VI01CH06-Sullivan ARI 2 September 2014 6:40

Balance and Stealth: The Role of Noncoding in the Regulation of Expression

Jennifer E. Cox and Christopher S. Sullivan

Department of Molecular Biosciences, The University of Texas at Austin, Austin, Texas 78712; email: [email protected]

Annu. Rev. Virol. 2014. 1:89–109 Keywords First published online as a Review in Advance on , autoregulation, latency, persistence, reactivation, stress June 2, 2014

The Annual Review of Virology is online at Abstract Access provided by 45.16.157.220 on 10/02/15. For personal use only. virology.annualreviews.org In the past two decades, our knowledge of gene regulation has been greatly This article’s doi:

Annual Review of Virology 2014.1:89-109. Downloaded from www.annualreviews.org expanded by the discovery of microRNAs (miRNAs). miRNAs are small (19– 10.1146/annurev-virology-031413-085439 24 nt) noncoding RNAs (ncRNAs) found in metazoans, plants, and some Copyright c 2014 by Annual Reviews. . They have been shown to regulate many cellular processes, includ- All rights reserved ing differentiation, maintenance of homeostasis, apoptosis, and the immune response. At present, there are over 300 known viral miRNAs encoded by di- verse virus families. One well-characterized function of some viral miRNAs is the regulation of viral transcripts. Host miRNAs can also regulate viral gene expression. We propose that viruses take advantage of both host and viral ncRNA regulation to balance replication and infectious state (for ex- ample, latent versus lytic infection). As miRNA regulation can be reversed upon certain cellular stresses, we hypothesize that ncRNAs can serve viruses as barometers for cellular stress.

89 VI01CH06-Sullivan ARI 2 September 2014 6:40

INTRODUCTION The “grandmother of punk,” American singer-songwriter Patti Smith, once said: “In life may Seed sequence: you proceed with balance and stealth.” Although the context differs, this sentiment accurately nucleotides 2–8 of a describes the challenges that face viruses that undergo persistent infections. That is, they must have mature miRNA; life cycles with balanced gene expression, maintained through stealthful mechanisms. Persistent particularly important viruses must optimize when, where, and how many virions are produced, while keeping host in determining target interactions cells alive and avoiding the immune response. A major tool viruses use to accomplish this is the noncoding RNAs. Here we review the current understanding of both host and viral non-protein- coding regulatory RNAs (ncRNAs) in the regulation of virus gene expression. Over the past decade there has been an exponential increase in our understanding of the role of ncRNAs in diverse areas of biology. There is an existing or emerging understanding of the mechanisms of action of several classes of ncRNA, including microRNAs (miRNAs) and long noncoding RNAs (lncRNAs). miRNAs are small RNAs (typically ∼22 nt) that repress gene ex- pression by directing the RNA-induced silencing complex (RISC) to target mRNA transcripts, thereby inhibiting and/or decreasing the abundance of bound transcripts (1, 2). Unlike miRNAs, lncRNAs comprise a heterogeneous group of RNAs (typically >200 nt) whose biogen- esis and mode of action are poorly defined. Regardless, some lncRNAs have been implicated in transcriptional and posttranscriptional control of gene expression (3). New classes of ncRNAs and associated functionalities will continue to emerge with the increased use of high-throughput RNA sequencing. Like other fields, virology has been greatly impacted by ncRNAs. Both virus- and host-encoded ncRNAs contribute to the regulation of virus gene expression and the host response to infection. At least nine different virus families, spanning much of the Baltimore classification scheme, encode ncRNAs (4–8). ncRNAs are most prevalent in DNA viruses, but some retroviruses as well as negative- and positive-strand RNA viruses also give rise to ncRNAs (8, 9; R.P. Kincaid, Y. Chen, J.E. Cox, A. Rethwilm & C.S. Sullivan, unpublished observations). Although these ncRNAs can be diverse in terms of sequence, biogenesis, size, and structure, some shared activities are common. These include evasion of host defenses and regulation of virus gene expression (10). The role of ncRNAs in host defenses against virus infection has been extensively reviewed (6, 11–15). Here, we focus on ncRNAs that regulate virus gene expression.

MicroRNAs Access provided by 45.16.157.220 on 10/02/15. For personal use only. miRNAs have been implicated in numerous virus-relevant processes, including the immune re- sponse, cell viability, and tumorigenesis (16–18). In the canonical pathway, miRNAs are derived Annual Review of Virology 2014.1:89-109. Downloaded from www.annualreviews.org from a primary transcript containing an approximately 70-nt stem-loop structure that is processed by the nuclear endonuclease Drosha and then the cytoplasmic endonuclease Dicer to give rise to the final ∼22-nt effector molecule (19). In general, miRNAs function as rheostats that serve to dampen gene expression (20). A single miRNA can have hundreds of different targets; conversely, a single transcript can have numerous docking sites for the same or multiple different miRNAs. In this manner, a complex web of potential interactions comes together to posttranscriptionally fine-tune gene expression (6). miRNAs typically bind to the 3 untranslated region (3 UTR) of their target transcripts with imperfect complementarity. Commonly, this interaction is mediated in large part by the seed sequence (nucleotides 2–8), which directly binds to the target transcript via Watson–Crick base-pairing (21, 22). This results in translational repression, often accompanied by enhanced turnover of targeted transcripts. Less frequently, miRNAs can bind with perfect complementarity

90 Cox · Sullivan VI01CH06-Sullivan ARI 2 September 2014 6:40

PERSISTENCE AND LATENCY

Following de novo infection, some viruses can establish long-term persistent infections. Often this mode of infection is cell type–specific. Latency represents a specific subtype of persistent infection, defined as an alternative gene expression program that does not result in the production of virus (38). A defining characteristic of latency is that it can be reversed to allow full virus gene expression and production of virions (38). Persistent infections are maintained through modulation of both viral and cellular gene expression. Currently, there is a limited understanding of the mechanisms that control the maintenance of persistence and the switch to lytic infection. As many virus-associated diseases (for example, some cancers) are a consequence of persistent infection or escape from persistence, it is imperative to better understand the mechanisms of viral persistence. However, many common laboratory models of virus infection have been selected for the ability of the virus to undergo robust lytic infections and thus may be inappropriate for modeling persistence. Thus, future work on existing models is warranted, as is the development of new laboratory models of persistent infection.

(∼22 out of 22 nt) and direct an irreversible cleavage of the target transcript, similar to siRNAs. This mode of regulation readily occurs in plants but is rarely seen in animals. However, some an- imal viral miRNAs direct cleavage of their antisense transcripts through this mechanism (23–27). It is interesting to note that miRNA silencing activity can be attenuated or even inverted in some circumstances, such as stress (28–33) (discussed below). In most cases, any target of a miRNA is only subtly regulated by a single miRNA-target interaction. This suggests that the sum of numerous subtle regulatory events accounts for the profound biological effects often associated with miRNAs (34). As such, miRNAs can serve to balance “transcriptional noise” and maintain homeostasis (34–37). Thus, global perturbations of miRNA activity can be expected to reduce the tolerance of any given cell to various stressors. To date, over 300 viral miRNAs have been discovered from diverse viruses (6, 38). Most of these viruses produce long-term, often lifelong, persistent infections (see sidebar, Persistence and Latency). miRNAs offer several advantages to a virus, including a small genomic footprint (most miRNAs can be encoded in less than 100 nt of genomic space) and presumed invisibility to the protein-based adaptive immune response. Although some viral miRNAs serve to block host defenses (including apoptosis and the adaptive and innate immune responses), an emerging concept is that many viral miRNAs regulate virus gene expression (10, 39, 40). This is of particular

Access provided by 45.16.157.220 on 10/02/15. For personal use only. importance to those viruses that undergo persistent infections, as they need to optimize gene expression to allow for appropriate induction of the lytic, productive phase of infection. Below, we present specific examples whereby diverse viruses utilize viral or host miRNAs to control virus Annual Review of Virology 2014.1:89-109. Downloaded from www.annualreviews.org gene expression. For a comprehensive overview of ncRNA regulation of viral transcripts, refer to Table 1.

VIRAL MicroRNAs

Herpesviruses Herpesviruses have large, double-stranded DNA genomes, typically greater than 100 kb, that code for numerous (>65) gene products. Herpesviruses can undergo latent infections in long- lived cells such as neurons or immune effector cells (the cell type depends on the infecting virus). Lytic infection proceeds via a temporally coordinated gene expression cascade from immediate early through late . Lytic infection can occur upon de novo infection or via exit from latency.

www.annualreviews.org • Viral MicroRNAs 91 VI01CH06-Sullivan ARI 2 September 2014 6:40

Table 1 Summary of viral and host miRNAs known to regulate viral transcripts Family Virus(es) miRNA Target Reference(s) Viral miRNAs Herpesviridae Epstein–Barr virus (EBV) miR-BART2 BALF5a 26, 77 miR-BART11 EBF1 150 miR-BART3, -BART5, LMP1 77, 127 -BART16, -BART17, -BART19, -BART20 miR-BART22 LMP2A 151 miR-BART10 BHRF1 77 miR-BART3, -BART11, EBNA2a 77 -BART20 Kaposi sarcoma–associated miR-K12-9, -K12-7 RTAa 57, 64 herpesvirus (KSHV) miR-K12-10 V-IL-6, V-FLIP, 57 V-cyclin, V-IRF3, LANA Human cytomegalovirus miR-UL36 UL138 54 (HCMV) miR-UL112 IE1a, UL112/113, 39, 40, 54 UL114, UL120/121 Herpes simplex virus 1 and 2 miR-H2 ICP0a 50, 152 (HSV-1 and -2) miR-H3, -H4 ICP34.5 50, 152, 153 Herpes simplex virus 1 (HSV-1) miR-H6 ICP4 44 Rhesus cytomegalovirus miR-Rh183-1 Rh186 53 (RhCMV) Bovine herpesvirus 1 (BoHV-1) ncRNA 1, 2 bICP0a 154 Gallid herpesvirus 1 (GaHV-1) miR-I5, -I6 ICP4 155 Marek disease virus (MDV) miR-M4 UL28, UL32a 156 miR-M7 ICP4 157 miR-M7 ICP27 157 Rhesus lymphocryptovirus miR-rL1-17, -rL1-5, -rL1-8, LMP1 127 Access provided by 45.16.157.220 on 10/02/15. For personal use only. (rLCV) -rL1-16, -rL1-23 Papillomaviridae Bandicoot papillomatosis miR-B1 Early transcriptsa 89

Annual Review of Virology 2014.1:89-109. Downloaded from www.annualreviews.org carcinomatosis virus type 1 (BPCV1) Bandicoot papillomatosis miR-B1 Early transcriptsa 89 carcinomatosis virus type 2 (BPCV2) Polyomaviridae BK polyomavirus (BKPyV) miR-B1 Early transcriptsa 24, 90 JC polyomavirus ( JCPyV) miR-J1 Early transcriptsa 24 Merkel cell polyomavirus miR-M1 Early transcriptsa 23 (MCPyV) Murine polyomavirus (MPyV) miR-M1 Early transcriptsa 27 Simian agent 12 (SA12) miR-S1 Early transcriptsa 87 Simian virus 40 (SV40) miR-S1 Early transcriptsa 25, 88 (Continued )

92 Cox · Sullivan VI01CH06-Sullivan ARI 2 September 2014 6:40

Table 1 (Continued) Family Virus(es) miRNA Target Reference(s) Viral miRNAs Baculoviridae Bombyx mori nuclear miR-8 IE-1 158 polyhedrosis virus (BmNPV) Ascoviridae Heliothis virescens ascovirus miR-1 DNA polymerase I 159 (HvAV) Unassigned Heliothis zea nudivirus 1 pag1-derived miRNAs hhia 160 (HzNV-1) Nimaviridae White spot syndrome virus miR-66 wsv094, wsv177 161 (WSSV) miR-68 wsv248, wsv309 161 Host miRNAs Human adenovirus type 5 miR-214 E1A 162 (HAdV-5) Hepadnaviridae Hepatitis B virus (HBV) miR-210 Pre-S1 region 164 miR-199a HBsAg 164 Flaviviridae Hepatitis C virus (HCV) miR-122 Binds genome 143 GB virus B (GBV-B) miR-122 Binds genome 163 Herpesviridae Epstein–Barr virus (EBV) miR-17, -142 LMP1, BHRF1 77 miR-7, -15, -103, -185, -747 EBNA2 77 Kaposi sarcoma–associated miR-498, -320d RTAa 165 herpesvirus (KSHV) miR-142 V-IL-6, V-IRF3 57 let-7 V-FLIP, V-cyclin, 57 LANA Nimaviridae White spot syndrome virus miR-7 wsv477 121 (WSSV) Orthomyxoviridae H1N1 influenza A virus let-7c M1 166 Parvoviridae Mink enteritis virus (MEV) miR-181b NS1 167 Picornaviridae Enterovirus 71 (EV71) miR-23b EV71 VPl 168 miR-296-5p Viral genome 169 Coxsackievirus B type 3 (CVB3) miR-342-5p 2C-coding sequence 170 miR-10a 3D-coding region 171

Access provided by 45.16.157.220 on 10/02/15. For personal use only. Retroviridae Avian leucosis virus subgroup J miR-1650 Gag 9 (ALV-J)  Human immunodeficiency virus miR-125b, -28, -150, -223, 3 mRNA 118 Annual Review of Virology 2014.1:89-109. Downloaded from www.annualreviews.org type 1 (HIV-1) -382 miR-29a Nef 117 NEAT1 Replication (increased 119 Rev-independent transport) Simian immunodeficiency virus miR-29a, -29b, -9, -146a Nef/U3 and R regions 172 (SIV) Togaviridae Eastern equine encephalitis virus miR-142 3 untranslated region 128 (EEEV)

aTarget transcripts that could reinforce persistent infection.

www.annualreviews.org • Viral MicroRNAs 93 VI01CH06-Sullivan ARI 2 September 2014 6:40

Optimizing the switch from latent to lytic infection is likely crucial to the fitness of the virus and is controlled by various factors, including chromatin modifications, transcription, and the host adaptive immune response—all of which can be altered by various stressors (41–43). Recently, Pre-miRNA: precursor miRNA that miRNAs have emerged as contributors to latent infection and the regulation of the switch to lytic is cleaved by Dicer to infection in diverse herpesviruses (10). form mature miRNAs Herpes simplex virus 1 (HSV-1) is the prototypic alphaherpesvirus. HSV-1 is associated with AntimiRs: synthetic various diseases including cold sores, genital lesions, and lethal encephalitis. HSV-1 encodes oligonucleotides that more than six pre-miRNAs within the latency-associated transcript (LAT) (44, 45). Interestingly, inhibit a specific the miRNAs and LAT are the only readily detectable transcripts during latency. Several HSV-1 microRNA miRNAs lie antisense (therefore with perfect complementarity) to the immunomodulatory ICP0 and ICP34.5 transcripts. As would be predicted, mutant viruses lacking HSV-1 miRNAs show increased expression of both ICP0 and ICP34.5 (44). ICP0 has also been implicated as necessary for reactivation (46–49). Importantly, infection of cultured Neuro2 cells at a low multiplicity of infection (MOI) with miRNA-mutant viruses resulted in increased viral replication. This phenotype was dependent on both MOI and cell type; infection of Neuro2 cells at a higher MOI or infection of NIH3T3 cells yielded no difference between wild-type and miRNA-mutant viruses (50). This suggests a possible role for the HSV-1 miRNAs in enforcing latency in vivo, although such a hypothesis awaits confirmation. Furthermore, the conditional nature of the increased replication phenotype offers a cautionary note for interpreting those negative studies in which miRNA mutant viruses show no apparent phenotype (25, 27). Human cytomegalovirus (HCMV) is a prevalent human betaherpesvirus that is typically asymp- tomatic in healthy individuals. Infection of the immunocompromised individual or the unborn fetus can result in life-threatening illnesses and birth defects. Unlike other herpesviruses, the betaherpesviruses do not express miRNAs from one or a few clusters but rather have numer- ous distinct miRNA loci (at least 14) dispersed throughout the genome (51). Although multiple HCMV miRNAs can directly regulate various HCMV transcripts, the functional relevance of most of these interactions remains unknown (52, 53). The best-studied HCMV miRNA, miR-UL112, clearly regulates multiple immediate early genes (39, 54). Consequently, ectopic expression of miR-UL112 can negatively regulate virus replication (54). Grey et al. (39) were among the first to demonstrate that miR-UL112 can negatively regulate multiple viral transcripts. Furthermore, engineered mutant viruses (null for either miR-UL112 or one of its direct viral target sites, e.g., IE-1) have altered target viral protein levels (40). These studies were performed in fibroblasts where HCMV undergoes a persistent infection, continuously producing virions. It remains un-

Access provided by 45.16.157.220 on 10/02/15. For personal use only. known what role HCMV miRNAs might play during latent infection; however, miR-UL112 is detectable in at least one model of latent infection (55). Furthermore, infection of a mouse model with a murine cytomegalovirus lacking two miRNAs resulted in less virus production in the sali- Annual Review of Virology 2014.1:89-109. Downloaded from www.annualreviews.org vary glands (56). The phenotype was dependent on both the initial viral titer and the genetic background of the mouse. It will be interesting to test a similar role for HCMV miRNAs in the establishment or maintenance of latency. Kaposi sarcoma–associated herpesvirus (KSHV) is a gammaherpesvirus associated with rare B cell and endothelial tumors in immunosuppressed patients. KSHV encodes 12 pre-miRNAs that have a wealth of established host and viral target transcripts (6, 10, 57–61). Although latency is the typical default pathway for gammaherpesviruses, expression of the immediate early gene product RTA is sufficient to initiate the lytic replication cycle (62). At least three KSHV miRNAs can neg- atively regulate RTA expression via direct binding to its 3 UTR (59, 63, 64). Furthermore, several host targets of KSHV miRNAs have been identified that can affect RTA transcription (65–68). Whereas inhibition of individual KSHV miRNAs via antimiRs can lead to a decreased or increased propensity to initiate lytic replication, deletion of 10 of the 12 pre-miRNAs gives rise to viruses with

94 Cox · Sullivan VI01CH06-Sullivan ARI 2 September 2014 6:40

an increased propensity to undergo lytic activation (64, 67–69). Thus, a model emerges whereby multiple miRNA-target interactions (of viral and host origin) combine to reinforce latent infection. In addition to RTA, cross-linking immunoprecipitation (CLIP) studies have revealed several CLIP: cross-linking other viral targets of KSHV miRNAs, including both lytic (V-IL-6, V-FLIP, V-cyclin, V-IRF3) immunoprecipitation and latent (LANA) transcripts (57, 59). The benefit to the virus of this regulation remains un- RFHVM: known. However, there are at least two examples of seed conservation between a KSHV miRNA retroperitoneal and two closely related rhesus gammaherpesviruses, termed RFHVM and RRV, which suggests fibromatosis that some of this regulation is likely evolutionarily conserved (70, 71). Additional viral transcripts herpesvirus from that may be regulated by miRNAs were uncovered by screening a library of KSHV 3 UTR re- pig-tailed macaque; porters (72). A similar analysis that we conducted revealed an unexpected degree of regulation, related to KSHV with over 50% of the KSHV 3 UTRs conveying significant negative regulation (73). This implies RRV: rhesus a possible global strategy of gene regulation. Interestingly, some of the 3 UTRs convey regu- rhadinovirus lation only in an infectious stage–specific context (that is, only in latent or lytic infection) (73). Although the mechanisms accounting for this stage-specific regulation are mostly unknown, we have characterized one such example. This atypical regulation occurs through Drosha cleavage of pre-miRNAs that are positioned in the 3 UTRs of Kaposin B and K12, as well as the coding portion of K12 (74). Drosha activity is high in latency, resulting in only a small portion of K12 and Kaposin B protein-coding transcripts escaping cleavage. However, during lytic infection, Drosha levels are reduced, with a consequent increase in the transcription of the K12 and Kaposin B loci. Because a positionally conserved pre-miRNA is found in the 3 UTR of the RFHVM K12 transcript, we predict that this mode of cis regulation is likely evolutionarily conserved (71). We note, however, that much of the regulatory potential of the KSHV 3 UTRs is independent of cis or trans effects of the KSHV miRNAs (73). Though well-studied for the RNA viruses, the possible widespread role of UTRs in DNA virus biology remains an exciting and understudied topic. Epstein–Barr virus (EBV) is a gammaherpesvirus prevalent in the human population and associ- ated with tumors, especially in immunosuppressed patients. EBV encodes at least 25 pre-miRNAs derived from one of two separate poly-pre-miRNA clusters (58, 75, 76). In the original description of EBV miRNAs, due to its antisense location, the transcript encoding the viral DNA polymerase (BALF5) was identified as a likely target of EBV miR-BART2 (76). This was later functionally validated by Barth et al. (26). Recently, CLIP strategies on EBV miRNAs have identified nu- merous host and a few viral target transcripts including LMP1, BHRF1, and EBNA2 (57, 77). Interestingly, no lytic transcripts (including BALF5) were identified as EBV miRNA targets, but it remains unclear whether this is due to the low levels of lytic transcripts that “leak” during latency or

Access provided by 45.16.157.220 on 10/02/15. For personal use only. whether EBV miRNAs do not target lytic transcripts in a meaningful way (10). However, transfec- tion of an antimiR against miR-BART6 leads to an approximately 2-fold increase in ZTA and RTA transcripts levels (78). As increased ZTA and RTA could promote lytic activation, these results are Annual Review of Virology 2014.1:89-109. Downloaded from www.annualreviews.org consistent with EBV utilizing miRNAs to optimize the latent/lytic switch. The authors propose that induction of ZTA and RTA is an indirect result of the global miRNA reduction resulting from downregulation of Dicer by miR-BART6 (although it remains possible that other targets could ac- count for this phenotype) (78). Despite these data, a mutant EBV defective for miRNA production did not display increased markers of lytic infection compared with a recombinant virus engineered to express the full repertoire of EBV miRNAs (79). Thus, it remains unknown whether there exists a context in which EBV utilizes viral miRNAs to preferentially reinforce latent over lytic infection.

Polyomaviruses Polyomaviruses (PyVs) are unrelated to herpesviruses in genome and structure, but they may share similar challenges in maintaining long-term persistent infections. These small DNA viruses are

www.annualreviews.org • Viral MicroRNAs 95 VI01CH06-Sullivan ARI 2 September 2014 6:40

prevalent in vertebrates, and at least 12 infect humans (80, 81). Although typically asymptomatic, in immunosuppressed patients some PyVs have a clear association with rare but serious diseases. These include Merkel cell carcinoma, progressive multifocal leukoencephalopathy, nephropathy, Autoregulation: regulation of a viral organ rejection in transplant patients, and trichodysplasia spinulosa (82–85). The prototypic PyVs transcript through a include murine polyomavirus (MPyV) and simian virus 40 (SV40) (86). Both of these emerged viral gene product as laboratory models due in part to their selected ability to undergo robust lytic infection in cultured cells. To date, virus-encoded miRNAs have been identified from at least six different PyVs (23–25, 27, 87). All are found antisense with perfect complementarity to the early transcripts that give rise to the multifunctional tumor antigens (T antigens). As would be predicted from their location, these miRNAs can direct siRNA-like cleavage of the early transcripts (23–25, 27). Although there are undoubtedly some important host targets, these transcripts differ among the PyVs, because the seed regions of most PyV miRNAs are not shared. This does not rule out a possible convergence to regulate similar pathways, as has been suggested for KSHV and EBV (57). As in the herpesviruses, the genomic positions of the PyV miRNAs are often conserved. Consequently, the ability to negatively regulate the early transcripts is also conserved. Two lines of evidence support the importance of miRNA-mediated autoregulation of early gene expression. First, closely related SV40 isolates can have miRNAs with different seed sequences (and therefore different host targets) while preserving similar regulatory activity on the early transcripts (88). Second, the PyV-related bandicoot papillomatosis carcinomatosis viruses (BPCVs) also encode miRNAs that autoregulate the early T antigen transcripts. However, unlike PyV miRNAs, BPCV miRNAs are not located antisense to and do not bind with perfect complementarity to the early transcripts (89). These findings support evolutionary pressure to regulate T antigen transcripts via miRNAs, as opposed to an alternative model of “tolerating” some degree of negative regulation as a necessary consequence of the antisense genomic location. Although the biological relevance of PyV miRNA regulation remains obscure, three studies implicate a role for the miRNAs in negatively regulating virus copy number (27, 90, 91). Thus, two non–mutually exclusive models emerge whereby miRNA-mediated autoregulation of the T antigen transcripts contributes to either (a) lytic infection and/or (b) persistent infection (88). It has been observed that a substantial fraction of the early transcripts are subjected to miRNA- mediated cleavage in lytic models, which suggests a role for autoregulation (25, 88). However, we note that in some cell culture models, no difference in virus replication for miRNA-mutant viruses is observed (25, 88). There is also evidence that suggests a role for autoregulation in persistence of both BK virus and SV40. The BK virus miRNA can limit viral replication in a cultured cell model,

Access provided by 45.16.157.220 on 10/02/15. For personal use only. and the SV40 miRNA affects virus copy number during long periods of in vivo infection (25, 90, 91). It is currently unknown whether PyVs maintain long-term persistence via true latency (that is, via alternative and reversible gene expression profiles). However, as for the herpesviruses, these Annual Review of Virology 2014.1:89-109. Downloaded from www.annualreviews.org data argue that PyV miRNAs play a role in persistent infection. As more is understood about PyV persistent infection, it will be interesting to determine whether the PyV miRNAs participate in any hypothetical herpesvirus-like switch to full-on lytic infection.

Retroviruses Like herpesviruses and polyomaviruses, some retroviruses cause long-term infections and encode miRNAs. Retroviruses generally possess RNA genomes, but they have unique replication cycles in- volving integration of a DNA form of the genome (92). In some instances, this integrated provirus can become transcriptionally silenced. When this occurs in a reactivatable state, it represents a true latent infection. At least three diverse retroviruses encode miRNAs [bovine leukemia virus (BLV), avian leucosis virus subgroup J (ALV-J), and some foamy viruses (FVs) (8, 9; R.P. Kincaid,

96 Cox · Sullivan VI01CH06-Sullivan ARI 2 September 2014 6:40

Y. Chen, J.E. Cox, A. Rethwilm & C.S. Sullivan, unpublished observations)], but some do not (8, 9, 93). The best characterized retroviral miRNAs include the BLV miRNA B4, which mimics the oncogenic host miRNA miR-29, and the African green monkey simian foamy virus (SFVagm) miRNAs S4 and S6, which mimic the host miRNAs miR-155 and miR-132, respectively. All three of these host miRNAs can be prosurvival and/or immunoevasive. Interestingly, the homologous genomic region coding for the majority of SFV miRNAs is sometimes lost when the prototypic virus is cultured ex vivo in cell culture (94–96). These results suggest a possible autoregulatory role for at least some retroviral miRNAs and raise the possibility that this regulation is advantageous in maintaining persistent infection in vivo (R.P. Kincaid, Y. Chen, J.E. Cox, A. Rethwilm & C.S. Sullivan, unpublished observations).

ATTENUATED MicroRNA REGULATION DURING STRESS As multiple virus families utilize miRNAs to control viral gene expression, it is interesting to note that miRNA-mediated regulation can be attenuated or even reversed during certain forms of stress. Cellular stressors that can foster reduced miRNA activity include starvation, amino acid depletion, endoplasmic reticulum stress, oxidative stress, and detection of intracellular or extracellular pathogen-associated molecular patterns (PAMPs) (29–33). The transition of some herpesviral and retroviral latent infections into productive virus replication can be promoted by some of these same stressors (66, 97–103). Thus, it is possible that one mechanism by which stress induces virus production is by alleviating miRNA-mediated repression of viral and/or host transcripts. In such a scenario, any miRNA target that is capable of initiating a prolytic feed-forward loop (for example, RTA in KSHV) could contribute to context-specific virus induction. Although it awaits validation, such a model implies that viruses may use miRNA-mediated regulation as a barometer for gauging cellular stress (Figure 1).

HOST MicroRNAs From a viral perspective, virus-encoded miRNAs seem to offer several advantages, including the ability to stealthily regulate numerous host and viral transcripts while occupying relatively little genomic space. Therefore, it is striking that many viruses, especially those with positive- and negative-strand RNA genomes, do not encode miRNAs. Even some DNA viruses with nuclear life cycles, including some of the papillomaviruses (PVs), do not encode miRNAs (104, 105). It is likely that the life cycles of these viruses do not require the fine-tuning regulation provided by

Access provided by 45.16.157.220 on 10/02/15. For personal use only. miRNAs, or that they obtain equivalent regulation via alternative strategies. One such strategy used by diverse viruses is to take advantage of host miRNAs. As discussed below, host miRNA– mediated regulation of virus gene expression can lead to consequential biological effects. Annual Review of Virology 2014.1:89-109. Downloaded from www.annualreviews.org The PVs are small viruses with circular, double-stranded DNA; they are the causative agents of benign warts and several human cancers (106). Two reports (107, 108) claim that several alphapapillomavirus types (6, 16, 18, 38, 45, 68) encode miRNAs. However, these reports have not been independently verified and suffer from a lower level of experimental verification than is currently the norm for the field. Thus, it is not widely accepted that PVs encode miRNAs. What is clear is that at least one PV, human papillomavirus 31 (HPV-31), does not encode its own miRNAs (104, 105). PVs can cause long-term persistent infections, whereby the genome is maintained as an episome in undifferentiated keratinocytes (109). Only upon cellular differentiation is the full repertoire of viral proteins expressed (109). Gunasekharan & Laimins (105) found in a raft culture model of keratinocyte differentiation that 93 host miRNAs were differentially expressed in the presence of the HPV-31 genome. One of these miRNAs, miR-145, directly targeted the

www.annualreviews.org • Viral MicroRNAs 97 VI01CH06-Sullivan ARI 2 September 2014 6:40

ab Latency: low stress Lytic: high stress

Stress

RISC RISC

RTA ICP0 RTA ICP0 ZTA Host targets ZTA Host targets T antigens UL28 T antigens UL28 UL32 UL32 IE1 hhi1 IE1 hhi1

RN miRNAmmiR vit actiactivityccttivittiv PProlyticrolytic mmiRNAiRNA transcriptstranscripts aactivityctivity

lytic PProlyticro pts scri transcriptstran

Figure 1 Hypothetical model: MicroRNAs function as a “barometer” of stress. Note that the figure represents a schematic version of infection rather than the life cycle of any specific virus. (a) During persistence or latency, miRNAs restrict expression of viral and/or host transcripts that promote lytic infection (e.g., Kaposi sarcoma–associated herpesvirus RTA, Epstein–Barr virus ZTA, polyomaviral T antigens). In this regard, miRNAs serve to balance prolytic transcripts and to maintain and reinforce persistence or latency. (b)Under stress conditions, miRNA suppressive activity is reduced, thus allowing increased expression of prolytic transcripts. Therefore, under stress conditions, alleviation of miRNA suppression can contribute to initiating or increasing lytic infection. Abbreviation: RISC, RNA-induced silencing complex.

polycistronic early viral transcripts that can give rise to multiple different viral proteins (105). This regulation has functional relevance, as forced expression of miR-145 results in reduced viral

Access provided by 45.16.157.220 on 10/02/15. For personal use only. genome amplification upon cell differentiation (105). In this way, HPV seems to manipulate host miR-145 levels to restrict genome amplification in undifferentiated cells, while allowing for full virus gene expression in differentiated cells. Thus, HPVs may use host miRNAs to the same ends Annual Review of Virology 2014.1:89-109. Downloaded from www.annualreviews.org as the PyVs and other viruses use autoregulatory viral miRNAs. In addition to viruses altering host miRNAs to regulate virus gene expression, some viruses have altered tissue distribution due to preexisting differences in host miRNAs. The North Amer- ican eastern equine encephalitis virus (EEEV) is an alphavirus that can be transmitted to humans via mosquito vectors (110). Although the case mortality rate is high, humans are generally con- sidered to be a dead-end host; birds are the major vertebrate hosts of this arbovirus (111). There are direct target sites for the host miRNA miR-142-3p in the 3 UTR of EEEV (112). Through the use of overexpression and inhibition strategies, it was demonstrated that miR-142 restricts EEEV replication. Interestingly, miR-142 expression is abundant only in hematopoietic cells, suggesting that this mechanism accounts for the lack of replication of EEEV in certain cell types such as macrophages and dendritic cells (112). In a mouse model, a mutant virus that is resistant to miR-142 regulation restores replication in macrophages (112). This gives rise to increased

98 Cox · Sullivan VI01CH06-Sullivan ARI 2 September 2014 6:40

interferon-α/β serum levels, and the virus is consequently less virulent (112). Even though mosquitoes do not express miR-142, the genomic region surrounding the docking sites for miR- 142 is important for replication in the mosquito. The authors interpret this as evidence that this region of the UTR has redundant functions in the arbovector, and they suggest that this phe- nomenon may contribute to the positive selective pressure required to maintain these miRNA target sites (112). An alternative, non–mutually exclusive model is that miR-142-mediated regula- tion in the avian hosts could be advantageous to the virus, perhaps by allowing evasion of the innate immune response while somehow still permitting high enough levels of viremia. Thus, viruses may use miRNAs not only to coordinate with the differentiation and stress status of infected cells but also to limit replication in particular tissues. In the past 10 years, there has been an ongoing debate involving miRNA regulation of HIV. In the midst of the controversy are contradicting reports that HIV encodes several miRNAs (60, 93, 113–116). Early reports based on low-throughput small-RNA sequencing and computational prediction technologies have not been recapitulated by other labs (60, 113). Thus, until it is independently verified that HIV encodes bioactive miRNAs with relevant functionality, caution is warranted toward any claims that HIV makes miRNAs. Additionally, HIV has been suggested to use host miRNAs to regulate its transcripts (117, 118). However, Whisnant et al. (113) recently demonstrated by using CLIP technologies that the HIV genome is largely resistant to host miRNA regulation. This study verified that a small subset of the published miRNAs do in fact directly contact a small fraction of HIV transcripts, but questions were raised as to whether this contact is sufficient for meaningful bioactivity (113). It is worth noting that a lncRNA, NEAT1, has also been shown to regulate HIV gene expression (119). Further studies are warranted to determine the effects of host ncRNAs on the HIV life cycle.

A SUGGESTED ROLE FOR HOST MicroRNAs AS A DIRECT ANTIVIRAL DEFENSE In principle, host miRNA–mediated negative regulation of viruses falls into one of three categories: (a) a natural host defense against specific viral pathogens, (b) a gene regulation tool employed by the virus, or (c) a consequence of the laboratory system used that does not occur during natural infection (120). Several papers have published models whereby miRNAs act as a direct defense against specific pathogens (121–123). However, we note that others have questioned this interpre- tation because in some instances these host miRNAs are not expressed in a physiologically relevant

Access provided by 45.16.157.220 on 10/02/15. For personal use only. context (120, 124, 125). This criticism is supported by experimental evidence demonstrating that some miRNAs are not at a sufficient copy number for bioactivity, or are not expressed in the appro- priate host or cell type during natural infection (120, 124, 125). Any model whereby host miRNAs Annual Review of Virology 2014.1:89-109. Downloaded from www.annualreviews.org serve to directly combat a specific virus will have to account for the inability of these viruses to evolve resistance. This is especially problematic given that only a single nucleotide change can be sufficient to abolish miRNA-mediated suppression (126). As discussed above, miRNA-mediated negative regulation can be advantageous to the virus because diverse viruses encode their own autoregulatory miRNAs. Indeed, some miRNA-mutant viruses replicate better in cell culture and/or in vivo models (25, 27, 50, 90, 91). Yet, these viral miRNAs are maintained in the wild, and their functionalities are typically conserved among divergent viruses (10, 53, 70, 75, 127). This underscores the reality that standard laboratory assays sometimes do not recapitulate the proviral advantages conveyed by miRNAs during natural infection. As new data emerge, it may be worth revisiting some of the early reports claiming direct miRNA antiviral effectors. In a notable study, Lecellier et al. (128) proposed that host miRNAs can be a direct antiviral defense. They demonstrated that miR-32 can repress primate foamy virus 1 (PFV-1;

www.annualreviews.org • Viral MicroRNAs 99 VI01CH06-Sullivan ARI 2 September 2014 6:40

note that PFV-1 now refers to prototype foamy virus 1) accumulation through a docking site located within the coding region of ORF2 that is also the 3 UTR of all the remaining PFV-1 mRNAs (128). Although this was argued as a miRNA-based antiviral defense, it remains to be Aptamer: typically an oligonucleotide (can determined whether this interaction occurs to a meaningful degree in vivo and, if so, whether it is be a protein) that disadvantageous to the fitness of the virus. Importantly, the recent discovery that PFV and other binds to a specific SFVs encode highly expressed, evolutionarily conserved miRNAs (R.P. Kincaid, Y. Chen, J.E. target molecule Cox, A. Rethwilm & C.S. Sullivan, unpublished observations) demonstrates that some miRNA regulation is clearly advantageous to overall PFV fitness.

OTHER CLASSES OF VIRAL NONCODING RNAs Decades before the discovery of miRNAs, virus-encoded ncRNAs were expected to play an im- portant role in the infectious cycle (14). Several different viruses, including herpesviruses and adenoviruses, encode robust amounts of longer ncRNAs (∼160 to >1,000 nt). Unlike miRNAs, these ncRNAs likely represent a hodgepodge of different mechanisms of action, and as such, a full understanding of how they function is still lacking. Two of these ncRNAs for which recent insights have been reported include the adenovirus virus-associated (VA) RNAs and the KSHV polyadenylated nuclear RNA. VA RNAs are encoded by the lung- and gastrointestinal tract–tropic adenoviruses. These are among the best studied and oldest known viral ncRNAs. These ∼165-nt RNAs are transcribed by RNA polymerase III during the lytic cycle to 107–108 copies/cell, ranking among the most numerically abundant RNAs (14, 129). Both primate and human adenoviruses can express up to two VA RNAs (130). VA I RNA functions like a natural aptamer that stoichiometrically binds to and inhibits (PKR) (131). Thus, this represents an important counterdefense used by the virus to evade the antiviral response. Additionally, VA I inhibition of PKR indirectly promotes viral gene expression by limiting translational repression associated with PKR activation. Interestingly, less than 1% of the VA RNAs is processed into miRNAs. Despite the inefficient processing, VA-derived miRNAs are the most abundant miRNAs (∼106 copies/cell) detected in a cell undergoing lytic infection (129). However, these miRNAs are likely not important during lytic infection. Unlike VA-null viruses, which are attenuated for virus production by 1–2 orders of magnitude, engineered mutant viruses that retain the larger VA structure but produce miRNAs with altered seeds behave like wild-type viruses (129). Although much less studied, adenoviruses can also produce a persistent infection (132, 133). Recently, it was shown in a cell culture model ∼

Access provided by 45.16.157.220 on 10/02/15. For personal use only. of persistence that the VA-derived miRNAs are readily detectable and account for 2.7% of the total RISC-associated RNAs (132, 134). Thus, it remains formally possible that VA miRNAs are important during persistent infection or in other unknown contexts, but this possibility has not Annual Review of Virology 2014.1:89-109. Downloaded from www.annualreviews.org yet been demonstrated (Figure 2). KSHV encodes a 1.1-kb ncRNA called the polyadenylated nuclear RNA (PAN). PAN is ex- pressed during lytic infection and is the most abundant viral transcript produced, topping out at 106–107 copies/cell (135). Although polyadenylated, PAN is nuclear and has an atypical 3 end structure that renders it resistant to nucleolytic turnover (136, 137). Until recently, no functions for PAN had been reported. However, in the past few years, several exciting reports have been published. Knockdown and knockout studies show that PAN is important for virus gene expression (138, 139). Importantly, PAN is required for expression of RTA, the master lytic switch protein (140). The published mechanism that can account for this is that PAN is a lncRNA that alters viral chromatin (139). Very recently, a small fraction of PAN has been shown to be associated with translationally active ribosomes (141). Though only a small fraction, this still would potentially account for some of the most abundant virally derived peptides made during infection (141). The

100 Cox · Sullivan VI01CH06-Sullivan ARI 2 September 2014 6:40

abLytic infection Persistent infection

DampensDampens RISC antiviral response ? Viral targets ProteinProtein IndirectlyIndirectly promotepromotess Host targets kkinaseinase R viral genegene expression

Figure 2 Model: role of virus-associated (VA) I RNA in adenovirus infection. (a) VA I RNAs are structured, highly abundant RNAs encoded by some primate viruses (including human adenoviruses). During lytic infection, VA I RNAs stoichiometrically bind to protein kinase R (PKR), preventing its activation. This limits the cellular antiviral response and prevents translation inhibition, thus allowing viral gene expression (128). (b) VA I RNA, and the related VA II RNA in some adenoviruses, can be inefficiently processed into miRNAs. These miRNAs are made during lytic infection but have also been detected in cell models of persistent infection (130). It is therefore possible that the VA RNA–derived miRNAs function to maintain persistent infection, but such a model remains untested. Abbreviation: RISC, RNA-induced silencing complex.

function(s) of these putative peptides and how translating polyribosomes are associated with the (mostly?) nuclear-restricted PAN remain a mystery.

WHAT ARE THEY GOOD FOR? With our current understanding of the impact of miRNAs on viral gene expression, miRNAs represent a plausible avenue of antiviral therapeutics. Notably, human clinical trials are already underway for therapeutics that target miR-122 to limit hepatitis C infection. Hepatitis C uses miR-122 to positively regulate virus copy number (142, 143). The success of this work suggests that targeting virally encoded miRNAs could also be a viable treatment strategy when the delivery barrier of miRNA therapeutics can be overcome. Others have engineered multiple host miRNA docking sites to restrict infection by gene therapy vectors (144–146). This approach shows promise in limiting oncolytic viruses to tumor tissues. A similar approach has been developed to restrict species specificity to laboratory model animals of highly pathogenic avian influenza (147). Finally,

Access provided by 45.16.157.220 on 10/02/15. For personal use only. in cases where viral genome copy number is limiting but miRNA copy number is high, viral miRNAs may be useful as biomarkers of disease. For example, this method has been proposed for

Annual Review of Virology 2014.1:89-109. Downloaded from www.annualreviews.org early detection of a herpesvirus associated with a fatal disease in elephants (148, 149).

SUMMARY POINTS 1. Diverse viruses, including DNA viruses and retroviruses, encode miRNAs. 2. Many viral miRNAs target host and/or viral transcripts as a mechanism to regulate virus gene expression. 3. miRNA regulation plays an important role in regulating persistent infection. 4. Host miRNAs are unlikely to be a direct defense against specific viral transcripts. 5. Viruses may use miRNAs as barometers to measure stress levels in host cells.

www.annualreviews.org • Viral MicroRNAs 101 VI01CH06-Sullivan ARI 2 September 2014 6:40

6. Longer ncRNAs are emerging as important players in virus infection. 7. Manipulating miRNA regulation shows promise in optimizing viral gene therapy vectors.

FUTURE ISSUES 1. New studies of animal models are needed to establish the role of ncRNAs in viral per- sistence in vivo. 2. Determining the full repertoire of virus families that encode miRNAs will help us un- derstand known viral miRNA functions. 3. Uncovering alternative mechanisms of regulation for persistent viruses that do not encode miRNAs is clinically important. 4. Entire new classes of viral and host ncRNAs relevant to infection likely await discovery. 5. Viral ncRNAs offer new therapeutics and biomarkers strategies.

DISCLOSURE STATEMENT The authors are not aware of any affiliations, memberships, funding, or financial holdings that might be perceived as affecting the objectivity of this review.

ACKNOWLEDGMENTS This work was supported by National Institutes of Health grant R01AI077746, a Burroughs Wellcome Investigators in Pathogenesis Award, Cancer Prevention and Research Institute of Texas grant RP110098, a National Science Foundation CAREER award, and a University of Texas at Austin Institute for Cellular and fellowship.

LITERATURE CITED 1. Lee RC, Feinbaum RL, Ambros V. 1993. The C. elegans heterochronic gene lin-4 encodes small RNAs Access provided by 45.16.157.220 on 10/02/15. For personal use only. with antisense complementarity to lin-14. Cell 75:843–54 2. Ulitsky I, Bartel DP. 2013. lincRNAs: genomics, evolution, and mechanisms. Cell 154:26–46 Annual Review of Virology 2014.1:89-109. Downloaded from www.annualreviews.org 3. Wang KC, Chang HY. 2011. Molecular mechanisms of long noncoding RNAs. Mol. Cell 43:904–14 4. Baltimore D. 1971. Expression of animal virus genomes. Bacteriol. Rev. 35:235–41 5. Funk A, Truong K, Nagasaki T, Torres S, Floden N, et al. 2010. RNA structures required for production of subgenomic flavivirus RNA. J. Virol. 84:11407–17 6. Grundhoff A, Sullivan CS. 2011. Virus-encoded microRNAs. Virology 411:325–43 7. Kincaid RP, Burke JM, Cox JC, de Villiers EM, Sullivan CS. 2013. A human torque teno virus encodes a microRNA that inhibits interferon signaling. PLoS Pathog. 9:e1003818 8. Kincaid RP, Burke JM, Sullivan CS. 2012. RNA virus microRNA that mimics a B-cell oncomiR. Proc. Natl. Acad. Sci. USA 109:3077–82 9. Yao Y, Smith LP, Nair V, Watson M. 2014. An avian retrovirus uses canonical expression and processing mechanisms to generate viral microRNA. J. Virol. 88:2–9 10. Kincaid RP, Sullivan CS. 2012. Virus-encoded microRNAs: an overview and a look to the future. PLoS Pathog. 8:e1003018

102 Cox · Sullivan VI01CH06-Sullivan ARI 2 September 2014 6:40

11. Sullivan CS. 2008. New roles for large and small viral RNAs in evading host defences. Nat. Rev. Genet. 9:503–7 12. Umbach JL, Cullen BR. 2009. The role of RNAi and microRNAs in animal virus replication and antiviral immunity. Genes Dev. 23:1151–64 13. Boss IW, Renne R. 2010. Viral miRNAs: tools for immune evasion. Curr. Opin. Microbiol. 13:540–45 14. Sullivan CS, Cullen BR. 2009. Non-coding regulatory RNAs of the DNA tumor viruses. In DNA Tumor Viruses, ed. B Damania, JM Pipas, pp. 645–82. New York: Springer 15. Sedger LM. 2013. MicroRNA control of interferons and interferon induced anti-viral activity. Mol. Immunol. 56:781–93 16. Subramanian S, Steer CJ. 2010. MicroRNAs as gatekeepers of apoptosis. J. Cell. Physiol. 223:289–98 17. Chen CZ, Schaffert S, Fragoso R, Loh C. 2013. Regulation of immune responses and tolerance: the microRNA perspective. Immunol. Rev. 253:112–28 18. Farazi TA, Hoell JI, Morozov P, Tuschl T. 2013. MicroRNAs in human cancer. Adv. Exp. Med. Biol. 774:1–20 19. Cullen BR. 2004. Transcription and processing of human microRNA precursors. Mol. Cell 16:861–65 20. Flynt AS, Lai EC. 2008. Biological principles of microRNA-mediated regulation: shared themes amid diversity. Nat. Rev. Genet. 9:831–42 21. Zisoulis DG, Lovci MT, Wilbert ML, Hutt KR, Liang TY, et al. 2010. Comprehensive discovery of endogenous Argonaute binding sites in Caenorhabditis elegans. Nat. Struct. Mol. Biol. 17:173–79 22. Lim LP, Lau NC, Garrett-Engele P, Grimson A, Schelter JM, et al. 2005. Microarray analysis shows that some microRNAs downregulate large numbers of target mRNAs. Nature 433:769–73 23. Seo GJ, Chen CJ, Sullivan CS. 2009. Merkel cell polyomavirus encodes a microRNA with the ability to autoregulate viral gene expression. Virology 383:183–87 24. Seo GJ, Fink LHL, O’Hara B, Atwood WJ, Sullivan CS. 2008. Evolutionarily conserved function of a viral microRNA. J. Virol. 82:9823–28 25. Sullivan CS, Grundhoff AT, Tevethia S, Pipas JM, Ganem D. 2005. SV40-encoded microRNAs 25. First demonstration regulate viral gene expression and reduce susceptibility to cytotoxic T cells. Nature 435:682–86 of a viral miRNA 26. Barth S, Pfuhl T, Mamiani A, Ehses C, Roemer K, et al. 2008. Epstein–Barr virus–encoded microRNA target—the miR-BART2 down-regulates the viral DNA polymerase BALF5. Nucleic Acids Res. 36:666–75 autoregulation of T 27. Sullivan CS, Sung CK, Pack CD, Grundhoff A, Lukacher AE, et al. 2009. Murine polyomavirus encodes antigen transcripts. a microRNA that cleaves early RNA transcripts but is not essential for experimental infection. Virology 387:157–67 28. Bhattacharyya SN, Habermacher R, Martine U, Closs EI, Filipowicz W. 2006. Stress-induced reversal of microRNA repression and mRNA P-body localization in human cells. Cold Spring Harb. Symp. Quant. Biol. 71:513–21 29. Leung AKL, Vyas S, Rood JE, Bhutkar A, Sharp PA, Chang P. 2011. Poly(ADP-ribose) regulates stress responses and microRNA activity in the cytoplasm. Mol. Cell 42:489–99 Access provided by 45.16.157.220 on 10/02/15. For personal use only. 30. Seo GJ, Kincaid RP, Phanaksri T, Burke JM, Pare JM, et al. 2013. Reciprocal inhibition between intracellular antiviral signaling and the RNAi machinery in mammalian cells. Cell Host Microbe 14:435–

Annual Review of Virology 2014.1:89-109. Downloaded from www.annualreviews.org 45 31. Bhattacharyya SN, Habermacher R, Martine U, Closs EI, Filipowicz W. 2006. Relief of microRNA- mediated translational repression in human cells subjected to stress. Cell 125:1111–24 32. Vasudevan S, Tong Y, Steitz JA. 2007. Switching from repression to activation: microRNAs can up- regulate translation. Science 318:1931–34 33. Vasudevan S, Steitz JA. 2007. AU-rich-element-mediated upregulation of translation by FXR1 and Argonaute 2. Cell 128:1105–18 34. Ebert MS, Sharp PA. 2012. Roles for microRNAs in conferring robustness to biological processes. Cell 149:515–24 35. Cassidy JJ, Jha AR, Posadas DM, Giri R, Venken KJT, et al. 2013. miR-9a minimizes the phenotypic impact of genomic diversity by buffering a transcription factor. Cell 155:1556–67 36. Mendell JT, Olson EN. 2012. MicroRNAs in stress signaling and human disease. Cell 148:1172–87 37. Pelaez´ N, Carthew RW. 2012. Biological robustness and the role of microRNAs: a network perspective. Curr. Top. Dev. Biol. 99:237–55

www.annualreviews.org • Viral MicroRNAs 103 VI01CH06-Sullivan ARI 2 September 2014 6:40

38. Kozomara A, Griffiths-Jones S. 2011. miRBase: integrating microRNA annotation and deep-sequencing data. Nucleic Acids Res. 39:D152–57 39. Demonstrated that 39. Grey F, Meyers H, White EA, Spector DH, Nelson J. 2007. A human cytomegalovirus-encoded viral miRNAs can microRNA regulates expression of multiple viral genes involved in replication. PLoS Pathog. regulate several 3:e163 different viral 40. Murphy E, Vanıcek´ J, Robins H, Shenk T, Levine AJ. 2008. Suppression of immediate-early viral transcripts. gene expression by herpesvirus-coded microRNAs: implications for latency. Proc. Natl. Acad. Sci. USA 105:5453–58 40. First to suggest that 41. Forte E, Luftig MA. 2011. The role of microRNAs in Epstein–Barr virus latency and lytic reactivation. miRNAs may regulate Microbes Infect. 13:1156–67 numerous viral 42. Lieberman PM. 2013. Keeping it quiet: chromatin control of gammaherpesvirus latency. Nat. Rev. transcripts in different Microbiol. 11:863–75 herpesviruses. 43. Wilson AC, Mohr I. 2012. A cultured affair: HSV latency and reactivation in neurons. Trends Microbiol. 20:604–11 44. Umbach JL, Kramer MF, Jurak I, Karnowski HW, Coen DM, Cullen BR. 2008. MicroRNAs ex- 44. Demonstrated that pressed by herpes simplex virus 1 during latent infection regulate viral mRNAs. Nature 454:780– HSV-1 encodes 83 autoregulatory 45. Jurak I, Griffiths A, Coen DM. 2011. Mammalian alphaherpesvirus miRNAs. Biochim. Biophys. Acta miRNAs; implied a role in maintaining latency. 1809:641–53 46. Boutell C, Everett RD. 2013. Regulation of alphaherpesvirus infections by the ICP0 family of proteins. J. Gen. Virol. 94:465–81 47. Everett RD. 2000. ICP0, a regulator of herpes simplex virus during lytic and latent infection. BioEssays 22:761–70 48. Everett RD, Parsy ML, Orr A. 2009. Analysis of the functions of herpes simplex virus type 1 regulatory protein ICP0 that are critical for lytic infection and derepression of quiescent viral genomes. J. Virol. 83:4963–77 49. Smith MC, Boutell C, Davido DJ. 2011. HSV-1 ICP0: paving the way for viral replication. Future Virol. 6:421–29 50. An HSV-1 miRNA 50. Flores O, Nakayama S, Whisnant AW, Javanbakht H, Cullen BR, Bloom DC. 2013. Mutational deletion mutant shows inactivation of herpes simplex virus 1 microRNAs identifies viral mRNA targets and reveals increased replication phenotypic effects in culture. J. Virol. 87:6589–603 under only some 51. Dolken¨ L, Pfeffer S, Koszinowski UH. 2009. Cytomegalovirus microRNAs. Virus Genes 38:355–64 experimental 52. Huang Y, Qi Y, Ma Y, He R, Ji Y, et al. 2013. Down-regulation of human cytomegalovirus UL138, a conditions. novel latency-associated determinant, by hcmv-miR-UL36. J. Biosci. 38:479–85 53. Hancock MH, Tirabassi RS, Nelson JA. 2012. Rhesus cytomegalovirus encodes seventeen microRNAs that are differentially expressed in vitro and in vivo. Virology 425:133–42 54. Stern-Ginossar N, Saleh N, Goldberg MD, Prichard M, Wolf DG, Mandelboim O. 2009. Analysis of Access provided by 45.16.157.220 on 10/02/15. For personal use only. human cytomegalovirus–encoded microRNA activity during infection. J. Virol. 83:10684–93 55. Fu M, Gao Y, Zhou Q, Zhang Q, Peng Y, et al. 2014. Human cytomegalovirus latent infection alters

Annual Review of Virology 2014.1:89-109. Downloaded from www.annualreviews.org the expression of cellular and viral microRNA. Gene 536:272–78 56. Dolken¨ L, Krmpotic A, Kothe S, Tuddenham L, Tanguy M, et al. 2010. Cytomegalovirus microRNAs facilitate persistent virus infection in salivary glands. PLoS Pathog. 6:e1001150 57. Gottwein E, Corcoran DL, Mukherjee N, Skalsky RL, Hafner M, et al. 2011. Viral microRNA targetome of KSHV-infected primary effusion lymphoma cell lines. Cell Host Microbe 10:515–26 58. Grundhoff A, Sullivan CS, Ganem D. 2006. A combined computational and microarray-based approach identifies novel microRNAs encoded by human gamma-herpesviruses. RNA 12:733–50 59. Haecker I, Gay LA, Yang Y, Hu J, Morse AM, et al. 2012. Ago HITS-CLIP expands understanding of Kaposi’s sarcoma–associated herpesvirus miRNA function in primary effusion lymphomas. PLoS Pathog. 8:e1002884 60. Pfeffer S, Sewer A, Lagos-Quintana M, Sheridan R, Sander C, et al. 2005. Identification of microRNAs of the herpesvirus family. Nat. Methods 2:269–76 61. Zhu Y, Haecker I, Yang Y, Gao SJ, Renne R. 2013. γ-Herpesvirus-encoded miRNAs and their roles in viral biology and pathogenesis. Curr. Opin. Virol. 3:266–75

104 Cox · Sullivan VI01CH06-Sullivan ARI 2 September 2014 6:40

62. Lukac DM, Kirshner JR, Ganem D. 1999. Transcriptional activation by the product of open reading frame 50 of Kaposi’s sarcoma–associated herpesvirus is required for lytic viral reactivation in B cells. J. Virol. 73:9348–61 63. Lin X, Liang D, He Z, Deng Q, Robertson ES, Lan K. 2011. miR-K12-7-5p encoded by Kaposi’s sarcoma–associated herpesvirus stabilizes the latent state by targeting viral ORF50/RTA. PLoS ONE 6:e16224 64. Lu F, Stedman W, Yousef M, Renne R, Lieberman PM. 2010. Epigenetic regulation of Kaposi’s 64. Demonstrated that sarcoma–associated herpesvirus latency by virus-encoded microRNAs that target Rta and the deletion of 10 of 12 cellular Rbl2-DNMT pathway. J. Virol. 84:2697–706 KSHV pre-miRNAs 65. Lei X, Bai Z, Ye F, Xie J, Kim CG, et al. 2010. Regulation of NF-κB inhibitor IκBα and viral results in increased lytic infection. replication by a KSHV microRNA. Nat. Cell Biol. 12:193–99 66. Liang D, Gao Y, Lin X, He Z, Zhao Q, et al. 2011. A human herpesvirus miRNA attenuates interferon signaling and contributes to maintenance of viral latency by targeting IKKε. Cell Res. 21:793–806 65. Demonstrated that a 67. Lu CC, Li Z, Chu CY, Feng J, Feng J, et al. 2010. MicroRNAs encoded by Kaposi’s sarcoma–associated KSHV miRNA can control viral replication herpesvirus regulate viral life cycle. EMBO Rep. 11:784–90 through the NFκB 68. Moody R, Zhu Y, Huang Y, Cui X, Jones T, et al. 2013. KSHV microRNAs mediate cellular trans- pathway. formation and tumorigenesis by redundantly targeting cell growth and survival pathways. PLoS Pathog. 9:e1003857 69. Ziegelbauer JM, Sullivan CS, Ganem D. 2009. Tandem array–based expression screens identify host mRNA targets of virus-encoded microRNAs. Nat. Genet. 41:130–34 70. Umbach JL, Strelow LI, Wong SW, Cullen BR. 2010. Analysis of rhesus rhadinovirus microRNAs expressed in virus-induced tumors from infected rhesus macaques. Virology 405:592–99 71. Bruce AG, Ryan JT, Thomas MJ, Peng X, Grundhoff A, et al. 2013. Next-generation sequence analysis of the genome of RFHVMn, the macaque homolog of Kaposi’s sarcoma (KS)-associated herpesvirus, from a KS-like tumor of a pig-tailed macaque. J. Virol. 87:13676–93 72. Bai Z, Huang Y, Li W, Zhu Y, Jung JU, et al. 2014. Genomewide mapping and screening of Kaposi’s  sarcoma–associated herpesvirus (KSHV) 3 untranslated regions identify bicistronic and polycistronic viral transcripts as frequent targets of KSHV microRNAs. J. Virol. 88:377–92 73. McClure LV, Kincaid RP, Burke JM, Grundhoff A, Sullivan CS. 2013. Comprehensive mapping and analysis of Kaposi’s sarcoma–associated herpesvirus 3 UTRs identify differential posttranscriptional control of gene expression in lytic versus latent infection. J. Virol. 87:12838–49 74. Lin YT, Sullivan CS. 2011. Expanding the role of Drosha to the regulation of viral gene expression. Proc. Natl. Acad. Sci. USA 108:11229–34 75. Cai X, Schafer¨ A, Lu S, Bilello JP, Desrosiers RC, et al. 2006. Epstein–Barr virus microRNAs are evolutionarily conserved and differentially expressed. PLoS Pathog. 2:e23 76. Pfeffer S, Zavolan M, Grasser¨ FA, Chien M, Russo JJ, et al. 2004. Identification of virus-encoded 76. First publication of

Access provided by 45.16.157.220 on 10/02/15. For personal use only. microRNAs. Science 304:734–36 viral miRNAs; 77. Riley KJ, Rabinowitz GS, Yario TA, Luna JM, Darnell RB, Steitz JA. 2012. EBV and human microRNAs suggested that co-target oncogenic and apoptotic viral and human genes during latency. EMBO J. 31:2207–21 miR-BART2 directly Annual Review of Virology 2014.1:89-109. Downloaded from www.annualreviews.org 78. Iizasa H, Wulff BE, Alla NR, Maragkakis M, Megraw M, et al. 2010. Editing of Epstein–Barr virus– regulates the BALF5 encoded BART6 microRNAs controls their Dicer targeting and consequently affects viral latency. polymerase transcript. J. Biol. Chem. 285:33358–70 79. Seto E, Moosmann A, Gromminger¨ S, Walz N, Grundhoff A, Hammerschmidt W. 2010. Micro RNAs of Epstein–Barr virus promote cell cycle progression and prevent apoptosis of primary human B cells. PLoS Pathog. 6:e1001063 80. DeCaprio JA, Garcea RL. 2013. A cornucopia of human polyomaviruses. Nat. Rev. Microbiol. 11:264–76 81. Korup S, Rietscher J, Calvignac-Spencer S, Trusch F, Hofmann J, et al. 2013. Identification of a novel human polyomavirus in organs of the gastrointestinal tract. PLoS ONE 8:e58021 82. Delbue S, Ferraresso M, Ghio L, Carloni C, Carluccio S, et al. 2013. A review on JC virus infection in kidney transplant recipients. Clin. Dev. Immunol. 2013:926391 83. Hirsch HH, Kardas P, Kranz D, Leboeuf C. 2013. The human JC polyomavirus ( JCPyV): virological background and clinical implications. APMIS 121:685–727

www.annualreviews.org • Viral MicroRNAs 105 VI01CH06-Sullivan ARI 2 September 2014 6:40

84. Rinaldo CH, Tylden GD, Sharma BN. 2013. The human polyomavirus BK (BKPyV): virological back- ground and clinical implications. APMIS 121:728–45 85. Kazem S, van der Meijden E, Feltkamp MCW. 2013. The trichodysplasia spinulosa–associated poly- omavirus: virological background and clinical implications. APMIS 121:770–82 86. Cole CN. 1996. Polyomaviridae: the viruses and their replication. In Fields Virology,ed.DMKnipe,PM Howley, pp. 1997–2043. Philadelphia: Lippincott-Raven. 3rd ed. 87. Cantalupo P, Doering A, Sullivan CS, Pal A, Peden KWC, et al. 2005. Complete nucleotide sequence of polyomavirus SA12. J. Virol. 79:13094–104 88. Chen CJ, Cox JE, Kincaid RP, Martinez A, Sullivan CS. 2013. Divergent microRNA targetomes of closely related circulating strains of a polyomavirus. J. Virol. 87:11135–47 89. Chen CJ, Kincaid RP, Seo GJ, Bennett MD, Sullivan CS. 2011. Insights into Polyomaviridae microRNA function derived from study of the bandicoot papillomatosis carcinomatosis viruses. J. Virol. 85:4487–500 90. Broekema NM, Imperiale MJ. 2013. miRNA regulation of BK polyomavirus replication during early infection. Proc. Natl. Acad. Sci. USA 110:8200–5 91. Zhang S, Sroller V, Zanwar P, Chen CJ, Halvorson SJ, et al. 2014. Viral microRNA effects on patho- genesis of polyomavirus SV40 infections in Syrian golden hamsters. PLoS Pathogens 10:e1003912 92. Goff SP. 2001. Retroviridae: the retroviruses and their replication. In Fields Virology,ed.DMKnipe, PM Howley, DE Griffin, RA Lamb, MA Martin, B Roizman, pp. 1871–922. Philadelphia: Lippincott Williams & Wilkins. 4th ed. 93. Lin J, Cullen BR. 2007. Analysis of the interaction of primate retroviruses with the human RNA inter- ference machinery. J. Virol. 81:12218–26 94. Schmidt M, Herchenroder¨ O, Heeney J, Rethwilm A. 1997. Long terminal repeat U3 length polymor- phism of human foamy virus. Virology 230:167–78 95. Rethwilm A, Baunach G, Netzer KO, Maurer B, Borisch B, ter Meulen V. 1990. Infectious DNA of the human spumaretrovirus. Nucleic Acids Res. 18:733–38 96. Neves M, Peri´ es` J, Saıb¨ A. 1998. Study of human foamy virus proviral integration in chronically infected murine cells. Res. Virol. 149:393–401 97. Gregory SM, West JA, Dillon PJ, Hilscher C, Dittmer DP, Damania B. 2009. Toll-like receptor signaling controls reactivation of KSHV from latency. Proc. Natl. Acad. Sci. USA 106:11725–30 98. Li X, Feng J, Sun R. 2011. Oxidative stress induces reactivation of Kaposi’s sarcoma–associated her- pesvirus and death of primary effusion lymphoma cells. J. Virol. 85:715–24 99. Lu C, Zeng Y, Huang Z, Huang L, Qian C, et al. 2005. Human herpesvirus 6 activates lytic cycle replication of Kaposi’s sarcoma–associated herpesvirus. Am. J. Pathol. 166:173–83 100. Palmisano I, Della Chiara G, D’Ambrosio RL, Huichalaf C, Brambilla P, et al. 2012. Amino acid starva- tion induces reactivation of silenced transgenes and latent HIV-1 provirus via down-regulation of histone deacetylase 4 (HDAC4). Proc. Natl. Acad. Sci. USA 109:E2284–93

Access provided by 45.16.157.220 on 10/02/15. For personal use only. 101. Qin D, Zeng Y, Qian C, Huang Z, Lv Z, et al. 2008. Induction of lytic cycle replication of Kaposi’s sarcoma–associated herpesvirus by herpes simplex virus type 1: involvement of IL-10 and IL-4. Cell Microbiol. 10:713–28 Annual Review of Virology 2014.1:89-109. Downloaded from www.annualreviews.org 102. Taylor GM, Raghuwanshi SK, Rowe DT, Wadowsky RM, Rosendorff A. 2011. Endoplasmic reticulum stress causes EBV lytic replication. Blood 118:5528–39 103. Vieira J, O’Hearn P, Kimball L, Chandran B, Corey L. 2001. Activation of Kaposi’s sarcoma–associated herpesvirus (human herpesvirus 8) lytic replication by human cytomegalovirus. J. Virol. 75:1378–86 104. Cai X, Li G, Laimins LA, Cullen BR. 2006. Human papillomavirus genotype 31 does not express detectable microRNA levels during latent or productive virus replication. J. Virol. 80:10890–93 105. Gunasekharan V, Laimins LA. 2013. Human papillomaviruses modulate microRNA 145 expression to directly control genome amplification. J. Virol. 87:6037–43 106. Crow JM. 2012. HPV: the global burden. Nature 488:S2–3 107. Gu W, An J, Ye P, Zhao KN, Antonsson A. 2011. Prediction of conserved microRNAs from skin and mucosal human papillomaviruses. Arch. Virol. 156:1161–71 108. Qian K, PietilaT,R¨ onty¨ M, Michon F, Frilander MJ, et al. 2013. Identification and validation of human papillomavirus encoded microRNAs. PLoS ONE 8:e70202

106 Cox · Sullivan VI01CH06-Sullivan ARI 2 September 2014 6:40

109. Doorbar J. 2005. The papillomavirus life cycle. J. Clin. Virol. 32(Suppl. 1):S7–15 110. Zacks MA, Paessler S. 2010. Encephalitic alphaviruses. Vet. Microbiol. 140:281–86 111. Weaver SC. 2005. Host range, amplification and arboviral disease emergence. In Infectious Diseases from Nature: Mechanisms of Viral Emergence and Persistence, ed. CJ Peters, CH Calisher, pp. 33–44. Vienna: Springer 112. Trobaugh DW, Gardner CL, Sun C, Haddow AD, Wang E, et al. 2013. RNA viruses can hijack 112. Demonstrated that vertebrate microRNAs to suppress innate immunity. Nature 506:245–48 host miRNAs can 113. Whisnant AW, Bogerd HP, Flores O, Ho P, Powers JG, et al. 2013. In-depth analysis of the interaction restrict tissue tropism of HIV-1 with cellular microRNA biogenesis and effector mechanisms. mBio 4:e000193 of virus infection in a 114. Klase Z, Kale P, Winograd R, Gupta MV, Heydarian M, et al. 2007. HIV-1 TAR element is processed nonengineered setting. by Dicer to yield a viral micro-RNA involved in chromatin remodeling of the viral LTR. BMC Mol. Biol. 8:63 115. Omoto S, Ito M, Tsutsumi Y, Ichikawa Y, Okuyama H, et al. 2004. HIV-1 nef suppression by virally encoded microRNA. Retrovirology 1:44 116. Schopman NCT, Willemsen M, Liu YP, Bradley T, van Kampen A, et al. 2012. Deep sequencing of virus-infected cells reveals HIV-encoded small RNAs. Nucleic Acids Res. 40:414–27 117. Ahluwalia JK, Khan SZ, Soni K, Rawat P, Gupta A, et al. 2008. Human cellular microRNA hsa-miR-29a interferes with viral nef protein expression and HIV-1 replication. Retrovirology 5:117 118. Huang J, Wang F, Argyris E, Chen K, Liang Z, et al. 2007. Cellular microRNAs contribute to HIV-1 latency in resting primary CD4+ T lymphocytes. Nat. Med. 13:1241–47 119. Zhang Q, Chen CY, Yedavalli VSRK, Jeang KT. 2013. NEAT1 long noncoding RNA and paraspeckle bodies modulate HIV-1 posttranscriptional expression. mBio 4:e00596–12 120. Skalsky RL, Cullen BR. 2010. Viruses, microRNAs, and host interactions. Annu. Rev. Microbiol. 64:123– 41 121. Huang T, Zhang X. 2012. Functional analysis of a crustacean microRNA in host-virus interactions. J. Virol. 86:12997–3004 122. Kelly EJ, Nace R, Barber GN, Russell SJ. 2010. Attenuation of vesicular stomatitis virus encephalitis through microRNA targeting. J. Virol. 84:1550–62 123. Pedersen IM, Cheng G, Wieland S, Volinia S, Croce CM, et al. 2007. Interferon modulation of cellular microRNAs as an antiviral mechanism. Nature 449:919–22 124. Cullen BR. 2013. How do viruses avoid inhibition by endogenous cellular microRNAs? PLoS Pathog. 9:e1003694 125. Sarasin-Filipowicz M, Krol J, Markiewicz I, Heim MH, Filipowicz W. 2009. Decreased levels of mi- croRNA miR-122 in individuals with hepatitis C responding poorly to interferon therapy. Nat. Med. 15:31–33 126. Berezikov E. 2011. Evolution of microRNA diversity and regulation in animals. Nat. Rev. Genet. 12:846– 60 Access provided by 45.16.157.220 on 10/02/15. For personal use only. 127. Skalsky RL, Kang D, Linnstaedt SD, Cullen BR. 2014. Evolutionary conservation of primate lym- phocryptovirus microRNA targets. J. Virol. 88:1617–35

Annual Review of Virology 2014.1:89-109. Downloaded from www.annualreviews.org 128. Lecellier C-H, Dunoyer P, Arar K, Lehmann-Che J, Eyquem S, et al. 2005. A cellular microRNA mediates antiviral defense in human cells. Science 308:557–60 129. Kamel W, Segerman B, Oberg¨ D, Punga T, Akusjarvi¨ G. 2013. The adenovirus VA RNA–derived miRNAs are not essential for lytic virus growth in tissue culture cells. Nucleic Acids Res. 41:4802–12 130. Ma Y, Mathews MB. 1996. Structure, function, and evolution of adenovirus-associated RNA: a phylo- genetic approach. J. Virol. 70:5083–99 131. Maran A, Mathews MB. 1988. Characterization of the double-stranded RNA implicated in the inhibition of protein synthesis in cells infected with a mutant adenovirus defective for VA RNA. Virology 164:106–13 132. Furuse Y, Ornelles DA, Cullen BR. 2013. Persistently adenovirus-infected lymphoid cells express micro- RNAs derived from the viral VAI and especially VAII RNA. Virology 447:140–45 133. Garnett CT, Erdman D, Xu W, Gooding LR. 2002. Prevalence and quantitation of species C adenovirus DNA in human mucosal lymphocytes. J. Virol. 76:10608–16 134. Aparicio O, Razquin N, Zaratiegui M, Narvaiza I, Fortes P. 2006. Adenovirus virus-associated RNA is processed to functional interfering RNAs involved in virus production. J. Virol. 80:1376–84

www.annualreviews.org • Viral MicroRNAs 107 VI01CH06-Sullivan ARI 2 September 2014 6:40

135. Sun R, Lin SF, Gradoville L, Miller G. 1996. Polyadenylylated nuclear RNA encoded by Kaposi sarcoma– associated herpesvirus. Proc. Natl. Acad. Sci. USA 93:11883–88 136. Mitton-Fry RM, DeGregorio SJ, Wang J, Steitz TA, Steitz JA. 2010. Poly(A) tail recognition by a viral RNA element through assembly of a triple helix. Science 330:1244–47 137. Conrad NK, Mili S, Marshall EL, Shu MD, Steitz JA. 2006. Identification of a rapid mammalian deadenylation-dependent decay pathway and its inhibition by a viral RNA element. Mol. Cell 24:943–53 138. Borah S, Darricarrere` N, Darnell A, Myoung J, Steitz JA. 2011. A viral nuclear noncoding RNA binds re- localized poly(A) binding protein and is required for late KSHV gene expression. PLoS Pathog. 7:e1002300 139. Rossetto CC, Tarrant-Elorza M, Verma S, Purushothaman P, Pari GS. 2013. Regulation of viral and cellular gene expression by Kaposi’s sarcoma–associated herpesvirus polyadenylated nuclear RNA. J. Virol. 87:5540–53 140. Rossetto CC, Pari G. 2012. KSHV PAN RNA associates with demethylases UTX and JMJD3 to activate lytic replication through a physical interaction with the virus genome. PLoS Pathog. 8:e1002680 141. Arias C, Weisburd B, Stern-Ginossar N, Mercier A, Madrid AS, et al. 2014. KSHV 2.0: a comprehensive annotation of the Kaposi’s sarcoma–associated herpesvirus genome using next-generation sequencing reveals novel genomic and functional features. PLoS Pathog. 10:e1003847 142. Haussecker D, Kay MA. 2010. miR-122 continues to blaze the trail for microRNA therapeutics. Mol. Ther. 18:240–42 143. Jopling CL, Yi M, Lancaster AM, Lemon SM, Sarnow P. 2005. Modulation of hepatitis C virus RNA abundance by a liver-specific microRNA. Science 309:1577–81 144. tenOever B. 2009. MicroManipulating viral-based therapeutics. Discov. Med. 8:51–54 145. Ibrisimoviˇ c´ M, Kneidinger D, Lion T, Klein R. 2013. An adenoviral vector–based expression and delivery system for the inhibition of wild-type adenovirus replication by artificial microRNAs. Antiviral Res. 97:10–23 146. Ylosm¨ aki¨ E, Hakkarainen T, Hemminki A, Visakorpi T, Andino R, Saksela K. 2008. Generation of a conditionally replicating adenovirus based on targeted destruction of E1A mRNA by a cell type–specific microRNA. J. Virol. 82:11009–15 147. Langlois RA, Albrecht RA, Kimble B, Sutton T, Shapiro JS, et al. 2013. MicroRNA-based strategy to mitigate the risk of gain-of-function influenza studies. Nat. Biotechnol. 31:844–47 148. Gall A, Palser A. 2013. An elephantine viral problem. Nat. Rev. Microbiol. 11:512 149. Furuse Y, Dastjerdi A, Steinbach F, Cullen BR. 2014. Analysis of viral microRNA expression by elephant endotheliotropic herpesvirus 1. Virology 454–55:102–8 150. Ross N, Gandhi MK, Nourse JP. 2013. The Epstein–Barr virus microRNA BART11-5p targets the early B-cell transcription factor EBF1. Am. J. Blood Res. 3:210–24 151. Lung RW, Tong JH, Sung Y, Leung P, Ng DC, et al. 2009. Modulation of LMP2A expression by a newly identified Epstein–Barr virus–encoded microRNA miR-BART22. Neoplasia 11:1174–84

Access provided by 45.16.157.220 on 10/02/15. For personal use only. 152. Tang S, Patel A, Krause PR. 2009. Novel less-abundant viral microRNAs encoded by herpes simplex virus 2 latency-associated transcript and their roles in regulating ICP34.5 and ICP0 mRNAs. J. Virol. 83:1433–42 Annual Review of Virology 2014.1:89-109. Downloaded from www.annualreviews.org 153. Tang S, Bertke AS, Patel A, Wang K, Cohen JI, Krause PR. 2008. An acutely and latently expressed herpes simplex virus 2 viral microRNA inhibits expression of ICP34.5, a viral neurovirulence factor. Proc. Natl. Acad. Sci. USA 105:10931–36 154. Jaber T, Workman A, Jones C. 2010. Small noncoding RNAs encoded within the bovine herpesvirus 1 latency-related gene can reduce steady-state levels of infected cell protein 0 (bICP0). J. Virol. 84:6297– 307 155. Waidner LA, Burnside J, Anderson AS, Bernberg EL, German MA, et al. 2011. A microRNA of infectious laryngotracheitis virus can downregulate and direct cleavage of ICP4 mRNA. Virology 411:25–31 156. Muylkens B, Coupeau D, Dambrine G, Trapp S, Rasschaert D. 2010. Marek’s disease virus microRNA designated Mdv1-pre-miR-M4 targets both cellular and viral genes. Arch. Virol. 155:1823–37 157. Strassheim S, Stik G, Rasschaert D, Laurent S. 2012. Mdv1-miR-M7-5p, located in the newly identified first intron of the latency-associated transcript of Marek’s disease virus, targets the immediate-early genes ICP4 and ICP27. J. Gen. Virol. 93:1731–42

108 Cox · Sullivan VI01CH06-Sullivan ARI 2 September 2014 6:40

158. Singh CP, Singh J, Nagaraju J. 2012. A baculovirus-encoded microRNA (miRNA) suppresses its host miRNA biogenesis by regulating the Exportin-5 cofactor Ran. J. Virol. 86:7867–79 159. Hussain M, Taft RJ, Asgari S. 2008. An insect virus–encoded microRNA regulates viral replication. J. Virol. 82:9164–70 160. Wu YL, Wu CP, Liu CYY, Hsu PWC, Wu EC, Chao YC. 2011. A non-coding RNA of insect HzNV-1 virus establishes latent viral infection through microRNA. Sci. Rep. 1:60 161. He Y, Yang K, Zhang X. 2014. Viral microRNAs targeting virus genes promote virus infection in shrimp in vivo. J. Virol. 88:1104–12 162. Yanagawa-Matsuda A, Kitamura T, Higashino F, Yamano S, Totsuka Y, Shindoh M. 2012. E1A expres- sion might be controlled by miR-214 in cells with low adenovirus productivity. Virus Res. 170:85–90 163. Sagan SM, Sarnow P, Wilson JA. 2013. Modulation of GB virus B RNA abundance by microRNA-122: dependence on and escape from microRNA-122 restriction. J. Virol. 87:7338–47 164. Zhang G, Li Y, Zheng S, Liu M, Li X, Tang H. 2010. Suppression of hepatitis B virus replication by microRNA-199a-3p and microRNA-210. Antiviral Res. 88:169–75 165. Yan Q, Li W, Tang Q, Yao S, Lv Z, et al. 2013. Cellular microRNAs 498 and 320d regulate herpes simplex virus 1 induction of Kaposi’s sarcoma–associated herpesvirus lytic replication by targeting RTA. PLoS ONE 8:e55832 166. Ma YJ, Yang J, Fan XL, Zhao HB, Hu W, et al. 2012. Cellular microRNA let-7c inhibits M1 protein expression of the H1N1 influenza A virus in infected human lung epithelial cells. J. Cell. Mol. Med. 16:2539–46 167. Sun JZ, Wang J, Yuan D, Wang S, Li Z, et al. 2013. Cellular microRNA miR-181b inhibits replication of mink enteritis virus by repression of non-structural protein 1 translation. PLoS ONE 8:e81515 168. Wen B, Dai H, Yang Y, Zhuang Y, Sheng R. 2013. MicroRNA-23b inhibits enterovirus 71 replication through downregulation of EV71 VPL protein. Intervirology 56:195–200 169. Zheng Z, Ke X, Wang M, He S, Li Q, et al. 2013. Human microRNA hsa-miR-296-5p suppresses enterovirus 71 replication by targeting the viral genome. J. Virol. 87:5645–56 170. Wang L, Qin Y, Tong L, Wu S, Wang Q, et al. 2012. miR-342-5p suppresses coxsackievirus B3 biosyn- thesis by targeting the 2C-coding region. Antiviral Res. 93:270–79 ∗ 171. Tong L, Lin L, Wu S, Guo Z, Wang T, et al. 2013. Mir-10a up-regulates coxsackievirus B3 biosynthesis by targeting the 3D-coding sequence. Nucleic Acids Res. 41:3760–71 172. Sisk JM, Witwer KW, Tarwater PM, Clements JE. 2013. SIV replication is directly downregulated by four antiviral miRNAs. Retrovirology 10:95 Access provided by 45.16.157.220 on 10/02/15. For personal use only. Annual Review of Virology 2014.1:89-109. Downloaded from www.annualreviews.org

www.annualreviews.org • Viral MicroRNAs 109 VI01-FrontMatter ARI 19 August 2014 7:19

Annual Review of Virology Contents Volume 1, 2014

Forty Years with Emerging Viruses C.J. Peters pppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppp1 Inventing Viruses William C. Summers ppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppp25 PHIRE and TWiV : Experiences in Bringing Virology to New Audiences Graham F. Hatfull and Vincent Racaniello pppppppppppppppppppppppppppppppppppppppppppppppppp37 Viruses and the Microbiota Christopher M. Robinson and Julie K. Pfeiffer pppppppppppppppppppppppppppppppppppppppppppppp55 Role of the Vector in Arbovirus Transmission Michael J. Conway, Tonya M. Colpitts, and Erol Fikrig ppppppppppppppppppppppppppppppppppp71 Balance and Stealth: The Role of Noncoding RNAs in the Regulation of Virus Gene Expression Jennifer E. Cox and Christopher S. Sullivan pppppppppppppppppppppppppppppppppppppppppppppppp89 Thinking Outside the Triangle: Replication Fidelity of the Largest RNA Viruses Everett Clinton Smith, Nicole R. Sexton, and Mark R. Denison ppppppppppppppppppppppppp111 The Placenta as a Barrier to Viral Infections Elizabeth Delorme-Axford, Yoel Sadovsky, and Carolyn B. Coyne ppppppppppppppppppppppp133

Access provided by 45.16.157.220 on 10/02/15. For personal use only. Cytoplasmic RNA Granules and Viral Infection Wei-Chih Tsai and Richard E. Lloyd pppppppppppppppppppppppppppppppppppppppppppppppppppppp147

Annual Review of Virology 2014.1:89-109. Downloaded from www.annualreviews.org Mechanisms of Virus Membrane Fusion Proteins Margaret Kielian pppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppp171 Oncolytic Poxviruses Winnie M. Chan and Grant McFadden ppppppppppppppppppppppppppppppppppppppppppppppppppp191 Herpesvirus Genome Integration into Telomeric Repeats of Host Cell Chromosomes Nikolaus Osterrieder, Nina Wallaschek, and Benedikt B. Kaufer pppppppppppppppppppppppp215

ix VI01-FrontMatter ARI 19 August 2014 7:19

Viral Manipulation of Plant Host Membranes Jean-Fran¸cois Lalibert´e and Huanquan Zheng pppppppppppppppppppppppppppppppppppppppppppp237 IFITM-Family Proteins: The Cell’s First Line of Antiviral Defense Charles C. Bailey, Guocai Zhong, I-Chueh Huang, and Michael Farzan ppppppppppppppp261 Glycan Engagement by Viruses: Receptor Switches and Specificity Luisa J. Str¨oh and Thilo Stehle ppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppp285 Remarkable Mechanisms in Microbes to Resist Phage Infections Ron L. Dy, Corinna Richter, George P.C. Salmond, and Peter C. Fineran ppppppppppppp307 Polydnaviruses: Nature’s Genetic Engineers Michael R. Strand and Gaelen R. Burke ppppppppppppppppppppppppppppppppppppppppppppppppppp333 Human Cytomegalovirus: Coordinating Cellular Stress, Signaling, and Metabolic Pathways Thomas Shenk and James C. Alwine ppppppppppppppppppppppppppppppppppppppppppppppppppppppp355 Vaccine Development as a Means to Control Dengue Virus Pathogenesis: Do We Know Enough? Theodore C. Pierson and Michael S. Diamond pppppppppppppppppppppppppppppppppppppppppppp375 Archaeal Viruses: Diversity, Replication, and Structure Nikki Dellas, Jamie C. Snyder, Benjamin Bolduc, and Mark J. Young ppppppppppppppppp399 AAV-Mediated Gene Therapy for Research and Therapeutic Purposes R. Jude Samulski and Nicholas Muzyczka ppppppppppppppppppppppppppppppppppppppppppppppppp427 Three-Dimensional Imaging of Viral Infections Cristina Risco, Isabel Fern´andez de Castro, Laura Sanz-S´anchez, Kedar Narayan, Giovanna Grandinetti, and Sriram Subramaniam pppppppppppppppppppppppppppppppppppp453 New Methods in Tissue Engineering: Improved Models for Viral Infection Vyas Ramanan, Margaret A. Scull, Timothy P. Sheahan, Charles M. Rice,

Access provided by 45.16.157.220 on 10/02/15. For personal use only. and Sangeeta N. Bhatia pppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppp475 Live Cell Imaging of Retroviral Entry Annual Review of Virology 2014.1:89-109. Downloaded from www.annualreviews.org Amy E. Hulme and Thomas J. Hope ppppppppppppppppppppppppppppppppppppppppppppppppppppppp501 Parvoviruses: Small Does Not Mean Simple Susan F. Cotmore and Peter Tattersall pppppppppppppppppppppppppppppppppppppppppppppppppppp517 Naked Viruses That Aren’t Always Naked: Quasi-Enveloped Agents of Acute Hepatitis Zongdi Feng, Asuka Hirai-Yuki, Kevin L. McKnight, and Stanley M. Lemon ppppppppp539

xContents VI01-FrontMatter ARI 19 August 2014 7:19

In Vitro Assembly of Retroviruses Di L. Bush and Volker M. Vogt pppppppppppppppppppppppppppppppppppppppppppppppppppppppppppp561 The Impact of Mass Spectrometry–Based Proteomics on Fundamental Discoveries in Virology Todd M. Greco, Benjamin A. Diner, and Ileana M. Cristea ppppppppppppppppppppppppppppp581 Viruses and the DNA Damage Response: Activation and Antagonism Micah A. Luftig ppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppppp605 Errata An online log of corrections to Annual Review of Virology articles may be found at http://www.annualreviews.org/errata/virology Access provided by 45.16.157.220 on 10/02/15. For personal use only. Annual Review of Virology 2014.1:89-109. Downloaded from www.annualreviews.org

Contents xi ANNUAL REVIEWS It’s about time. Your time. It’s time well spent.

Now Available from Annual Reviews: Annual Review of Virology virology.annualreviews.org • Volume 1 • September 2014 Editor: Lynn W. Enquist, Princeton University The Annual Review of Virology captures and communicates exciting advances in our understanding of viruses of animals, plants, bacteria, archaea, fungi, and protozoa. Reviews highlight new ideas and directions in basic virology, viral disease mechanisms, virus-host interactions, and cellular and immune responses to virus infection, and reinforce the position of viruses as uniquely powerful probes of cellular function.

Complimentary online access to the first volume will be available until September 2015.

TABLE OF CONTENTS: • Forty Years with Emerging Viruses, C.J. Peters • Polydnaviruses: Nature’s Genetic Engineers, Michael R. Strand, • Inventing Viruses, William C. Summers Gaelen R. Burke • PHIRE and TWiV: Experiences in Bringing Virology to New Audiences, • Human Cytomegalovirus: Coordinating Cellular Stress, Signaling, Graham F. Hatfull, Vincent Racaniello and Metabolic Pathways, Thomas Shenk, James C. Alwine • Viruses and the Microbiota, Christopher M. Robinson, Julie K. Pfeiffer • Vaccine Development as a Means to Control Dengue Virus Pathogenesis: Do We Know Enough? Theodore C. Pierson, • Role of the Vector in Arbovirus Transmission, Michael J. Conway, Michael S. Diamond Tonya M. Colpitts, Erol Fikrig • Archaeal Viruses: Diversity, Replication, and Structure, Nikki Dellas, • Balance and Stealth: The Role of Noncoding RNAs in the Regulation Jamie C. Snyder, Benjamin Bolduc, Mark J. Young of Virus Gene Expression, Jennifer E. Cox, Christopher S. Sullivan • AAV-Mediated Gene Therapy for Research and Therapeutic Purposes, • Thinking Outside the Triangle: Replication Fidelity of the Largest RNA R. Jude Samulski, Nicholas Muzyczka Viruses, Everett Clinton Smith, Nicole R. Sexton, Mark R. Denison • Three-Dimensional Imaging of Viral Infections, Cristina Risco, • The Placenta as a Barrier to Viral Infections, Isabel Fernández de Castro, Laura Sanz-Sánchez, Kedar Narayan, Elizabeth Delorme-Axford, Yoel Sadovsky, Carolyn B. Coyne Giovanna Grandinetti, Sriram Subramaniam Wei-Chih Tsai, • Cytoplasmic RNA Granules and Viral Infection, • New Methods in Tissue Engineering: Improved Models for Viral Richard E. Lloyd Infection, Vyas Ramanan, Margaret A. Scull, Timothy P. Sheahan, • Mechanisms of Virus Membrane Fusion Proteins, Margaret Kielian Charles M. Rice, Sangeeta N. Bhatia Access provided by 45.16.157.220 on 10/02/15. For personal use only. • Oncolytic Poxviruses, Winnie M. Chan, Grant McFadden • Live Cell Imaging of Retroviral Entry, Amy E. Hulme, Thomas J. Hope • Herpesvirus Genome Integration into Telomeric Repeats of Host • Parvoviruses: Small Does Not Mean Simple, Susan F. Cotmore,

Annual Review of Virology 2014.1:89-109. Downloaded from www.annualreviews.org Cell Chromosomes, Nikolaus Osterrieder, Nina Wallaschek, Peter Tattersall Benedikt B. Kaufer • Naked Viruses That Aren’t Always Naked: Quasi-Enveloped Agents • Viral Manipulation of Plant Host Membranes, Jean-François Laliberté, of Acute Hepatitis, Zongdi Feng, Asuka Hirai-Yuki, Kevin L. McKnight, Huanquan Zheng Stanley M. Lemon • IFITM-Family Proteins: The Cell’s First Line of Antiviral Defense, • In Vitro Assembly of Retroviruses, Di L. Bush, Volker M. Vogt Charles C. Bailey, Guocai Zhong, I-Chueh Huang, Michael Farzan • The Impact of Mass Spectrometry–Based Proteomics on Fundamental • Glycan Engagement by Viruses: Receptor Switches and Specificity, Discoveries in Virology, Todd M. Greco, Benjamin A. Diner, Luisa J. Ströh, Thilo Stehle Ileana M. Cristea • Remarkable Mechanisms in Microbes to Resist Phage Infections, • Viruses and the DNA Damage Response: Activation and Antagonism, Ron L. Dy, Corinna Richter, George P.C. Salmond, Peter C. Fineran Micah A. Luftig

ANNUAL REVIEWS | Connect With Our Experts

Tel: 800.523.8635 (us/can) | Tel: 650.493.4400 | Fax: 650.424.0910 | Email: [email protected]