza virus by using codon usage bias. J Virol 89:10762–10773. doi: 10.1128/JVI.01443-15. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 222.Wang B, Yang C, Tekes G, Mueller S, Paul A, Whelan SPJ, Wimmer E. 2015. Recoding of the vesicular stomatitis virus L gene by computer-aided design provides a live, attenuated vaccine candidate. mBio 6:e00237-15. doi: 10.1128/mBio.00237-15. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 223.Broadbent AJ, Santos CP, Anafu A, Wimmer E, Mueller S, Subbarao K. 2016. Evaluation of the attenuation, immunogenicity, and efficacy of a live virus vaccine generated by codon-pair bias de-optimization of the 2009 pandemic H1N1 influenza virus, in ferrets. Vaccine 34:563–570. doi: 10.1016/j.vaccine.2015.11.054. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 224.Chevance FFV, Hughes KT. 2017. Case for the genetic code as a triplet of triplets. Proc Natl Acad Sci U S A 114:4745–4750. doi: 10.1073/pnas.1614896114. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 225.Jacobs WM, Shakhnovich EI. 2017. Evidence of evolutionary selection for cotranslational folding. Proc Natl Acad Sci U S A 114:11434–11439. doi: 10.1073/pnas.1705772114. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 226.McMullan LK, Grakoui A, Evans MJ, Mihalik K, Puig M, Branch AD, Feinstone SM, Rice CM. 2007. Evidence for a functional RNA element in the hepatitis C virus core gene. Proc Natl Acad Sci U S A 104:2879–2884. doi: 10.1073/pnas.0611267104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 227.Jiang B, Monroe SS, Koonin EV, Stine SE, Glass RI. 1993. RNA sequence of astrovirus: distinctive genomic organization and a putative retrovirus-like ribosomal frameshifting signal that directs the viral replicase synthesis. Proc Natl Acad Sci U S A 90:10539–10543. doi: 10.1073/pnas.90.22.10539. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 228.Xu Z, Choi J, Yen TS, Lu W, Strohecker A, Govindarajan S, Chien D, Selby MJ, Ou J. 2001. Synthesis of a novel hepatitis C virus protein by ribosomal frameshift. EMBO J 20:3840–3848. doi: 10.1093/emboj/20.14.3840. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 229.Kolakofsky D. 2016. Paramyxovirus RNA synthesis, mRNA editing, and genome hexamer phase: a review. Virology 498:94–98. doi: 10.1016/j.virol.2016.08.018. [DOI] [PubMed] [Google Scholar]
  • 230.Moomau C, Musalgaonkar S, Khan YA, Jones JE, Dinman JD. 2016. Structural and functional characterization of programmed ribosomal frameshift signals in West Nile virus strains reveals high structural plasticity among cis-acting RNA elements. J Biol Chem 291:15788–15795. doi: 10.1074/jbc.M116.735613. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 231.Kendra JA, de la Fuente C, Brahms A, Woodson C, Bell TM, Chen B, Khan YA, Jacobs JL, Kehn-Hall K, Dinman JD. 2017. Ablation of programmed −1 ribosomal frameshifting in Venezuelan equine encephalitis virus results in attenuated neuropathogenicity. J Virol 91:e01766-16. doi: 10.1128/JVI.01766-16. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 232.Burns CC, Campagnoli R, Shaw J, Vincent A, Jorba J, Kew O. 2009. Genetic inactivation of poliovirus infectivity by increasing the frequencies of CpG and UpA dinucleotides within and across synonymous capsid region codons. J Virol 83:9957–9969. doi: 10.1128/JVI.00508-09. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 233.Atkinson NJ, Witteveldt J, Evans DJ, Simmonds P. 2014. The influence of CpG and UpA dinucleotide frequencies on RNA virus replication and characterization of the innate cellular pathways underlying virus attenuation and enhanced replication. Nucleic Acids Res 42:4527–4545. doi: 10.1093/nar/gku075. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 234.Stern A, Yeh MT, Zinger T, Smith M, Wright C, Ling G, Nielsen R, Macadam A, Andino R. 2017. The evolutionary pathway to virulence of an RNA virus. Cell 169:35.e19–46.e19. doi: 10.1016/j.cell.2017.03.013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 235.Takata MA, Goncalves-Carneiro D, Zang TM, Soll SJ, York A, Blanco-Melo D, Bieniasz PD. 2017. CG dinucleotide suppression enables antiviral defence targeting non-self RNA. Nature 550:124–127. doi: 10.1038/nature24039. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 236.Fros JJ, Dietrich I, Alshaikhahmed K, Passchier TC, Evans DJ, Simmonds P. 2017. CpG and UpA dinucleotides in both coding and non-coding regions of echovirus 7 inhibit replication initiation post-entry. Elife 6:e29112. doi: 10.7554/eLife.29112. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 237.Wang CL, Jiang P, Sand C, Paul AV, Wimmer E. 2012. Alanine scanning of poliovirus 2CATPase reveals new genetic evidence that capsid protein/2CATPase interactions are essential for morphogenesis. J Virol 86:9964–9975. doi: 10.1128/JVI.00914-12. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 238.Dmitrieva TM, Alexeevski AV, Shatskaya GS, Tolskaya EA, Gmyl AP, Khitrina EV, Agol VI. 2007. Significance of the C-terminal amino acid residue in mengovirus RNA-dependent RNA polymerase. Virology 365:79–91. doi: 10.1016/j.virol.2007.02.038. [DOI] [PubMed] [Google Scholar]
  • 239.Gullberg M, Polacek C, Botner A, Belsham GJ. 2013. Processing of the VP1/2A junction is not necessary for production of foot-and-mouth disease virus empty capsids and infectious viruses: characterization of “self-tagged” particles. J Virol 87:11591–11603. doi: 10.1128/JVI.01863-13. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 240.Kristensen T, Normann P, Gullberg M, Fahnoe U, Polacek C, Rasmussen TB, Belsham GJ. 2017. Determinants of the VP1/2A junction cleavage by the 3C protease in foot-and-mouth disease virus infected cells. J Gen Virol 98:385–395. doi: 10.1099/jgv.0.000664. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 241.Cheong CJ, Varani G, Tinoco I. 1990. Solution structure of an unusually stable RNA hairpin, 5′ggac(UUCG)gucc. Nature 346:680–682. doi: 10.1038/346680a0. [DOI] [PubMed] [Google Scholar]
  • 242.Varani G, Cheong CJ, Tinoco I. 1991. Structure of an unusually stable RNA hairpin. Biochemistry 30:3280–3289. doi: 10.1021/bi00227a016. [DOI] [PubMed] [Google Scholar]
  • 243.Melchers WJG, Zoll J, Tessari M, Bakhmutov DV, Gmyl AP, Agol VI, Heus HA. 2006. A GCUA tetranucleotide loop found in the poliovirus oriL by in vivo SELEX (un)expectedly forms a YNMG-like structure: extending the YNMG family with GYYA. RNA 12:1671–1682. doi: 10.1261/rna.113106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 244.Kuge S, Nomoto A. 1987. Construction of viable deletion and insertion mutants of the Sabin strain of type 1 poliovirus: function of the 5′ noncoding sequence in viral replication. J Virol 61:1478–1487. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 245.Slobodskaya OR, Gmyl AP, Maslova SV, Tolskaya EA, Viktorova EG, Agol VI. 1996. Poliovirus neurovirulence correlates with the presence of a cryptic AUG upstream of the initiator codon. Virology 221:141–150. doi: 10.1006/viro.1996.0360. [DOI] [PubMed] [Google Scholar]
  • 246.Haller AA, Semler BL. 1992. Linker scanning mutagenesis of the internal ribosome entry site of poliovirus RNA. J Virol 66:5075–5086. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 247.Burke KL, Evans DJ, Jenkins O, Meredith J, D’Souza EDA, Almond JW. 1989. A cassette vector for the construction of antigen chimeras of poliovirus. J Gen Virol 70:2475–2479. doi: 10.1099/0022-1317-70-9-2475. [DOI] [PubMed] [Google Scholar]
  • 248.Dedieu JF, Ronco J, van der Werf S, Hogle JM, Henin Y, Girard M. 1992. Poliovirus chimeras expressing sequences from the principal neutralization domain of human immunodeficiency virus type 1. J Virol 66:3161–3167. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 249.Lemon SM, Barclay W, Ferguson M, Murphy P, Jing L, Burke K, Wood D, Katrak K, Sangar D, Minor PD, Almond JW. 1992. Immunogenicity and antigenicity of chimeric picornaviruses which express hepatitis A virus (HAV) peptide sequences: evidence for a neutralization domain near the amino terminus of VP1 of HAV. Virology 188:285–295. doi: 10.1016/0042-6822(92)90758-H. [DOI] [PubMed] [Google Scholar]
  • 250.Andino R, Silvera D, Suggett SD, Achacoso PL, Miller CJ, Baltimore D, Feinberg MB. 1994. Engineering poliovirus as a vaccine vector for the expression of diverse antigens. Science 265:1448–1451. doi: 10.1126/science.8073288. [DOI] [PubMed] [Google Scholar]
  • 251.Mattion NM, Reilly PA, DiMichele SJ, Crowley JC, Weeks-Levy C. 1994. Attenuated poliovirus strain as a live vector: expression of regions of rotavirus outer capsid protein VP7 by using recombinant Sabin 3 viruses. J Virol 68:3925–3933. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 252.Mattion NM, Reilly PA, Camposano E, Wu SL, DiMichele SJ, Ishizaka ST, Fantini SE, Crowley JC, Weeks-Levy C. 1995. Characterization of recombinant polioviruses expressing regions of rotavirus VP4, hepatitis B surface antigen, and herpes simplex virus type 2 glycoprotein D. J Virol 69:5132–5137. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 253.Yim TJ, Tang SB, Andino R. 1996. Poliovirus recombinants expressing hepatitis B virus antigens elicited a humoral immune response in susceptible mice. Virology 218:61–70. doi: 10.1006/viro.1996.0166. [DOI] [PubMed] [Google Scholar]
  • 254.Tang SB, van Rij R, Silvera D, Andino R. 1997. Toward a poliovirus-based simian immunodeficiency virus vaccine: correlation between genetic stability and immunogenicity. J Virol 71:7841–7850. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 255.Cho SP, Lee B, Min MK. 2000. Recombinant polioviruses expressing hepatitis B virus-specific cytotoxic T-lymphocyte epitopes. Vaccine 18:2878–2885. doi: 10.1016/S0264-410X(00)00060-8. [DOI] [PubMed] [Google Scholar]
  • 256.Crotty S, Miller CJ, Lohman BL, Neagu MR, Compton L, Lu D, Lu FXS, Fritts L, Lifson JD, Andino R. 2001. Protection against simian immunodeficiency virus vaginal challenge by using Sabin poliovirus vectors. J Virol 75:7435–7452. doi: 10.1128/JVI.75.16.7435-7452.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 257.Dobrikova EY, Florez P, Gromeier M. 2003. Structural determinants of insert retention of poliovirus expression vectors with recombinant IRES elements. Virology 311:241–253. doi: 10.1016/S0042-6822(03)00191-0. [DOI] [PubMed] [Google Scholar]
  • 258.Kim DS, Cho YL, Kim BG, Lee SH, Nam JH. 2010. Systematic analysis of attenuated coxsackievirus expressing a foreign gene as a viral vaccine vector. Vaccine 28:1234–1240. doi: 10.1016/j.vaccine.2009.11.017. [DOI] [PubMed] [Google Scholar]
  • 259.Teterina NL, Lauber C, Jensen KS, Levenson EA, Gorbalenya AE, Ehrenfeld E. 2011. Identification of tolerated insertion sites in poliovirus non-structural proteins. Virology 409:1–11. doi: 10.1016/j.virol.2010.09.028. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 260.Teterina NL, Levenson EA, Ehrenfeld E. 2010. Viable polioviruses that encode 2A proteins with fluorescent protein tags. J Virol 84:1477–1488. doi: 10.1128/JVI.01578-09. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 261.Seago J, Jackson T, Doel C, Fry E, Stuart D, Harmsen MM, Charleston B, Juleff N. 2012. Characterization of epitope-tagged foot-and-mouth disease virus. J Gen Virol 93:2371–2381. doi: 10.1099/vir.0.043521-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 262.Park JN, Ko MK, Kim RH, Park ME, Lee SY, Yoon JE, Choi JH, You SH, Park JW, Lee KN, Chun JE, Kim SM, Tark D, Lee HS, Ko YJ, Kim B, Lee MH, Park JH. 2016. Construction of stabilized and tagged foot-and-mouth disease virus. J Virol Methods 237:187–191. doi: 10.1016/j.jviromet.2016.09.013. [DOI] [PubMed] [Google Scholar]
  • 263.Caine EA, Osorio JE. 2017. In vivo imaging with bioluminescent enterovirus 71 allows for real-time visualization of tissue tropism and viral spread. J Virol 91:e01759-16. doi: 10.1128/JVI.01759-16. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 264.Zhang F, Perez-Martin E, Juleff N, Charleston B, Seago J. 2017. A replication-competent foot-and-mouth disease virus expressing a luciferase reporter. J Virol Methods 247:38–44. doi: 10.1016/j.jviromet.2017.05.011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 265.Molla A, Jang SK, Paul AV, Reuer Q, Wimmer E. 1992. Cardioviral internal ribosomal entry site is functional in a genetically engineered dicistronic poliovirus. Nature 356:255–257. doi: 10.1038/356255a0. [DOI] [PubMed] [Google Scholar]
  • 266.Molla A, Paul AV, Schmid M, Jang SK, Wimmer E. 1993. Studies on dicistronic polioviruses implicate viral proteinase 2Apro in RNA replication. Virology 196:739–747. doi: 10.1006/viro.1993.1531. [DOI] [PubMed] [Google Scholar]
  • 267.Borman AM, Deliat FG, Kean KM. 1994. Sequences within the poliovirus internal ribosome entry segment control viral RNA synthesis. EMBO J 13:3149–3157. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 268.Alexander L, Lu HH, Wimmer E. 1994. Polioviruses containing picornavirus type 1 and/or type 2 internal ribosomal entry site elements: genetic hybrids and the expression of a foreign gene. Proc Natl Acad Sci U S A 91:1406–1410. doi: 10.1073/pnas.91.4.1406. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 269.Lu HH, Alexander L, Wimmer E. 1995. Construction and genetic analysis of dicistronic polioviruses containing open reading frames for epitopes of human immunodeficiency virus type 1 gp120. J Virol 69:4797–4806. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 270.Cao XM, Wimmer E. 1995. Intragenomic complementation of a 3AB mutant in dicistronic polioviruses. Virology 209:315–326. doi: 10.1006/viro.1995.1263. [DOI] [PubMed] [Google Scholar]
  • 271.Behura M, Mohapatra JK, Pandey LK, Das B, Bhatt M, Subramaniam S, Pattnaik B. 2016. The carboxy-terminal half of nonstructural protein 3A is not essential for foot-and-mouth disease virus replication in cultured cell lines. Arch Virol 161:1295–1305. doi: 10.1007/s00705-016-2805-z. [DOI] [PubMed] [Google Scholar]
  • 272.Baranowski E, Ruiz-Jarabo CM, Lim F, Domingo E. 2001. Foot-and-mouth disease virus lacking the VP1 G-H loop: the mutant spectrum uncovers interactions among antigenic sites for fitness gain. Virology 288:192–202. doi: 10.1006/viro.2001.1096. [DOI] [PubMed] [Google Scholar]
  • 273.Lawrence P, Pacheco JM, Uddowla S, Hollister J, Kotecha A, Fry E, Rieder E. 2013. Foot-and-mouth disease virus (FMDV) with a stable FLAG epitope in the VP1 G-H loop as a new tool for studying FMDV pathogenesis. Virology 436:150–161. doi: 10.1016/j.virol.2012.11.001. [DOI] [PubMed] [Google Scholar]
  • 274.Escarmis C, Dopazo J, Davila M, Palma EL, Domingo E. 1995. Large deletions in the 5′-untranslated region of foot-and-mouth-disease virus of serotype C. Virus Res 35:155–167. doi: 10.1016/0168-1702(94)00091-P. [DOI] [PubMed] [Google Scholar]
  • 275.Subramaniam S, Mohapatra JK, Das B, Sanyal A, Pattnaik B. 2015. Genetic and antigenic analysis of foot-and-mouth disease virus serotype O responsible for outbreaks in India during 2013. Infect Genet Evol 30:59–64. doi: 10.1016/j.meegid.2014.12.009. [DOI] [PubMed] [Google Scholar]
  • 276.Forss S, Schaller H. 1982. A tandem repeat gene in a picornavirus. Nucleic Acids Res 10:6441–6450. doi: 10.1093/nar/10.20.6441. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 277.Herod MR, Gold S, Lasecka-Dykes L, Wright C, Ward JC, McLean TC, Forrest S, Jackson T, Tuthill TJ, Rowlands DJ, Stonehouse NJ. 2017. Genetic economy in picornaviruses: foot-and-mouth disease virus replication exploits alternative precursor cleavage pathways. PLoS Pathog 13:e1006666. doi: 10.1371/journal.ppat.1006666. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 278.Pilipenko EV, Blinov VM, Agol VI. 1990. Gross rearrangements within the 5′-untranslated region of the picornaviral genomes. Nucleic Acids Res 18:3371–3375. doi: 10.1093/nar/18.11.3371. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 279.Tsuchiaka S, Rahpaya SS, Otomaru K, Aoki H, Kishimoto M, Naoi Y, Omatsu T, Sano K, Okazaki-Terashima S, Katayama Y, Oba M, Nagai M, Mizutani T. 2017. Identification of a novel bovine enterovirus possessing highly divergent amino acid sequences in capsid protein. BMC Microbiol 17:18. doi: 10.1186/s12866-016-0923-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 280.Simon-Loriere E, Holmes EC. 2013. Gene duplication is infrequent in the recent evolutionary history of RNA viruses. Mol Biol Evol 30:1263–1269. doi: 10.1093/molbev/mst044. [DOI] [PubMed] [Google Scholar]
  • 281.Ahmed A, Haider SH, Parveen S, Arshad M, Alsenaidy HA, Baaboud AO, Mobaireek KF, AlSaadi MM, Alsenaidy AM, Sullender W. 2016. Co-circulation of 72 bp duplication group A and 60 bp duplication group B respiratory syncytial virus (RSV) strains in Riyadh, Saudi Arabia during 2014. PLoS One 11:e0166145. doi: 10.1371/journal.pone.0166145. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 282.Villordo SM, Carballeda JM, Filomatori CV, Gamarnik AV. 2016. RNA structure duplications and flavivirus host adaptation. Trends Microbiol 24:270–283. doi: 10.1016/j.tim.2016.01.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 283.Le Guillou-Guillemette H, Pivert A, Bouthry E, Henquell C, Petsaris O, Ducancelle A, Veillon P, Vallet S, Alain S, Thibault V, Abravanel F, Rosenbere AA, Andre-Garnier E, Bour JB, Baazia Y, Trimoulet P, Andre P, Gaudy-Graffin C, Bettinger D, Larrat S, Signori-Schmuck A, Saoudin H, Pozzetto B, Lagathu G, Minjolle-Cha S, Stoll-Keller F, Pawlotsky JM, Izopet J, Payan C, Lunel-Fabiani F, Lemaire C. 2017. Natural non-homologous recombination led to the emergence of a duplicated V3-NS5A region in HCV-1b strains associated with hepatocellular carcinoma. PLoS One 12:e0174651. doi: 10.1371/journal.pone.0174651. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 284.Gulyaeva A, Dunowska M, Hoogendoorn E, Giles J, Samborskiy D, Gorbalenya AE. 2017. Domain organization and evolution of the highly divergent 5′ coding region of genomes of arteriviruses, including the novel possum nidovirus. J Virol 91:e02096-16. doi: 10.1128/JVI.02096-16. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 285.Lu HH, Wimmer E. 1996. Poliovirus chimeras replicating under the translational control of genetic elements of hepatitis C virus reveal unusual properties of the internal ribosomal entry site of hepatitis C virus. Proc Natl Acad Sci U S A 93:1412–1417. doi: 10.1073/pnas.93.4.1412. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 286.Pilipenko EV, Viktorova EG, Guest ST, Agol VI, Roos RP. 2001. Cell-specific proteins regulate viral RNA translation and virus-induced disease. EMBO J 20:6899–6908. doi: 10.1093/emboj/20.23.6899. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 287.Kloc A, Diaz-San Segundo F, Schafer EA, Rai DK, Kenney M, de los Santos T, Rieder E. 2017. Foot-and-mouth disease virus 5′-terminal S fragment is required for replication and modulation of the innate immune response in host cells. Virology 512:132–143. doi: 10.1016/j.virol.2017.08.036. [DOI] [PubMed] [Google Scholar]
  • 288.Elena SF, Sole RV, Sardanyes J. 2010. Simple genomes, complex interactions: epistasis in RNA virus. Chaos 20:026106. doi: 10.1063/1.3449300. [DOI] [PubMed] [Google Scholar]
  • 289.Moratorio G, Henningsson R, Barbezange C, Carrau L, Borderia AV, Blanc H, Beaucourt S, Poirier EZ, Vallet T, Boussier J, Mounce BC, Fontes M, Vignuzzi M. 2017. Attenuation of RNA viruses by redirecting their evolution in sequence space. Nat Microbiol 2:17088. doi: 10.1038/nmicrobiol.2017.88. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 290.Escarmis C, Lazaro E, Arias A, Domingo E. 2008. Repeated bottleneck transfers can lead to non-cytocidal forms of a cytopathic virus: implications for viral extinction. J Mol Biol 376:367–379. doi: 10.1016/j.jmb.2007.11.042. [DOI] [PubMed] [Google Scholar]
  • 291.Agol VI. 2006. Molecular mechanisms of poliovirus variation and evolution. Curr Top Microbiol Immunol 299:211–259. [DOI] [PubMed] [Google Scholar]
  • 292.Barr J, Fearns R. 2010. How RNA viruses maintain their genome integrity. J Gen Virol 91:1373–1387. doi: 10.1099/vir.0.020818-0. [DOI] [PubMed] [Google Scholar]
  • 293.Cords CE, Holland JJ. 1964. Replication of poliovirus RNA induced by heterologous virus. Proc Natl Acad Sci U S A 51:1080–1082. doi: 10.1073/pnas.51.6.1080. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 294.Holland JJ, Cords CE. 1964. Maturation of poliovirus RNA with capsid protein coded by heterologous enteroviruses. Proc Natl Acad Sci U S A 51:1082–1085. doi: 10.1073/pnas.51.6.1082. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 295.Agol VI, Shirman GA. 1964. Interaction of guanidine-sensitive and guanidine-dependent variants of poliovirus in mixedly infected cells. Biochem Biophys Res Commun 17:28–33. doi: 10.1016/0006-291X(64)90295-5. [DOI] [Google Scholar]
  • 296.Agol VI, Shirman GA. 1966. Formation of virus particles via enzyme systems and structural proteins induced by another, “assistant” virus. Fed Proc Transl Suppl 25:315–317. [PubMed] [Google Scholar]
  • 297.Wecker E, Lederhilger G. 1964. Genomic masking produced by double-infection of HeLa cells with heterotypic polioviruses. Proc Natl Acad Sci U S A 52:705–709. doi: 10.1073/pnas.52.3.705. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 298.Wilke CO, Novella IS. 2003. Phenotypic mixing and hiding may contribute to memory in viral quasispecies. BMC Microbiol 3:11. doi: 10.1186/1471-2180-3-11. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 299.Sardanyes J, Elena SF. 2010. Error threshold in RNA quasispecies models with complementation. J Theor Biol 265:278–286. doi: 10.1016/j.jtbi.2010.05.018. [DOI] [PubMed] [Google Scholar]
  • 300.Shirogane Y, Watanabe S, Yanagi Y. 2013. Cooperation: another mechanism of viral evolution. Trends Microbiol 21:320–324. doi: 10.1016/j.tim.2013.05.004. [DOI] [PubMed] [Google Scholar]
  • 301.Domingo E, Schuster P (ed). 2016. Current topics in microbiology and immunology, vol 392 Quasispecies: from theory to experimental systems. Springer, Cham, Switzerland. doi: 10.1007/978-3-319-23898-2. [DOI] [Google Scholar]
  • 302.Klump WM, Bergmann I, Muller BC, Ameis D, Kandolf R. 1990. Complete nucleotide-sequence of infectious coxsackievirus B3 cDNA: two initial 5′ uridine residues are regained during plus-strand RNA synthesis. J Virol 64:1573–1583. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 303.Harmon SA, Richards OC, Summers DF, Ehrenfeld E. 1991. The 5′-terminal nucleotides of hepatitis A virus RNA, but not poliovirus RNA, are required for infectivity. J Virol 65:2757–2760. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 304.Miner JC, Chen AA, Garcia AE. 2016. Free-energy landscape of a hyperstable RNA tetraloop. Proc Natl Acad Sci U S A 113:6665–6670. doi: 10.1073/pnas.1603154113. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 305.Bottaro S, Gil-Ley A, Bussi G. 2016. RNA folding pathways in stop motion. Nucleic Acids Res 44:5883–5891. doi: 10.1093/nar/gkw239. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 306.D'Ascenzo L, Leonarski F, Vicens Q, Auffinger P. 2017. Revisiting GNRA and UNCG folds: U-turns versus Z-turns in RNA hairpin loops. RNA 23:259–269. doi: 10.1261/rna.059097.116. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 307.Cann AJ, Stanway G, Hughes PJ, Minor PD, Evans DMA, Schild GC, Almond JW. 1984. Reversion to neurovirulence of the live-attenuated Sabin type 3 oral poliovirus vaccine. Nucleic Acids Res 12:7787–7792. doi: 10.1093/nar/12.20.7787. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 308.Evans DMA, Dunn G, Minor PD, Schild GC, Cann AJ, Stanway G, Almond JW, Currey K, Maizel JV. 1985. Increased neurovirulence associated with a single nucleotide change in a noncoding region of the Sabin type 3 poliovaccine genome. Nature 314:548–550. doi: 10.1038/314548a0. [DOI] [PubMed] [Google Scholar]
  • 309.Minor PD, Dunn G. 1988. The effect of sequences in the 5′ non-coding region on the replication of polioviruses in the human gut. J Gen Virol 69:1091–1096. doi: 10.1099/0022-1317-69-5-1091. [DOI] [PubMed] [Google Scholar]
  • 310.Macadam AJ, Arnold C, Howlett J, John A, Marsden S, Taffs F, Reeve P, Hamada N, Wareham K, Almond J, Cammack N, Minor PD. 1989. Reversion of the attenuated and temperature-sensitive phenotypes of the Sabin type 3 strain of poliovirus in vaccinees. Virology 172:408–414. doi: 10.1016/0042-6822(89)90183-9. [DOI] [PubMed] [Google Scholar]
  • 311.Muzychenko AR, Lipskaya GY, Maslova SV, Svitkin YV, Pilipenko EV, Nottay BK, Kew OM, Agol VI. 1991. Coupled mutations in the 5′-untranslated region of the Sabin poliovirus strains during in vivo passages: structural and functional implications. Virus Res 21:111–122. doi: 10.1016/0168-1702(91)90002-D. [DOI] [PubMed] [Google Scholar]
  • 312.Kew OM, Sutter RW, de Gourville EM, Dowdle WR, Pallansch MA. 2005. Vaccine-derived polioviruses and the endgame strategy for global polio eradication. Annu Rev Microbiol 59:587–635. doi: 10.1146/annurev.micro.58.030603.123625. [DOI] [PubMed] [Google Scholar]
  • 313.Agol VI. 2006. Vaccine-derived polioviruses. Biologicals 34:103–108. doi: 10.1016/j.biologicals.2006.02.007. [DOI] [PubMed] [Google Scholar]
  • 314.Savolainen-Kopra C, Blomqvist S. 2010. Mechanisms of genetic variation in polioviruses. Rev Med Virol 20:358–371. doi: 10.1002/rmv.663. [DOI] [PubMed] [Google Scholar]
  • 315.Skinner MA, Racaniello VR, Dunn G, Cooper J, Minor PD, Almond JW. 1989. New model for the secondary structure of the 5′ non-coding RNA of poliovirus is supported by biochemical and genetic data that also show that RNA secondary structure is important in neurovirulence. J Mol Biol 207:379–392. doi: 10.1016/0022-2836(89)90261-1. [DOI] [PubMed] [Google Scholar]
  • 316.Svitkin YV, Maslova SV, Agol VI. 1985. The genomes of attenuated and virulent poliovirus strains differ in their in vitro translation efficiencies. Virology 147:243–252. doi: 10.1016/0042-6822(85)90127-8. [DOI] [PubMed] [Google Scholar]
  • 317.Svitkin YV, Pestova TV, Maslova SV, Agol VI. 1988. Point mutations modify the response of poliovirus RNA to a translation initiation factor: a comparison of neurovirulent and attenuated strains. Virology 166:394–404. doi: 10.1016/0042-6822(88)90510-7. [DOI] [PubMed] [Google Scholar]
  • 318.Svitkin YV, Cammack N, Minor PD, Almond JW. 1990. Translation deficiency of the Sabin type 3 poliovirus genome: association with an attenuating mutation C472-U. Virology 175:103–109. doi: 10.1016/0042-6822(90)90190-3. [DOI] [PubMed] [Google Scholar]
  • 319.Svitkin YV, Alpatova GA, Lipskaya GA, Maslova SV, Agol VI, Kew O, Meerovitch K, Sonenberg N. 1993. Towards development of an in vitro translation test for poliovirus neurovirulence. Dev Biol Stand 78:27–32. [PubMed] [Google Scholar]
  • 320.Kawamura N, Kohara M, Abe S, Komatsu T, Tago K, Arita M, Nomoto A. 1989. Determinants in the 5′ noncoding region of poliovirus Sabin 1 RNA that influence the attenuation phenotype. J Virol 63:1302–1309. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 321.Westrop GD, Wareham KA, Evans DM, Dunn G, Minor PD, Magrath DI, Taffs F, Marsden S, Skinner MA, Schild GC, Almond JW. 1989. Genetic basis of attenuation of the Sabin type 3 oral poliovirus vaccine. J Virol 63:1338–1344. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 322.Famulare M, Chang S, Iber J, Zhao K, Adeniji JA, Bukbuk D, Baba M, Behrend M, Burns CC, Oberste MS. 2016. Sabin vaccine reversion in the field: a comprehensive analysis of Sabin-like poliovirus isolates in Nigeria. J Virol 90:317–331. doi: 10.1128/JVI.01532-15. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 323.Christodoulou C, Colbere-Garapin F, Macadam A, Taffs LF, Marsden S, Minor P, Horaud F. 1990. Mapping of mutations associated with neurovirulence in monkeys infected with Sabin 1 poliovirus revertants selected at high temperature. J Virol 64:4922–4929. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 324.Rowe A, Ferguson GL, Minor PD, Macadam AJ. 2000. Coding changes in the poliovirus protease 2A compensate for 5′ NCR domain V disruptions in a cell-specific manner. Virology 269:284–293. doi: 10.1006/viro.2000.0244. [DOI] [PubMed] [Google Scholar]
  • 325.Jang SK, Pestova TV, Hellen CUT, Witherell GW, Wimmer E. 1990. Cap-independent translation of picornavirus RNAs: structure and function of the internal ribosomal entry site. Enzyme 44:292–309. doi: 10.1159/000468766. [DOI] [PubMed] [Google Scholar]
  • 326.Pestova TV, Hellen CUT, Wimmer E. 1991. Translation of poliovirus RNA: role of an essential cis-acting oligopyrimidine element within the 5′ nontranslated region and involvement of a cellular 57-kilodalton protein. J Virol 65:6194–6204. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 327.Nicholson R, Pelletier J, Le SY, Sonenberg N. 1991. Structural and functional analysis of the ribosome landing pad of poliovirus type 2: in vivo translation studies. J Virol 65:5886–5894. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 328.Iizuka N, Kohara M, Haginoyamagishi K, Abe S, Komatsu T, Tago K, Arita M, Nomoto A. 1989. Construction of less neurovirulent polioviruses by introducing deletions into the 5′ noncoding sequence of the genome. J Virol 63:5354–5363. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 329.Stewart SR, Semler BL. 1998. RNA structure adjacent to the attenuation determinant in the 5′-non-coding region influences poliovirus viability. Nucleic Acids Res 26:5318–5326. doi: 10.1093/nar/26.23.5318. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 330.La Monica N, Racaniello VR. 1989. Differences in replication of attenuated and neurovirulent polioviruses in human neuroblastoma cell line SH-SY5Y. J Virol 63:2357–2360. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 331.Agol VI, Drozdov SG, Ivannikova TA, Kolesnikova MS, Korolev MB, Tolskaya EA. 1989. Restricted growth of attenuated poliovirus strains in cultured cells of a human neuroblastoma. J Virol 63:4034–4038. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 332.Haller AA, Stewart SR, Semler BL. 1996. Attenuation stem-loop lesions in the 5′ noncoding region of poliovirus RNA: neuronal cell-specific translation defects. J Virol 70:1467–1474. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 333.Hoffman MA, Palmenberg AC. 1995. Mutational analysis of the J-K stem-loop region of the encephalomyocarditis virus IRES. J Virol 69:4399–4406. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 334.Hoffman MA, Palmenberg AC. 1996. Revertant analysis of J-K mutations in the encephalomyocarditis virus internal ribosomal entry site detects an altered leader protein. J Virol 70:6425–6430. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 335.Pilipenko EV, Gmyl AP, Maslova SV, Khitrina EV, Agol VI. 1995. Attenuation of Theiler's murine encephalomyelitis virus by modifications of the oligopyrimidine/AUG tandem, a host-dependent translational cis-element. J Virol 69:864–870. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 336.Dildine SL, Semler BL. 1989. The deletion of 41 proximal nucleotides reverts a poliovirus mutant containing a temperature-sensitive lesion in the 5′ noncoding region of genomic RNA. J Virol 63:847–862. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 337.Sarnow P, Bernstein HD, Baltimore D. 1986. A poliovirus temperature-sensitive RNA synthesis mutant located in a noncoding region of the genome. Proc Natl Acad Sci U S A 83:571–575. doi: 10.1073/pnas.83.3.571. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 338.Jacobson SJ, Konings DAM, Sarnow P. 1993. Biochemical and genetic evidence for a pseudoknot structure at the 3′ terminus of the poliovirus RNA genome and its role in viral RNA amplification. J Virol 67:2961–2971. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 339.Neufeld KL, Galarza JM, Richards OC, Summers DF, Ehrenfeld E. 1994. Identification of terminal adenylyl transferase activity of the poliovirus polymerase 3Dpol. J Virol 68:5811–5818. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 340.Kusov YY, Gosert R, Gauss-Muller V. 2005. Replication and in vivo repair of the hepatitis A virus genome lacking the poly(A) tail. J Gen Virol 86:1363–1368. doi: 10.1099/vir.0.80644-0. [DOI] [PubMed] [Google Scholar]
  • 341.Rieder E, Paul AV, Kim DW, van Boom JH, Wimmer E. 2000. Genetic and biochemical studies of poliovirus cis-acting replication element cre in relation to VPg uridylylation. J Virol 74:10371–10380. doi: 10.1128/JVI.74.22.10371-10380.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 342.Mason PW, Bezborodova SV, Henry TM. 2002. Identification and characterization of a cis-acting replication element (cre) adjacent to the internal ribosome entry site of foot-and-mouth disease virus. J Virol 76:9686–9694. doi: 10.1128/JVI.76.19.9686-9694.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 343.Smithee S, Tracy S, Chapman NM. 2016. Reversion to wildtype of a mutated and nonfunctional coxsackievirus B3 CRE(2C). Virus Res 220:136–149. doi: 10.1016/j.virusres.2016.04.016. [DOI] [PubMed] [Google Scholar]
  • 344.Omata T, Kohara M, Kuge S, Komatsu T, Abe S, Semler BL, Kameda A, Itoh H, Arita M, Wimmer E, Nomoto A. 1986. Genetic analysis of the attenuation phenotype of poliovirus type 1. J Virol 58:348–358. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 345.Ren R, Moss EG, Racaniello VR. 1991. Identification of two determinants that attenuate vaccine-related type 2 poliovirus. J Virol 65:1377–1382. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 346.Macadam AJ, Pollard SR, Ferguson G, Skuce R, Wood D, Almond JW, Minor PD. 1993. Genetic basis of attenuation of the Sabin type 2 vaccine strain of poliovirus in primates. Virology 192:18–26. doi: 10.1006/viro.1993.1003. [DOI] [PubMed] [Google Scholar]
  • 347.Bouchard MJ, Lam DH, Racaniello VR. 1995. Determinants of attenuation and temperature sensitivity in the type 1 poliovirus Sabin vaccine. J Virol 69:4972–4978. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 348.Kew OM, Nottay BK. 1984. Evolution of the oral poliovaccine strains in humans occurs by both mutation and intramolecular recombination, p 357–362. In Chanock R, Lerner R (ed), Modern approaches to vaccines. Cold Spring Harbor Laboratory Press, Cold Spring Harbor, NY. [Google Scholar]
  • 349.Cammack N, Phillips A, Dunn G, Patel V, Minor PD. 1988. Intertypic genomic rearrangements of poliovirus strains in vaccinees. Virology 167:507–514. doi: 10.1016/S0042-6822(88)90113-4. [DOI] [PubMed] [Google Scholar]
  • 350.Lipskaya GY, Muzychenko AR, Kutitova OK, Maslova SV, Equestre M, Drozdov SG, Perez Bercoff R, Agol VI. 1991. Frequent isolation of intertypic poliovirus recombinants with serotype 2 specificity from vaccine-associated polio cases. J Med Virol 35:290–296. doi: 10.1002/jmv.1890350415. [DOI] [PubMed] [Google Scholar]
  • 351.Furione M, Guillot S, Otelea D, Balanant J, Candrea A, Crainic R. 1993. Polioviruses with natural recombinant genomes isolated from vaccine-associated paralytic poliomyelitis. Virology 196:199–208. doi: 10.1006/viro.1993.1468. [DOI] [PubMed] [Google Scholar]
  • 352.Driesel G, Diedrich S, Kunkel U, Schreier E. 1995. Vaccine-associated cases of poliomyelitis over a 30 year period in East Germany. Eur J Epidemiol 11:647–654. doi: 10.1007/BF01720298. [DOI] [PubMed] [Google Scholar]
  • 353.Korotkova E, Laassri M, Zagorodnyaya T, Petrovskaya S, Rodionova E, Cherkasova E, Gmyl A, Ivanova OE, Eremeeva TP, Lipskaya GY, Agol VI, Chumakov K. 2017. Pressure for pattern-specific intertypic recombination between Sabin polioviruses: evolutionary implications. Viruses 9:353. doi: 10.3390/v9110353. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 354.Rico-Hesse R, Pallansch MA, Nottay BK, Kew OM. 1987. Geographic distribution of wild poliovirus type 1 genotypes. Virology 160:311–322. doi: 10.1016/0042-6822(87)90001-8. [DOI] [PubMed] [Google Scholar]
  • 355.Guillot S, Caro V, Cuervo N, Korotkova E, Combiescu M, Persu A, Aubert-Combiescu A, Delpeyroux F, Crainic R. 2000. Natural genetic exchanges between vaccine and wild poliovirus strains in humans. J Virol 74:8434–8443. doi: 10.1128/JVI.74.18.8434-8443.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 356.Liu HM, Zheng DP, Zhang LB, Oberste MS, Pallansch MA, Kew OM. 2000. Molecular evolution of a type 1 wild-vaccine poliovirus recombinant during widespread circulation in China. J Virol 74:11153–11161. doi: 10.1128/JVI.74.23.11153-11161.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 357.Combelas N, Holmblat B, Joffret ML, Colbere-Garapin F, Delpeyroux F. 2011. Recombination between poliovirus and coxsackie A viruses of species C: a model of viral genetic plasticity and emergence. Viruses 3:1460–1484. doi: 10.3390/v3081460. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 358.Arita M, Zhu SL, Yoshida H, Yoneyama T, Miyamura T, Shimizu H. 2005. A Sabin 3-derived poliovirus recombinant contained a sequence homologous with indigenous human enterovirus species C in the viral polymerase coding region. J Virol 79:12650–12657. doi: 10.1128/JVI.79.20.12650-12657.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 359.de la Torre JC, Giachetti C, Semler BL, Holland JJ. 1992. High-frequency of single base transitions and extreme frequency of precise multiple base reversion mutations in poliovirus. Proc Natl Acad Sci U S A 89:2531–2535. doi: 10.1073/pnas.89.7.2531. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 360.Wessels E, Notebaart RA, Duijsings D, Lanke K, Vergeer B, Melchers WJG, van Kuppeveld FJM. 2006. Structure-function analysis of the coxsackievirus protein 3A: identification of residues important for dimerization, viral RNA replication, and transport inhibition. J Biol Chem 281:28232–28243. doi: 10.1074/jbc.M601122200. [DOI] [PubMed] [Google Scholar]
  • 361.Wang CL, Ma HC, Wimmer E, Jiang P, Paul AV. 2014. A C-terminal, cysteine-rich site in poliovirus 2CATPase is required for morphogenesis. J Gen Virol 95:1255–1265. doi: 10.1099/vir.0.062497-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 362.Asare E, Mugavero J, Jiang P, Wimmer E, Paul AV. 2016. A single amino acid substitution in poliovirus nonstructural protein 2CATPase causes conditional defects in encapsidation and uncoating. J Virol 90:6174–6186. doi: 10.1128/JVI.02877-15. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 363.Charini WA, Todd S, Gutman GA, Semler BL. 1994. Transduction of a human RNA sequence by poliovirus. J Virol 68:6547–6552. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 364.Lowry K, Woodman A, Cook J, Evans DJ. 2014. Recombination in enteroviruses is a biphasic replicative process involving the generation of greater than genome length ‘imprecise’ intermediates. PLoS Pathog 10:e1004191. doi: 10.1371/journal.ppat.1004191. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 365.Muslin C, Joffret ML, Pelletier I, Blondel B, Delpeyroux F. 2015. Evolution and emergence of enteroviruses through intra- and inter-species recombination: plasticity and phenotypic impact of modular genetic exchanges in the 5′ untranslated region. PLoS Pathog 11:e1005266. doi: 10.1371/journal.ppat.1005266. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 366.Li JP, Baltimore D. 1988. Isolation of poliovirus 2C mutants defective in viral RNA synthesis. J Virol 62:4016–4021. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 367.Li JP, Baltimore D. 1990. An intragenic revertant of a poliovirus 2C mutant has an uncoating defect. J Virol 64:1102–1107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 368.Cao XM, Kuhn RJ, Wimmer E. 1993. Replication of poliovirus RNA containing two VPg coding sequences leads to a specific deletion event. J Virol 67:5572–5578. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 369.Arias A, Perales C, Escarmis C, Domingo E. 2010. Deletion mutants of VPg reveal new cytopathology determinants in a picornavirus. PLoS One 5:e10735. doi: 10.1371/journal.pone.0010735. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 370.Liu Y, Wang CL, Mueller S, Paul AV, Wimmer E, Jiang P. 2010. Direct interaction between two viral proteins, the nonstructural protein 2CATPase and the capsid protein VP3, is required for enterovirus morphogenesis. PLoS Pathog 6:e1001066. doi: 10.1371/journal.ppat.1001066. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 371.Song YT, Paul AV, Wimmer E. 2012. Evolution of poliovirus defective interfering particles expressing Gaussia luciferase. J Virol 86:1999–2010. doi: 10.1128/JVI.05871-11. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 372.Huang AS. 1973. Defective interfering viruses. Annu Rev Microbiol 27:101–117. doi: 10.1146/annurev.mi.27.100173.000533. [DOI] [PubMed] [Google Scholar]
  • 373.Cole CN, Smoler D, Wimmer E, Baltimore D. 1971. Defective interfering particles of poliovirus. I. Isolation and physical properties. J Virol 7:478–485. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 374.Lazzarini RA, Keene JD, Schubert M. 1981. The origins of defective interfering particles of the negative-strand RNA viruses. Cell 26:145–154. doi: 10.1016/0092-8674(81)90298-1. [DOI] [PubMed] [Google Scholar]
  • 375.Pathak KB, Nagy PD. 2009. Defective interfering RNAs: foes of viruses and friends of virologists. Viruses 1:895–919. doi: 10.3390/v1030895. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 376.Li D, Lott WB, Lowry K, Jones A, Thu HM, Aaskov J. 2011. Defective interfering viral particles in acute dengue infections. PLoS One 6:e19447. doi: 10.1371/journal.pone.0019447. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 377.Pesko KN, Fitzpatrick KA, Ryan EM, Shi P-Y, Zhang B, Lennon NJ, Newman RM, Henn MR, Ebel GD. 2012. Internally deleted WNV genomes isolated from exotic birds in New Mexico: function in cells, mosquitoes, and mice. Virology 427:10–17. doi: 10.1016/j.virol.2012.01.028. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 378.Saira K, Lin X, DePasse JV, Halpin R, Twaddle A, Stockwell T, Angus B, Cozzi-Lepri A, Delfino M, Dugan V, Dwyer DE, Freiberg M, Horban A, Losso M, Lynfield R, Wentworth DN, Holmes EC, Davey R, Wentworth DE, Ghedin E, INSIGHT FLU002 Study Group, INSIGHT FLU003 Study Group. 2013. Sequence analysis of in vivo defective interfering-like RNA of influenza A H1N1 pandemic virus. J Virol 87:8064–8074. doi: 10.1128/JVI.00240-13. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 379.Killip MJ, Young DF, Gatherer D, Ross CS, Short JAL, Davison AJ, Goodbourn S, Randall RE. 2013. Deep sequencing analysis of defective genomes of parainfluenza virus 5 and their role in interferon induction. J Virol 87:4798–4807. doi: 10.1128/JVI.03383-12. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 380.Timm C, Akpinar F, Yin J. 2014. Quantitative characterization of defective virus emergence by deep sequencing. J Virol 88:2623–2632. doi: 10.1128/JVI.02675-13. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 381.Poirier EZ, Mounce BC, Rozen-Gagnon K, Hooikaas PJ, Stapleford KA, Moratorio G, Vignuzzi M. 2016. Low fidelity polymerases of alphaviruses recombine at higher rates to overproduce defective interfering particles. J Virol 90:2446–2454. doi: 10.1128/JVI.02921-15. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 382.Williams ESCP, Morales NM, Wasik BR, Brusic V, Whelan SPJ, Turner PE. 2016. Repeatable population dynamics among vesicular stomatitis virus lineages evolved under high co-infection. Front Microbiol 7:370. doi: 10.3389/fmicb.2016.00370. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 383.Ziegler CM, Eisenhauer P, Bruce EA, Weir ME, King BR, Klaus JP, Krementsov DN, Shirley DJ, Ballif BA, Botten J. 2016. The lymphocytic choriomeningitis virus matrix protein PPXY late domain drives the production of defective interfering particles. PLoS Pathog 12:e1005501. doi: 10.1371/journal.ppat.1005501. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 384.Marriott AC, Dimmock NJ. 2010. Defective interfering viruses and their potential as antiviral agents. Rev Med Virol 20:51–62. doi: 10.1002/rmv.641. [DOI] [PubMed] [Google Scholar]
  • 385.Jaworski E, Routh A. 2017. Parallel ClickSeq and Nanopore sequencing elucidates the rapid evolution of defective-interfering RNAs in flock house virus. PLoS Pathog 13:e1006365. doi: 10.1371/journal.ppat.1006365. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 386.Cole CN, Baltimore D. 1973. Defective interfering particles of poliovirus. II. Nature of the defect. J Mol Biol 76:325–343. [DOI] [PubMed] [Google Scholar]
  • 387.Nomoto A, Jacobson A, Lee YF, Dunn J, Wimmer E. 1979. Defective interfering particles of poliovirus: mapping of the deletion and evidence that the deletions in the genomes of DI(1), DI(2) and DI(3) are located in the same region. J Mol Biol 128:179–196. doi: 10.1016/0022-2836(79)90125-6. [DOI] [PubMed] [Google Scholar]
  • 388.Garcia-Arriaza J, Manrubia SC, Toja M, Domingo E, Escarmis C. 2004. Evolutionary transition toward defective RNAs that are infectious by complementation. J Virol 78:11678–11685. doi: 10.1128/JVI.78.21.11678-11685.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 389.Ojosnegros S, Garcia-Arriaza J, Escarmis C, Manrubia SC, Perales C, Arias A, Mateu MG, Domingo E. 2011. Viral genome segmentation can result from a trade-off between genetic content and particle stability. PLoS Genet 7:e1001344. doi: 10.1371/journal.pgen.1001344. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 390.Moreno E, Ojosnegros S, Garcia-Arriaz J, Escarmis C, Domingo E, Perales C. 2014. Exploration of sequence space as the basis of viral RNA genome segmentation. Proc Natl Acad Sci U S A 111:6678–6683. doi: 10.1073/pnas.1323136111. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 391.Thompson J, Dasgupta I, Fuchs M, Iwanami T, Karasev AV, Petrzik K, Sanfaçon H, Tzanetakis I, Vlugt R, Wetzel T, Yoshikawa N. 2017. ICTV virus taxonomy profile: Secoviridae. J Gen Virol 98:529–531. doi: 10.1099/jgv.0.000779. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 392.Yuan T, Wang H, Li C, Yang D, Zhou G, Yu L. 2017. T135I substitution in the nonstructural protein 2C enhances foot-and-mouth disease virus replication. Virus Genes 53:840–847. doi: 10.1007/s11262-017-1480-9. [DOI] [PubMed] [Google Scholar]
  • 393.Gromeier M, Alexander L, Wimmer E. 1996. Internal ribosomal entry site substitution eliminates neurovirulence in intergeneric poliovirus recombinants. Proc Natl Acad Sci U S A 93:2370–2375. doi: 10.1073/pnas.93.6.2370. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 394.Jahan N, Wimmer E, Mueller S. 2011. A host-specific, temperature-sensitive translation defect determines the attenuation phenotype of a human rhinovirus/poliovirus chimera, PV1(RIPO). J Virol 85:7225–7235. doi: 10.1128/JVI.01804-09. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 395.Gromeier M, Lachmann S, Rosenfeld MR, Gutin PH, Wimmer E. 2000. Intergeneric poliovirus recombinants for the treatment of malignant glioma. Proc Natl Acad Sci U S A 97:6803–6808. doi: 10.1073/pnas.97.12.6803. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 396.Brown MC, Dobrikova EY, Dobrikov MI, Walton RW, Gemberling SL, Nair SK, Desjardins A, Sampson JH, Friedman HS, Friedman AH, Tyler DS, Bigner DD, Gromeier M. 2014. Oncolytic polio virotherapy of cancer. Cancer 120:3277–3286. doi: 10.1002/cncr.28862. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 397.Florez de Sessions P, Dobrikova E, Gromeier M. 2007. Genetic adaptation to untranslated region-mediated enterovirus growth deficits by mutations in the nonstructural proteins 3AB and 3CD. J Virol 81:8396–8405. doi: 10.1128/JVI.00321-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 398.Bradrick SS, Lieben EA, Carden BM, Romero JR. 2001. A predicted secondary structural domain within the internal ribosome entry site of echovirus 12 mediates a cell-type-specific block to viral replication. J Virol 75:6472–6481. doi: 10.1128/JVI.75.14.6472-6481.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 399.Zhao WD, Lahser FC, Wimmer E. 2000. Genetic analysis of a poliovirus/hepatitis C virus (HCV) chimera: interaction between the poliovirus cloverleaf and a sequence in the HCV 5′ nontranslated region results in a replication phenotype. J Virol 74:6223–6226. doi: 10.1128/JVI.74.13.6223-6226.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 400.Johnson VH, Semler BL. 1988. Defined recombinants of poliovirus and coxsackievirus: sequence-specific deletions and functional substitutions in the 5′-noncoding regions of viral RNAs. Virology 162:47–57. doi: 10.1016/0042-6822(88)90393-5. [DOI] [PubMed] [Google Scholar]
  • 401.Zell R, Klingel K, Sauter M, Fortmuller U, Kandolf R. 1995. Coxsackieviral proteins functionally recognize the polioviral cloverleaf structure of the 5′-NTR of a chimeric enterovirus RNA: influence of species-specific host cell factors on virus growth. Virus Res 39:87–103. doi: 10.1016/0168-1702(95)00075-5. [DOI] [PubMed] [Google Scholar]
  • 402.Jiang P, Liu Y, Ma HC, Paul AV, Wimmer E. 2014. Picornavirus morphogenesis. Microbiol Mol Biol Rev 78:418–437. doi: 10.1128/MMBR.00012-14. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 403.Monroe SS, Schlesinger S. 1983. RNAs from 2 independently isolated defective interfering particles of Sindbis virus contain a cellular tRNA sequence at their 5′ ends. Proc Natl Acad Sci U S A 80:3279–3283. doi: 10.1073/pnas.80.11.3279. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 404.Khatchikian D, Orlich M, Rott R. 1989. Increased viral pathogenicity after insertion of a 28S ribosomal RNA sequence into the hemagglutinin gene of an influenza virus. Nature 340:156–157. doi: 10.1038/340156a0. [DOI] [PubMed] [Google Scholar]
  • 405.Becher P, Tautz N. 2011. RNA recombination in pestiviruses: cellular RNA sequences in viral genomes highlight the role of host factors for viral persistence and lethal disease. RNA Biol 8:216–224. doi: 10.4161/rna.8.2.14514. [DOI] [PubMed] [Google Scholar]
  • 406.Beaucourt S, Vignuzzi M. 2014. Ribavirin: a drug active against many viruses with multiple effects on virus replication and propagation. Molecular basis of ribavirin resistance. Curr Opin Virol 8:10–15. doi: 10.1016/j.coviro.2014.04.011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 407.Crotty S, Maag D, Arnold JJ, Zhong WD, Lau JYN, Hong Z, Andino R, Cameron CE. 2000. The broad-spectrum antiviral ribonucleoside ribavirin is an RNA virus mutagen. Nat Med 6:1375–1379. doi: 10.1038/82191. [DOI] [PubMed] [Google Scholar]
  • 408.Crotty S, Cameron CE, Andino R. 2001. RNA virus error catastrophe: direct molecular test by using ribavirin. Proc Natl Acad Sci U S A 98:6895–6900. doi: 10.1073/pnas.111085598. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 409.Domingo E, Schuster P. 2016. What is a quasispecies? Historical origins and current scope. Curr Top Microbiol Immunol 392:1–22. doi: 10.1007/82_2015_453. [DOI] [PubMed] [Google Scholar]
  • 410.Levi LI, Gnadig NF, Beaucourt S, McPherson MJ, Baron B, Arnold JJ, Vignuzzi M. 2010. Fidelity variants of RNA dependent RNA polymerases uncover an indirect, mutagenic activity of amiloride compounds. PLoS Pathog 6:e1001163. doi: 10.1371/journal.ppat.1001163. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 411.Agudo R, Ferrer-Orta C, Arias A, de la Higuera I, Perales C, Perez-Luque R, Verdaguer N, Domingo E. 2010. A multi-step process of viral adaptation to a mutagenic nucleoside analogue by modulation of transition types leads to extinction-escape. PLoS Pathog 6:e1001072. doi: 10.1371/journal.ppat.1001072. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 412.Mihalik KB, Feigelstock DA. 2013. Sensitivity of a ribavirin resistant mutant of hepatitis C virus to other antiviral drugs. PLoS One 8:e74027. doi: 10.1371/journal.pone.0074027. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 413.de la Higuera I, Ferrer-Orta C, de Avila AI, Perales C, Sierra M, Singh K, Sarafianos SG, Dehouck Y, Bastolla U, Verdaguer N, Domingo E. 2017. Molecular and functional bases of selection against a mutation bias in an RNA virus. Genome Biol Evol 9:1212–1228. doi: 10.1093/gbe/evx075. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 414.Pincus SE, Diamond DC, Emini EA, Wimmer E. 1986. Guanidine-selected mutants of poliovirus: mapping of point mutations to polypeptide 2C. J Virol 57:638–646. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 415.Pincus SE, Wimmer E. 1986. Production of guanidine-resistant and guanidine-dependent poliovirus mutants from cloned cDNA: mutations in polypeptide 2C are directly responsible for altered guanidine sensitivity. J Virol 60:793–796. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 416.Pincus SE, Rohl H, Wimmer E. 1987. Guanidine-dependent mutants of poliovirus: identification of 3 classes with different growth requirements. Virology 157:83–88. doi: 10.1016/0042-6822(87)90316-3. [DOI] [PubMed] [Google Scholar]
  • 417.Baltera RF, Tershak DR. 1989. Guanidine-resistant mutants of poliovirus have distinct mutations in peptide 2C. J Virol 63:4441–4444. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 418.Tolskaya EA, Romanova LI, Kolesnikova MS, Gmyl AP, Gorbalenya AE, Agol VI. 1994. Genetic studies on the poliovirus 2C protein, an NTPase: a plausible mechanism of guanidine effect on the 2C function and evidence for the importance of 2C oligomerization. J Mol Biol 236:1310–1323. doi: 10.1016/0022-2836(94)90060-4. [DOI] [PubMed] [Google Scholar]
  • 419.Rodriguez PL, Carrasco L. 1993. Poliovirus protein 2C has ATPase and GTPase activities. J Biol Chem 268:8105–8110. [PubMed] [Google Scholar]
  • 420.Mirzayan C, Wimmer E. 1994. Biochemical studies on poliovirus polypeptide 2C: evidence for ATPase activity. Virology 199:176–187. doi: 10.1006/viro.1994.1110. [DOI] [PubMed] [Google Scholar]
  • 421.Gorbalenya AE, Koonin EV, Donchenko AP, Blinov VM. 1988. A conserved NTP-motif in putative helicases. Nature 333:22. [DOI] [PubMed] [Google Scholar]
  • 422.Gorbalenya AE, Koonin EV, Wolf YI. 1990. A new superfamily of putative NTP-binding domains encoded by genomes of small DNA and RNA viruses. FEBS Lett 262:145–148. doi: 10.1016/0014-5793(90)80175-I. [DOI] [PubMed] [Google Scholar]
  • 423.Rozovics JM, Semler BL. 2010. Genome replication. I. The players, p 107–125. In Ehrenfeld E, Domingo E, Roos RP (ed), The picornaviruses. ASM Press, Washington, DC. [Google Scholar]
  • 424.Agol VI. 1965. Guanidine-sensitive, guanidine-resistant and guanidine-dependent variants of poliovirus. Some observations and a general hypothesis, p 187–199. In Poliomyelitis and other enteroviral infections. Institute of Poliomyelitis and Viral Encephalitides, Moscow, Russia: (In Russian.) [Google Scholar]
  • 425.Agudo R, de la Higuera I, Arias A, Grande-Perez A, Domingo E. 2016. Involvement of a joker mutation in a polymerase-independent lethal mutagenesis escape mechanism. Virology 494:257–266. doi: 10.1016/j.virol.2016.04.023. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 426.Eltahla AA, Luciani F, White PA, Lloyd AR, Bull RA. 2015. Inhibitors of the hepatitis C virus polymerase; mode of action and resistance. Viruses 7:5206–5224. doi: 10.3390/v7102868. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 427.Escarmis C, Davila M, Charpentier N, Bracho A, Moya A, Domingo E. 1996. Genetic lesions associated with Muller's ratchet in an RNA virus. J Mol Biol 264:255–267. doi: 10.1006/jmbi.1996.0639. [DOI] [PubMed] [Google Scholar]
  • 428.Escarmis C, Davila M, Domingo E. 1999. Multiple molecular pathways for fitness recovery of an RNA virus debilitated by operation of Muller's ratchet. J Mol Biol 285:495–505. doi: 10.1006/jmbi.1998.2366. [DOI] [PubMed] [Google Scholar]
  • 429.Escarmis C, Gomez-Mariano G, Davila M, Lazaro E, Domingo E. 2002. Resistance to extinction of low fitness virus subjected to plaque-to-plaque transfers: diversification by mutation clustering. J Mol Biol 315:647–661. doi: 10.1006/jmbi.2001.5259. [DOI] [PubMed] [Google Scholar]
  • 430.Drake JW, Holland JJ. 1999. Mutation rates among RNA viruses. Proc Natl Acad Sci U S A 96:13910–13913. doi: 10.1073/pnas.96.24.13910. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 431.Sanjuan R, Nebot MR, Chirico N, Mansky LM, Belshaw R. 2010. Viral mutation rates. J Virol 84:9733–9748. doi: 10.1128/JVI.00694-10. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 432.Smith EC, Sexton NR, Denison MR. 2014. Thinking outside the triangle: replication fidelity of the largest RNA viruses. Annu Rev Virol 1:111–132. doi: 10.1146/annurev-virology-031413-085507. [DOI] [PubMed] [Google Scholar]
  • 433.Sexton NR, Smith EC, Blanc H, Vignuzzi M, Peersen OB, Denison MR. 2016. Homology-based identification of a mutation in the coronavirus RNA-dependent RNA polymerase that confers resistance to multiple mutagens. J Virol 90:7415–7428. doi: 10.1128/JVI.00080-16. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 434.Stapleford KA, Rozen-Gagnon K, Das PK, Saul S, Poirier EZ, Blanc H, Vidalain PO, Merits A, Vignuzzi M. 2015. Viral polymerase-helicase complexes regulate replication fidelity to overcome intracellular nucleotide depletion. J Virol 89:11233–11244. doi: 10.1128/JVI.01553-15. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 435.Minskaia E, Hertzig T, Gorbalenya AE, Campanacci V, Cambillau C, Canard B, Ziebuhr J. 2006. Discovery of an RNA virus 3′ → 5′ exoribonuclease that is critically involved in coronavirus RNA synthesis. Proc Natl Acad Sci U S A 103:5108–5113. doi: 10.1073/pnas.0508200103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 436.Eckerle LD, Lu X, Sperry SM, Choi L, Denison MR. 2007. High fidelity of murine hepatitis virus replication is decreased in nsp14 exoribonuclease mutants. J Virol 81:12135–12144. doi: 10.1128/JVI.01296-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 437.Eckerle LD, Becker MM, Halpin RA, Li K, Venter E, Lu XT, Scherbakova S, Graham RL, Baric RS, Stockwell TB, Spiro DJ, Denison MR. 2010. Infidelity of SARS-CoV Nsp14-exonuclease mutant virus replication is revealed by complete genome sequencing. PLoS Pathog 6:e1000896. doi: 10.1371/journal.ppat.1000896. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 438.Smith EC, Blanc H, Vignuzzi M, Denison MR. 2013. Coronaviruses lacking exoribonuclease activity are susceptible to lethal mutagenesis: evidence for proofreading and potential therapeutics. PLoS Pathog 9:e1003565. doi: 10.1371/journal.ppat.1003565. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 439.Smith EC, Case JB, Blanc H, Isakov O, Shomron N, Vignuzzi M, Denison MR. 2015. Mutations in coronavirus nonstructural protein 10 decrease virus replication fidelity. J Virol 89:6418–6426. doi: 10.1128/JVI.00110-15. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 440.Posthuma CC, te Velthuis AJW, Snijder EJ. 2017. Nidovirus RNA polymerases: complex enzymes handling exceptional RNA genomes. Virus Res 234:58–73. doi: 10.1016/j.virusres.2017.01.023. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 441.Graham RL, Becker MM, Eckerle LD, Bolles M, Denison MR, Baric RS. 2012. A live, impaired-fidelity coronavirus vaccine protects in an aged, immunocompromised mouse model of lethal disease. Nat Med 18:1820–1826. doi: 10.1038/nm.2972. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 442.Makino S, Keck JG, Stohlman SA, Lai MMC. 1986. High-frequency RNA recombination of murine coronaviruses. J Virol 57:729–737. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 443.Baric RS, Fu K, Schaad MC, Stohlman SA. 1990. Establishing a genetic recombination map for murine coronavirus strain A59 complementation groups. Virology 177:646–656. doi: 10.1016/0042-6822(90)90530-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 444.Gorbalenya AE. 2008. Genomics and evolution of the Nidovirales, p 15–28. In Perlman S, Gallagher T, Snijder EJ (ed), Nidoviruses. ASM Press, Washington, DC. [Google Scholar]
  • 445.Denison MR, Graham RL, Donaldson EF, Eckerle LD, Baric RS. 2011. Coronaviruses: an RNA proofreading machine regulates replication fidelity and diversity. RNA Biol 8:270–279. doi: 10.4161/rna.8.2.15013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 446.Taucher C, Berger A, Mandl CW. 2010. A trans-complementing recombination trap demonstrates a low propensity of flaviviruses for intermolecular recombination. J Virol 84:599–611. doi: 10.1128/JVI.01063-09. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 447.Allison R, Thompson C, Ahlquist P. 1990. Regeneration of a functional RNA virus genome by recombination between deletion mutants and requirement for cowpea chlorotic mottle virus 3a and coat genes for systemic infection. Proc Natl Αcad Sci U S A 87:1820–1824. doi: 10.1073/pnas.87.5.1820. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 448.Zwart MP, Willemsen A, Daros JA, Elena SF. 2014. Experimental evolution of pseudogenization and gene loss in a plant RNA virus. Mol Biol Evol 31:121–134. doi: 10.1093/molbev/mst175. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 449.Willemsen A, Zwart MP, Higueras P, Sardanyes J, Elena SF. 2016. Predicting the stability of homologous gene duplications in a plant RNA virus. Genome Biol Evol 8:3065–3082. doi: 10.1093/gbe/evw219. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 450.Palasingam K, Shaklee PN. 1992. Reversion of Qβ RNA phage mutants by homologous RNA recombination. J Virol 66:2435–2442. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 451.Olsthoorn RCL, van Duin J. 1996. Evolutionary reconstruction of a hairpin deleted from the genome of an RNA virus. Proc Natl Acad Sci U S A 93:12256–12261. doi: 10.1073/pnas.93.22.12256. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 452.Di H, Madden JC, Morantz EK, Tang HY, Graham RL, Baric RS, Brinton MA. 2017. Expanded subgenomic mRNA transcriptome and coding capacity of a nidovirus. Proc Natl Acad Sci U S A 114:E8895–E8904. doi: 10.1073/pnas.1706696114. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 453.Kulasegaran-Shylini R, Atasheva S, Gorenstein DG, Frolov I. 2009. Structural and functional elements of the promoter encoded by the 5′ untranslated region of the Venezuelan equine encephalitis virus genome. J Virol 83:8327–8339. doi: 10.1128/JVI.00586-09. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 454.Michel G, Petrakova O, Atasheva S, Frolov I. 2007. Adaptation of Venezuelan equine encephalitis virus lacking 51-nt conserved sequence element to replication in mammalian and mosquito cells. Virology 362:475–487. doi: 10.1016/j.virol.2007.01.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 455.Jiang Y, Cheng CP, Serviene E, Shapka N, Nagy PD. 2010. Repair of lost 5′ terminal sequences in tombusviruses: rapid recovery of promoter- and enhancer-like sequences in recombinant RNAs. Virology 404:96–105. doi: 10.1016/j.virol.2010.04.025. [DOI] [PubMed] [Google Scholar]
  • 456.Simon-Buela L, Osaba L, Garcia JA, Lopez-Moya JJ. 2000. Preservation of 5′-end integrity of a potyvirus genomic RNA is not dependent on template specificity. Virology 269:377–382. doi: 10.1006/viro.2000.0229. [DOI] [PubMed] [Google Scholar]
  • 457.Kuhn RJ, Hong Z, Strauss JH. 1990. Mutagenesis of the 3′ nontranslated region of Sindbis virus RNA. J Virol 64:1465–1476. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 458.Hill KR, Hajjou M, Hu JY, Raju R. 1997. RNA-RNA recombination in Sindbis virus: roles of the 3′ conserved motif, poly(A) tail, and nonviral sequences of template RNAs in polymerase recognition and template switching. J Virol 71:2693–2704. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 459.Raju R, Hajjou M, Hill KR, Botta V, Botta S. 1999. In vivo addition of poly(A) tail and AU-rich sequences to the 3′ terminus of the Sindbis virus RNA genome: a novel 3′-end repair pathway. J Virol 73:2410–2419. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 460.Men RH, Bray M, Clark D, Chanock RM, Lai CJ. 1996. Dengue type 4 virus mutants containing deletions in the 3′ noncoding region of the RNA genome: analysis of growth restriction in cell culture and altered viremia pattern and immunogenicity in rhesus monkeys. J Virol 70:3930–3937. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 461.van Leeuwen HC, Liefhebber JMP, Spaan WJA. 2006. Repair and polyadenylation of a naturally occurring hepatitis C virus 3′ nontranslated region-shorter variant in selectable replicon cell lines. J Virol 80:4336–4343. doi: 10.1128/JVI.80.9.4336-4343.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 462.Eyre NS, Johnson SM, Eltahla AA, Aloi M, Aloia AL, McDevitt CA, Bull RA, Beard MR. 2017. Genome-wide mutagenesis of dengue virus reveals plasticity of the NS1 protein and enables generation of infectious tagged reporter viruses. J Virol 91:e01455-17. doi: 10.1128/JVI.01455-17. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 463.Zust R, Miller TB, Goebel SJ, Thiel V, Masters PS. 2008. Genetic interactions between an essential 3′ cis-acting RNA pseudoknot, replicase gene products, and the extreme 3′ end of the mouse coronavirus genome. J Virol 82:1214–1228. doi: 10.1128/JVI.01690-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 464.Liu PH, Yang D, Carter K, Masud F, Leibowitz JL. 2013. Functional analysis of the stem loop S3 and S4 structures in the coronavirus 3′ UTR. Virology 443:40–47. doi: 10.1016/j.virol.2013.04.021. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 465.Peng YH, Lin CH, Lin CN, Lo CY, Tsai TL, Wu HY. 2016. Characterization of the role of hexamer AGUAAA and poly(A) tail in coronavirus polyadenylation. PLoS One 11:e0165077. doi: 10.1371/journal.pone.0165077. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 466.Eggen R, Verver J, Wellink J, De Jong A, Goldbach R, van Kammen A. 1989. Improvements of the infectivity of in vitro transcripts from cloned cowpea mosaic virus cDNA: impact of terminal nucleotide sequences. Virology 173:447–455. doi: 10.1016/0042-6822(89)90557-6. [DOI] [PubMed] [Google Scholar]
  • 467.Jupin I, Bouzoubaa S, Richards K, Jonard G, Guilley H. 1990. Multiplication of beet necrotic yellow vein virus RNA 3 lacking a 3′ poly(A) tail is accompanied by reappearance of the poly(A) tail and a novel short U-rich tract preceding it. Virology 178:281–284. doi: 10.1016/0042-6822(90)90404-F. [DOI] [PubMed] [Google Scholar]
  • 468.Guilford PJ, Beck DL, Forster RLS. 1991. Influence of the poly(A) tail and putative polyadenylation signal on the infectivity of white clover mosaic potexvirus. Virology 182:61–67. doi: 10.1016/0042-6822(91)90648-U. [DOI] [PubMed] [Google Scholar]
  • 469.Tacahashi Y, Uyeda I. 1999. Restoration of the 3′ end of potyvirus RNA derived from poly(A)-deficient infectious cDNA clones. Virology 265:147–152. doi: 10.1006/viro.1999.0027. [DOI] [PubMed] [Google Scholar]
  • 470.Wang CC, Hsu YC, Wu HC, Wu HN. 2017. Insights into the coordinated interplay of the sHP hairpin and its co-existing and mutually-exclusive dengue virus terminal RNA elements for viral replication. Virology 505:56–70. doi: 10.1016/j.virol.2017.02.007. [DOI] [PubMed] [Google Scholar]
  • 471.Basu M, Brinton MA. 2011. West Nile virus (WNV) genome RNAs with up to three adjacent mutations that disrupt long distance 5′-3′ cyclization sequence basepairs are viable. Virology 412:220–232. doi: 10.1016/j.virol.2011.01.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 472.Dreher TW. 2009. Role of tRNA-like structures in controlling plant virus replication. Virus Res 139:217–229. doi: 10.1016/j.virusres.2008.06.010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 473.Rao ALN, Kao CC. 2015. The brome mosaic virus 3′ untranslated sequence regulates RNA replication, recombination, and virion assembly. Virus Res 206:46–52. doi: 10.1016/j.virusres.2015.02.007. [DOI] [PubMed] [Google Scholar]
  • 474.Dreher TW, Goodwin JB. 1998. Transfer RNA mimicry among tymoviral genomic RNAs ranges from highly efficient to vestigial. Nucleic Acids Res 26:4356–4364. doi: 10.1093/nar/26.19.4356. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 475.Bujarski JJ, Kaesberg P. 1986. Genetic recombination between RNA components of a multipartite plant virus. Nature 321:528–531. doi: 10.1038/321528a0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 476.Rao ALN, Dreher TW, Marsh LE, Hall TC. 1989. Telomeric function of the tRNA-like structure of brome mosaic virus RNA. Proc Natl Acad Sci U S A 86:5335–5339. doi: 10.1073/pnas.86.14.5335. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 477.Hema M, Gopinath K, Kao C. 2005. Repair of the tRNA-like CCA sequence in a multipartite positive-strand RNA virus. J Virol 79:1417–1427. doi: 10.1128/JVI.79.3.1417-1427.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 478.Rao ALN. 2006. Sensitivity of brome mosaic virus RNA1 replication to mutations in the 3′ tRNA-like structure implies a requirement for sustained synthesis of replicase protein 1a. Arch Virol 151:721–733. doi: 10.1007/s00705-005-0658-y. [DOI] [PubMed] [Google Scholar]
  • 479.Dalmay T, Russo M, Burgyan J. 1993. Repair in vivo of altered 3′-terminus of cymbidium ringspot tombusvirus RNA. Virology 192:551–555. doi: 10.1006/viro.1993.1071. [DOI] [PubMed] [Google Scholar]
  • 480.Dalmay T, Rubino L. 1995. Replication of cymbidium ringspot virus satellite RNA mutants. Virology 206:1092–1098. doi: 10.1006/viro.1995.1032. [DOI] [PubMed] [Google Scholar]
  • 481.Nagy PD, Carpenter CD, Simon AE. 1997. A novel 3′-end repair mechanism in an RNA virus. Proc Natl Acad Sci U S A 94:1113–1118. doi: 10.1073/pnas.94.4.1113. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 482.Burgyan J, Garcia-Arenal F. 1998. Template-independent repair of the 3′ end of cucumber mosaic virus satellite RNA controlled by RNAs 1 and 2 of helper virus. J Virol 72:5061–5066. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 483.Guan H, Simon AE. 2000. Polymerization of nontemplate bases before transcription initiation at the 3′ ends of templates by an RNA-dependent RNA polymerase: an activity involved in 3′ end repair of viral RNAs. Proc Natl Acad Sci U S A 97:12451–12456. doi: 10.1073/pnas.97.23.12451. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 484.Carpenter CD, Simon AE. 1998. Analysis of sequences and predicted structures required for viral satellite RNA accumulation by in vivo genetic selection. Nucleic Acids Res 26:2426–2432. doi: 10.1093/nar/26.10.2426. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 485.Carpenter CD, Simon AE. 1996. In vivo restoration of biologically active 3′ ends of virus-associated RNAs by nonhomologous RNA recombination and replacement of a terminal motif. J Virol 70:478–486. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 486.Carpenter CD, Simon AE. 1996. In vivo repair of 3′-end deletions in a TCV satellite RNA may involve two abortive synthesis and priming events. Virology 226:153–160. doi: 10.1006/viro.1996.0641. [DOI] [PubMed] [Google Scholar]
  • 487.Sanjuan R, Domingo-Calap P. 2016. Mechanisms of viral mutation. Cell Mol Life Sci 73:4433–4448. doi: 10.1007/s00018-016-2299-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 488.Boni MF, Zhou Y, Taubenberger JK, Holmes EC. 2008. Homologous recombination is very rare or absent in human influenza A virus. J Virol 82:4807–4811. doi: 10.1128/JVI.02683-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 489.Han GZ, Worobey M. 2011. Homologous recombination in negative sense RNA viruses. Viruses 3:1358–1373. doi: 10.3390/v3081358. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 490.Reguera J, Cusack S, Kolakofsky D. 2014. Segmented negative strand RNA virus nucleoprotein structure. Curr Opin Virol 5:7–15. doi: 10.1016/j.coviro.2014.01.003. [DOI] [PubMed] [Google Scholar]
  • 491.Qin ZM, Sun L, Ma BC, Cui ZZ, Zhu YP, Kitamura Y, Liu WJ. 2008. F gene recombination between genotype II and VII Newcastle disease virus. Virus Res 131:299–303. doi: 10.1016/j.virusres.2007.10.001. [DOI] [PubMed] [Google Scholar]
  • 492.He CQ, Xie ZX, Han GZ, Dong JB, Wang D, Liu JB, Ma LY, Tang XF, Liu XP, Pang YS, Li GR. 2009. Homologous recombination as an evolutionary force in the avian influenza A virus. Mol Biol Evol 26:177–187. doi: 10.1093/molbev/msn238. [DOI] [PubMed] [Google Scholar]
  • 493.He CQ, Meng SL, Yan HY, Ding NZ, He HB, Yan JX, Xu GL. 2012. Isolation and identification of a novel rabies virus lineage in China with natural recombinant nucleoprotein gene. PLoS One 7:e49992. doi: 10.1371/journal.pone.0049992. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 494.He CQ, Ding NZ. 2012. Discovery of severe fever with thrombocytopenia syndrome bunyavirus strains originating from intragenic recombination. J Virol 86:12426–12430. doi: 10.1128/JVI.01317-12. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 495.Chong YL, Padhi A, Hudson PJ, Poss M. 2010. The effect of vaccination on the evolution and population dynamics of avian paramyxovirus 1. PLoS Pathog 6:e1000872. doi: 10.1371/journal.ppat.1000872. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 496.Zhang R, Wang XT, Su JL, Zhao JX, Zhang GZ. 2010. Isolation and analysis of two naturally-occurring multi-recombination Newcastle disease viruses in China. Virus Res 151:45–53. doi: 10.1016/j.virusres.2010.03.015. [DOI] [PubMed] [Google Scholar]
  • 497.Lukashev AN, Klimentov AS, Smirnova SE, Dzagurova TK, Drexler JF, Gmyl AP. 2016. Phylogeography of Crimean Congo hemorrhagic fever virus. PLoS One 11:e0166744. doi: 10.1371/journal.pone.0166744. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 498.Lee SH, Kim WK, No JS, Kim JA, Il Kim J, Gu SH, Kim HC, Klein TA, Park MS, Song JW. 2017. Dynamic circulation and genetic exchange of a shrew-borne hantavirus, Imjin virus, in the Republic of Korea. Sci Rep 7:44369. doi: 10.1038/srep44369. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 499.He M, Guan SY, He CQ. 2017. Evolution of rice stripe virus. Mol Phylogenet Evol 109:343–350. doi: 10.1016/j.ympev.2017.02.002. [DOI] [PubMed] [Google Scholar]
  • 500.Cao DJ, Barroj M, Hoshino Y. 2008. Porcine rotavirus bearing an aberrant gene stemming from an intergenic recombination of the NSP2 and NSP5 genes is defective and interfering. J Virol 82:6073–6077. doi: 10.1128/JVI.00121-08. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 501.He CQ, Ding NZ, He M, Li SN, Wang XM, He HB, Liu XF, Guo HS. 2010. Intragenic recombination as a mechanism of genetic diversity in bluetongue virus. J Virol 84:11487–11495. doi: 10.1128/JVI.00889-10. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 502.Woods RJ. 2015. Intrasegmental recombination does not contribute to the long-term evolution of group A rotavirus. Infect Genet Evol 32:354–360. doi: 10.1016/j.meegid.2015.03.035. [DOI] [PubMed] [Google Scholar]
  • 503.Esona MD, Roy S, Rungsrisuriyachai K, Sanchez J, Vasquez L, Gomez V, Rios LA, Bowen MD, Vazquez M. 2017. Characterization of a triple-recombinant, reassortant rotavirus strain from the Dominican Republic. J Gen Virol 98:134–142. doi: 10.1099/jgv.0.000688. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 504.Treanor JJ, Buja R, Murphy BR. 1991. Intragenic suppression of a deletion mutation of the nonstructural gene of an influenza A virus. J Virol 65:4204–4210. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 505.Stojdl DF, Lichty BD, ten Oever BR, Paterson JM, Power AT, Knowles S, Marius R, Reynard J, Poliquin L, Atkins H, Brown EG, Durbin RK, Durbin JE, Hiscott J, Bell JC. 2003. VSV strains with defects in their ability to shutdown innate immunity are potent systemic anti-cancer agents. Cancer Cell 4:263–275. doi: 10.1016/S1535-6108(03)00241-1. [DOI] [PubMed] [Google Scholar]
  • 506.Garijo R, Cuevas JM, Briz A, Sanjuan R. 2016. Constrained evolvability of interferon suppression in an RNA virus. Sci Rep 6:24722. doi: 10.1038/srep24722. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 507.Le Nouën C, McCarty T, Brown M, Smith ML, Lleras R, Dolan MA, Mehedi M, Yang L, Luongo C, Liang B, Munir S, DiNapoli JM, Mueller S, Wimmer E, Collins PL, Buchholz UJ. 2017. Genetic stability of genome-scale deoptimized RNA virus vaccine candidates under selective pressure. Proc Natl Acad Sci U S A 114:E386–E395. doi: 10.1073/pnas.1619242114. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 508.McDonald SM, Nelson MI, Turner PE, Patton JT. 2016. Reassortment in segmented RNA viruses: mechanisms and outcomes. Nat Rev Microbiol 14:448–460. doi: 10.1038/nrmicro.2016.46. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 509.Geoghegan JL, Duchene S, Holmes EC. 2017. Comparative analysis estimates the relative frequencies of co-divergence and cross-species transmission within viral families. PLoS Pathog 13:e1006215. doi: 10.1371/journal.ppat.1006215. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 510.de Visser JA, Hermisson J, Wagner GP, Meyers LA, Bagheri-Chaichian H, Blanchard JL, Chao L, Cheverud JM, Elena SF, Fontana W, Gibson G, Hansen TF, Krakauer D, Lewontin RC, Ofria C, Rice SH, von Dassow G, Wagner A, Whitlock MC. 2003. Perspective: evolution and detection of genetic robustness. Evolution 57:1959–1972. [DOI] [PubMed] [Google Scholar]
  • 511.Pierangeli A, Bucci M, Pagnotti P, Degener AM, Perez Bercoff R. 1995. Mutational analysis of the 3′-terminal extra-cistronic region of poliovirus RNA: secondary structure is not the only requirement for minus strand RNA replication. FEBS Lett 374:327–332. doi: 10.1016/0014-5793(95)01127-Z. [DOI] [PubMed] [Google Scholar]
  • 512.Melchers WJG, Bakkers JMJE, Bruins Slot HJ, Galama JMD, Agol VI, Pilipenko EV. 2000. Cross-talk between orientation-dependent recognition determinants of a complex control RNA element, the enterovirus oriR. RNA 6:976–987. doi: 10.1017/S1355838200000480. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 513.Agol VI, Paul AV, Wimmer E. 1999. Paradoxes of the replication of picornaviral genomes. Virus Res 62:129–147. doi: 10.1016/S0168-1702(99)00037-4. [DOI] [PubMed] [Google Scholar]
  • 514.Burch CL, Chao L. 1999. Evolution by small steps and rugged landscapes in the RNA virus phi6. Genetics 151:921–927. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 515.Steinhauer DA, Holland JJ. 1987. Rapid evolution of RNA viruses. Annu Rev Microbiol 41:409–433. doi: 10.1146/annurev.mi.41.100187.002205. [DOI] [PubMed] [Google Scholar]
  • 516.Wagner A, Stadler PF. 1999. Viral RNA and evolved mutational robustness. J Exp Zool 285:119–127. doi:. [DOI] [PubMed] [Google Scholar]
  • 517.Wagner A. 2008. Robustness and evolvability: a paradox resolved. Proc Biol Sci 275:91–100. doi: 10.1098/rspb.2007.1137. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 518.Cuevas JM, Moya A, Sanjuan R. 2009. A genetic background with low mutational robustness is associated with increased adaptability to a novel host in an RNA virus. J Evol Biol 22:2041–2048. doi: 10.1111/j.1420-9101.2009.01817.x. [DOI] [PubMed] [Google Scholar]
  • 519.Draghi JA, Parsons TL, Wagner GP, Plotkin JB. 2010. Mutational robustness can facilitate adaptation. Nature 463:353–355. doi: 10.1038/nature08694. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 520.Elena SF. 2012. RNA virus genetic robustness: possible causes and some consequences. Curr Opin Virol 2:525–530. doi: 10.1016/j.coviro.2012.06.008. [DOI] [PubMed] [Google Scholar]
  • 521.Lauring AS, Frydman J, Andino R. 2013. The role of mutational robustness in RNA virus evolution. Nat Rev Microbiol 11:327–336. doi: 10.1038/nrmicro3003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 522.Goldhill D, Lee A, Williams ESCP, Turner PE. 2014. Evolvability and robustness in populations of RNA virus ϕ6. Front Microbiol 5:35. doi: 10.3389/fmicb.2014.00035. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 523.Stern A, Bianco S, Yeh MT, Wright C, Butcher K, Tang C, Nielsen R, Andino R. 2014. Costs and benefits of mutational robustness in RNA viruses. Cell Rep 8:1026–1036. doi: 10.1016/j.celrep.2014.07.011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 524.Cervera H, Lalic J, Elena SF. 2016. Efficient escape from local optima in a highly rugged fitness landscape by evolving RNA virus populations. Proc Biol Sci 283:20160984. doi: 10.1098/rspb.2016.0984. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 525.Sanjuan R. 2016. Viral mutation rates, p 7–27. In Weaver SC, Denison MR, Roossinck MJ, Vignuzzi M (ed), Virus evolution: current research and future directions. Caister Academic Press, Wymondham, United Kingdom. [Google Scholar]
  • 526.Barr JN, Fearns R. 2016. Genetic instability of RNA viruses, p 21–35. In Kovalchuk I, Kovalchuk O (ed), Genome stability from virus to human application. Academic Press, Boston, MA. [Google Scholar]
  • 527.Holmes EC. 2010. The comparative genomics of viral emergence. Proc Natl Acad Sci U S A 107:1742–1746. doi: 10.1073/pnas.0906193106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 528.Elena SF, Fraile A, Garcia-Arenal F. 2014. Evolution and emergence of plant viruses. Adv Virus Res 88:161–191. doi: 10.1016/B978-0-12-800098-4.00003-9. [DOI] [PubMed] [Google Scholar]
  • 529.Longdon B, Brockhurst MA, Russell CA, Welch JJ, Jiggins FM. 2014. The evolution and genetics of virus host shifts. PLoS Pathog 10:e1004395. doi: 10.1371/journal.ppat.1004395. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 530.Bedhomme S, Hillung J, Elena SF. 2015. Emerging viruses: why they are not jacks of all trades? Curr Opin Virol 10:1–6. doi: 10.1016/j.coviro.2014.10.006. [DOI] [PubMed] [Google Scholar]
  • 531.Mandl JN, Ahmed R, Barreiro LB, Daszak P, Epstein JH, Virgin HW, Feinberg MB. 2015. Reservoir host immune responses to emerging zoonotic viruses. Cell 160:20–35. doi: 10.1016/j.cell.2014.12.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 532.Lam TTY, Zhu HC, Guan Y, Holmes EC. 2016. Genomic analysis of the emergence, evolution, and spread of human respiratory RNA viruses. Annu Rev Genomics Hum Genet 17:193–218. doi: 10.1146/annurev-genom-083115-022628. [DOI] [PubMed] [Google Scholar]
  • 533.Graci JD, Cameron CE. 2002. Quasispecies, error catastrophe, and the antiviral activity of ribavirin. Virology 298:175–180. doi: 10.1006/viro.2002.1487. [DOI] [PubMed] [Google Scholar]
  • 534.Vignuzzi M, Stone JK, Andino R. 2005. Ribavirin and lethal mutagenesis of poliovirus: molecular mechanisms, resistance and biological implications. Virus Res 107:173–181. doi: 10.1016/j.virusres.2004.11.007. [DOI] [PubMed] [Google Scholar]
  • 535.Ojosnegros S, Perales C, Mas A, Domingo E. 2011. Quasispecies as a matter of fact: viruses and beyond. Virus Res 162:203–215. doi: 10.1016/j.virusres.2011.09.018. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 536.Presloid JB, Novella IS. 2015. RNA viruses and RNAi: quasispecies implications for viral escape. Viruses 7:3226–3240. doi: 10.3390/v7062768. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 537.Perales C, Domingo E. 2016. Antiviral strategies based on lethal mutagenesis and error threshold. Curr Top Microbiol Immunol 392:323–339. doi: 10.1007/82_2015_459. [DOI] [PubMed] [Google Scholar]

Articles from Microbiology and Molecular Biology Reviews : MMBR are provided here courtesy of American Society for Microbiology (ASM)

RESOURCES