Available online at www.sciencedirect.com

ScienceDirect MicroRNAs and oncolytic viruses Autumn J Ruiz and Stephen J Russell MicroRNAs regulate gene expression in mammalian cells and often exhibit tissue-specific expression patterns. Incorporation of microRNA target sequences can be used to control exogenous gene expression and viral tropism in specific tissues to enhance the therapeutic indices of oncolytic viruses expressing therapeutic transgenes. Continued development of this targeting strategy has resulted in the generation of unattenuated oncolytic viruses with enhanced potency, broad species-tropisms and reduced off-target toxicities in multipletissues simultaneously. Furthermore, oncolytic viruses have been used to enhance the delivery, duration and therapeutic efficacy of microRNA-based therapeutics designed to either restore or inhibit the function of dysregulated microRNAs in cancer cells. Recent efforts focused on combining oncolytic virotherapy and microRNA regulation have generated increasingly potent and safe cancer therapeutics. Address Department of Molecular Medicine, Mayo Clinic College of Medicine, Rochester, MN 55905, United States Corresponding author: Russell, Stephen J ([email protected])

Current Opinion in Virology 2015, 13:40–48

roles in regulating cellular processes such as cell cycle, differentiation, metabolism, apoptosis, autophagy, signaling, inflammation, and migration [13–17]. Therefore, miRNA deregulation is commonly associated with disease, including many types of cancer [18–20]. miRNA signatures can distinguish between normal and cancerous cells, as well as different cancer subtypes, specific oncogenic abnormalities, and cancer stem cells [21]. An overview of cancers with miRNA signatures is given in Table 2. Major efforts have been focused on exploiting miRNA regulation for developing new and enhanced therapeutic interventions; as such the fields of miRNAbased therapeutics and oncolytic viruses (OVs) have converged. Although well tolerated, the clinical activity of both of these therapies as stand-alone treatments remains limited [22,23,24]. Increasing evidence suggests that enhanced delivery, increased potency, and combinatorial therapies are necessary for these therapeutics to reach their full potential. Here we review how OVs have been coupled with miRNA regulation to enhance the efficacy and translatability of miRNA-based therapeutics and oncolytic virotherapy and highlight major advances in the field over the past several years.

This review comes from a themed issue on Oncolytic viruses Edited by Grant McFadden and John Bell

http://dx.doi.org/10.1016/j.coviro.2015.03.007 1879-6257/# 2015 Published by Elsevier B.V.

Introduction MicroRNAs (miRNAs) are small, 19–25 nt, non-coding RNAs expressed by all multicellular eukaryotes that mediate post-transcriptional regulation of gene expression [1–5]. miRNAs silence gene expression by binding sequence-complementary target elements in messenger RNAs (mRNAs), commonly found in the 30 untranslated region (UTR), preventing translation or accelerating transcript degradation. Canonical and non-canonical miRNA biogenesis pathways are described in Figure 1, and the reader is referred to Refs. [4,6,7] for more detail. More than 2500 mature miRNAs have been identified in human cells [8–12] and while some miRNAs are ubiquitously expressed, many are enriched within certain tissues (Table 1) and their expression can be lineage-specific, activation, or differentiation stage-specific. They are known to regulate 60% of human genes and play major Current Opinion in Virology 2015, 13:40–48

MicroRNA regulated transgene expression Introduction of exogenous therapeutic genes often requires cell-specific expression in order to avoid offtarget effects. The idea of exploiting miRNA regulation to restrict gene expression in specific tissues was initially demonstrated using hematopoietic-specific miR-142-3p targets to inhibit lentiviral delivered transgene expression in antigen presenting cells [25,26]. By incorporating miRNA target sequences into the 30 UTR of the transgene, expression was significantly reduced in tissues expressing the specific miRNA, but maintained in those that did not. This concept was quickly adopted for preventing off-target toxicities and increasing tissue specificity of virally encoded genetic cassettes [27,28]. Exogenous gene expression from viral vectors including adenovirus (Ad), adeno-associated virus (AAV), and baculovirus can be controlled and the therapeutic indices of these treatments enhanced by including miRNA response elements (MREs) [29–31]. The utility of this targeting strategy has been widely demonstrated and tissue-specificity has been enhanced for a large number of virally encoded therapeutic genes. Most recently, MREs have been used to control the expression of the TNF-related apoptosis-inducing ligand (TRAIL), a cytokine that selectively triggers apoptosis and that demonstrated robust anticancer activity in preclinical studies. Clinical analysis of TRAIL-receptor agonists revealed that although well tolerated these agents did not have www.sciencedirect.com

MicroRNAs and oncolytic viruses Ruiz and Russell 41

Figure 1

Drosha-independent “Mirtron”

Canonical RNA Pol II

AAAA m7G

m7G

AAAA

exon

m7G

exon

Dicer-independent

AAAA

m7G

AAAA

Spliceosome Drosha

Drosha

DGCR8

DGCR8 RAN-GTP Exportin 5

Ran-GDP

Dicer

Ran-GDP

P Body

TRBP m7G

Deadenylation mRNA degradation Ago 2

PARN

mRNA Storage

Ago 1-4

Passenger Strand Degraded

Ago 2

m7G

AAAA RISC Current Opinion in Virology

Schematic representation of miRNA biogenesis and function. Canonical. miRNAs are transcribed as capped, polyadenylated primary miRNA (primiRNA) hairpins, typically by RNA polymerase II, which are then recognized by the nuclear processor complex (Drosha-DGCR8) and cleaved 22 nt down the miRNA stem. The resulting pre-miRNA hairpins bearing 2 nt 30 overhangs are translocated to the cytoplasm by exportin 5 and are further processed by Dicer-TRBP into double-stranded miRNA duplexes. miRNA duplexes are loaded into Argonaute (AGO)-containing RNAinduced silencing complexes (RISCs), followed by duplex unwinding, mature miRNA retention and passenger strand degradation. Some passengers are functional miRNAs and do not get degraded. The miRNA directs the RISC to partially complementary sites in target mRNAs resulting in translational repression or accelerated transcript degradation by retargeting mRNAs to P bodies. If the miRNA recognizes a sequence of perfect complementarity, the transcript can undergo endonucleolytic cleavage. Drosha-independent. Short intronic hairpins termed mirtons represent the most common alternative miRNA biogenesis pathway. These hairpins are spliced and debranched to form pre-miRNAs, which are subsequently processed through the canonical pathway. Dicer-independent. Pre-miRNAs are produced by Drosha and exported to the cytoplasm, presumably by exportin 5. These pre-miRNAs are loaded on AGO2, which cleaves the pre-miRNA stem, and are then further trimmed by the 30 –50 exonuclease poly(A)-specific ribonuclease (PARN). The mature miRNA-loaded RISC then mediates gene silencing.

www.sciencedirect.com

Current Opinion in Virology 2015, 13:40–48

42 Oncolytic viruses

Table 1 MicroRNAs enriched in specific tissues Tissue/cell type Bone Cardiac Endocrine Glands Hematopoietic Kidney Liver Lung Muscle Nervous System

Ovary Pancreas Spleen Testis

Enriched miRNAs

References

miR-377; miR-92a; miR-483 miR-1; miR-126; miR-133a/b; miR-208a; miR-302; miR-367 miR-375 miR-16; miR-142; miR-150; miR-223 miR-10a/b; miR-30c; miR-146a; miR-192; miR-196a/b; miR-200a; miR-204; miR-215; miR-216 miR-92a; miR-122; miR-192; miR-483 miR-126 miR-1; miR-95; miR-128a; miR-133a/b; miR-134; miR-193a; miR-206 miR-7; miR-9; miR-31; miR-33a; miR-93; miR-95; miR-99b; miR-124a; miR-125a/b; miR-128a/b; miR-129; miR-138; miR-149; miR-186; miR-199a/b, miR-212; miR-214; miR-137; miR-143; miR-153; miR-218; miR-323; miR-346; miR-330; miR-708 miR-542-5p miR-216a/b; miR-217; miR-375 miR-146a; miR-223 miR-10b; miR-23a; miR-34b/c; miR-134; miR-187; miR-202; miR-204; miR-449a; miR-506; miR-507; miR-508; miR-509; miR-510; miR-513; miR-514; miR-892b

[74] [74,75] [76] [75,76] [74,75] [74,75] [74] [74–76] [74–76]

[75] [74,75] [74] [74,75]

*Given the volume of manuscripts involved in obtaining the above data, reviews were cited instead of original reports to limit the number of references.

robust antitumor activity. One possible explanation for this was that relatively weak agonists were chosen because of initial concerns over potential hepatotoxicity. Therefore, stronger agonists may provide a better clinical outcome as long as the off-target toxicities are prevented. Several groups showed that MREs could confer tumorspecificity and reduce toxicity of TRAIL therapy against

glioma, uveal melanoma, prostate, and osteosarcoma models [32,33]. MREs can also be used to facilitate the manufacture of viruses expressing transgenes. Savdaminova et al. [34] used MREs to generate modified Ads expressing nucleases for cell type specific genome editing. Nuclease

Table 2 Cancers with miRNA signatures Cancer Breast Bladder CLL Colon Colorectal Cancer Esophagus Glioblastoma

Down-regulated miRNA Let-7; miR-31; miR-34; miR-200; miR-126; miR-29 miR-34; miR-143; miR-133a/b miR-15/16; miR-29 miR-Let-7a-2; miR-34, miR-107 miR-143, miR-145

Kidney Leukemia Liver

miR-221/222; miR-34; miR-7; miR-124; miR-128; miR-218 miR-34 miR-29 # miR-34; miR-122a; miR-26a;

Lung

miR-Let-7; miR-34; miR-29

Lymphoma Myeloma Ovarian

miR-155; miR-29

Pancreas Prostate Stomach Thyroid

miR-Let-7a-2; miR-31; miR-34; miR-143; miR-199a; miR-200 miR-107; miR-96; miR-196 miR-Let-7c; miR-125b; miR-145; miR-15/16 miR-Let-7a-2; miR-31; miR-126;

Up-regulated miRNA miR-21; miR-221/222; miR-10b; miR-17-92; miR-155 miR-21; miR-155 miR-21; miR-196; miR-17-92; miR-155 miR-31 miR-21; miR-10b; miR-196 miR-21; #miR-10b; miR-196

miR-21; miR-155 # miR-21; #miR-221/222; miR-181b miR-21; miR-221/222; miR-17-92; miR-155 miR-155 miR-21, miR-181b miR-21, miR-200a/b/c

References [20,77] [75,77] [75,77,78] [20,77] [75,78] [20] [20,77,78,79] [77] [77,78] [20,75,77,78] [20,75,77,78]

[75,77,78] [20,77] [20,75,77]

miR-21 miR-21

[20,77] [75,77]

miR-21 miR-221/222; miR-146

[20,77] [77]

*Given the volume of manuscripts involved in obtaining the above data, reviews were cited instead of original reports to limit the number of references. miRNA-based therapeutics currently undergoing preclinical and clinical analysis.

#

Current Opinion in Virology 2015, 13:40–48

www.sciencedirect.com

MicroRNAs and oncolytic viruses Ruiz and Russell 43

expression was inhibited in the producer cells allowing virus rescue without restricted expression in target cells.

Modifying virus tropism MicroRNA targeting can also be used to regulate oncolytic viral tropism to enhance tumor specificity and eliminate undesirable toxicities [35]. This was originally demonstrated by incorporating muscle-specific MREs into the 30 UTR of Coxsackievirus A21 (CVA21), which normally causes fatal myositis [36]. Inclusion of MREs eliminated muscle toxicity but did not compromise oncolytic activity. This strategy has since been applied to many viruses and may facilitate clinical translation of OVs with enhanced potency. The safety of therapeutic viruses can be improved by deleting essential viral genes, however this is often associated with viral attenuation and decreased potency [37–40]. By contrast, MREs can control gene expression without resulting in attenuation. Deletions within the E1A and E1B genes are often used to generate tumor-selective conditionally replicating Ads, but result in viral attenuation. Yao et al. [41] constructed an oncolytic Ad with a MRE controlled E1A gene. This virus, OA-4MREs, resulted in increased viral output and cytotoxicity in cultured and primary glioma cells compared to ONYX-015, a clinically tested Ad with an E1B region deletion. By contrast to ONYX-015, OA-4MREs inhibited neuro-toxicity and hepatotoxicity, displayed enhanced antitumor activity and prolonged survival of tumor-bearing mice. Callegari et al. [42] targeted miR-199 to regulate oncolytic Ad replication for improved antitumor activity and reduced hepatotoxicity in both tumorbearing, newborn mice treated intratumorally and in immune-competent miR-221 transgenic mice with diethylnitrosamine induced liver cancer treated intravenously (IV). The importance here is that few preclinical studies are conducted in IV-treated, tumor-bearing, immune competent mice, a model that most closely mimics a clinical setting. While preclinical data suggest miRNA targeting may be sufficient to eliminate toxicities in animal models, combining targeting strategies, especially for unattenuated viruses may prove more prudent in humans. Placing an E1A gene under liver-specific miR-122 control in Ad6, which has reduced liver toxicity, significantly improved safety versus unmodified Ad5 or Ad6 and allowed administration of elevated doses [38]. The tropism of unattenuated oncolytic herpes simplex viruses (oHSVs) can also be modified by MREs [43]. Most recently, Mazzacurati et al. [44] developed a wild-type oHSV modified to bind epidermal growth factor receptor and epidermal growth factor receptor variant III and encode a miR-124 regulated ICP4 gene. Replication in normal brain tissue and toxicity were reduced in nude mice and antitumor activity against primary human glioblastoma multiforme tumors was maintained. www.sciencedirect.com

Multi-tissue detargeting Multiple tissue tropisms can be regulated simultaneously by targeting a ubiquitous miRNA or multiple different miRNAs. The Let-7 family of miRNAs function as tumor suppressors and are down-regulated in a large number of cancers. miR-Let-7 response elements have been used for broad control of live attenuated poliovirus vaccines, vesicular stomatitis virus (VSV), vaccinia virus, Ad, and oHSV [41,44,45–47]. However, Let-7 levels in different tissues are variable and target recognition is redundant among family members. Thus, even though one family member is down-regulated, other members may have ample expression levels to reduce therapeutic efficacy in the tumor and increase the possibility of miRNA saturation in normal cells. Several studies have used multiple miRNA targets simultaneously to redirect viral tropism and this has proven just as efficacious as using ubiquitous targets [26,41,48,49,50]. This may prove the preferred method as incorporation of different miRNA targets decreases the potential for saturating a single miRNA. Furthermore, the ability to simultaneously eliminate various off-target toxicities increases the potential for clinical translation of OVs with increased potency. Fu et al. [51] combined miR-Let-7, miR-122, and miR-124 targets to control the glycoprotein H gene of oHSV, with a liver-specific promoter to generate a highly targeted, potent oHSV.

Versatility of miRNA targeting DNA and RNA viruses are both amenable to miRNA targeting. To date, the majority of miRNA-targeted RNA viruses are positive-sense, single-stranded viruses. These viruses have proven highly responsive to this targeting strategy, likely because both viral genomes and mRNAs are targeted. By contrast, only a few negative-sense RNA viruses including influenza A virus (IAV), VSV and measles virus (MV) have been targeted using MREs. miRNAtargeting of negative-strand RNA viruses is hypothesized to be more difficult as the accessibility of MREs will likely be hindered because encapsidation of the viral genome occurs concomitantly with transcription and replication. MREs oriented to target the viral genome of VSV or IAV did not result in miRNA-mediated silencing [52,53]. Therefore miRNA responses in negative sense RNA viruses are likely limited to mRNAs, increasing the burden for miRNA targeting by allowing continual production of mRNAs and increasing the potential for saturation of miRNAs. MicroRNA-targeting has been applied to IAV to generate a safe and effective live-attenuated vaccine as well as a virus that is transmissible among ferrets, but not among humans [54,55]. Additionally, Chua et al. [56] used miRNA-targeting technology to facilitate mechanistic studies on IAV replication. Interestingly, miRNA-targeting of the non-structural protein 1 resulted in significantly down-regulated expression both on an mRNA and proCurrent Opinion in Virology 2015, 13:40–48

44 Oncolytic viruses

tein level, but did not affect virus replication in vitro or in vivo. This demonstrates the importance of selecting genes whose expression level dramatically affects viral replication for optimal targeting. MV has also proven amenable to miRNA targeting [57]. Most recently a multi-tissue detargeted MV regulated by miR-7, miR-122 and miR-148a was generated [49]. Viral spread in the presence of these miRNAs was restricted, but antitumor activity against pancreatic carcinoma xenografts was unaltered. Unfortunately, the utility of this targeting strategy is difficult to assess because MV cannot replicate in murine cells. However, this study may have clinical relevance as MV has shown clinical efficacy at extremely high doses [58]. As the dose and potency of this virus are elevated to maximize therapeutic efficacy, toxicities in off-target tissues may be an issue. Studies focused on redirecting the tropism of VSV demonstrated that miRNA targeting of the polymerase gene (L) was more efficacious than targeting the gene encoding the matrix protein (M). This could be due to several different factors including, firstly, lower levels of L protein due to the transcriptional gradient of VSV; secondly, larger impact on viral replication that may mimic targeting of viral genomes; or thirdly, higher accessibility of the miRNA target. VSV tropism was redirected and neurotoxicity reduced when miR-125b response elements were used to control L protein expression. 90% of mice injected with this virus were protected against neuropathogenesis following an intracerebral injection of a lethal dose for unmodified VSV. However, the one mouse that developed encephalitis did not have mutations within the target sequences indicating that this MRE configuration was insufficient for eliminating neurotoxicity. The in vitro and in vivo results of this targeted virus were inferior to that observed with other targeted viruses and this may be attributed to various factors. miRNA silencing is more efficient when viral genomes are targeted compared to viral mRNAs, however this has not been achieved in negative-sense RNA viruses. VSV is a rapidly replicating virus and mRNAs will continually be synthesized increasing the potential for miRNA recognition escape and or saturation. Targeting of a single gene may not be sufficient and control of various genes simultaneously may be warranted. Location within the 30 UTR may not be optimal for target insertion and enhanced control may be achieved by incorporating the miRNA target within the open reading frame of the gene as was demonstrated for Sindbis virus [59]. Finally, the miRNAs chosen may not be optimal. It was recently demonstrated that inhibition of several miRNAs including miR-125b increased susceptibility of human bronchial epithelial cells to VSV infection [60]. The affects of viruses and targeted miRNAs on the cellular transcriptome, antiviral state and inflammatory response need to be determined in order to engineer safe and effective miRNA-targeting strategies. Current Opinion in Virology 2015, 13:40–48

Guidelines for optimal targeting Many factors contribute to targeting efficacy [61]. Elevated miRNA expression levels correlate with increased regulatory activity and certain miRNAs are more efficient at gene silencing than others [52,61]. Completely complementary target sequences also promote miRNA-gene regulation and reduce the potential for miRNA saturation because perfectly complementary sequences can be endonucleolytically cleaved allowing rapid recycling of the miRNA. This does require that the miRNA associate with an Argonaute protein with endonucleolytic activity. Increasing the copy number of target sequences generally will enhance targeting efficiency, however this is not always the case and defining an optimal target number for each miRNA is recommended. The level of targeted gene expression, attributed to the replication kinetics and transcription gradient of the virus, can affect the regulatory capacity of the MREs as will the level of competing mRNAs targeted by the miRNA being utilized. Finally, the insertion site is a major factor in dictating suppressive activity. While most MRE elements are found within the 30 UTR of mRNAs, relocation to the 50 UTR or the open reading frame of a gene or genome can be more efficacious [59]. Insertion sites need to allow high accessibility of the miRNA target, which can be hindered by secondary structures normally formed by the virus or additional structures introduced by the incorporated target sequences. Studies in our laboratory revealed that while various MREs inserted into the 30 UTR of Mengovirus regulate viral replication, recognition of Let-7a inserted at the same site is prohibited (Ruiz et al. unpublished data). In general, selecting a miRNA with high expression levels, and incorporating 4–6 copies of MREs with complete complementarity, in a site with low secondary structure and high conservation is an optimal starting place when ablation of gene expression is desired [61]. However, trial and error, both at the level of viral genome engineering to make stable recombinant viruses, and at the whole organism level in a relevant model is absolutely required for evaluating and optimizing miRNA regulation.

Potential caveats While miRNA targeting offers many advantages to traditional strategies, it is not without potential caveats. Given the amplification potential and high error rates of replication-competent viruses, there are two common concerns when using this technology. The first is the potential evolution of escape mutants, which has been demonstrated with miRNA-targeted CVA21, poliovirus, and Dengue virus [36,45,62]. Using perfectly complementary target sequences in tandem and at various stable locations throughout the genome decreases the potential for the evolution of escape mutants [59,61]. Additionally, MREs are differentially effective and susceptible to different selective pressures to generate escape mutations, which may be target and/or tissue-specific [63]. www.sciencedirect.com

MicroRNAs and oncolytic viruses Ruiz and Russell 45

Figure 2

Normal Cell TS-miRNA

Viral protein expression Transgene expression Viral genome transcription

No targeted miRNA

TS promoter

Host miRNA mimic No targeted genes

Viral protein expression Transgene expression Viral replication

Viral gene Transgene MREs

Malignant Cell

miRNA-RISC

OncomiR

Viral replication Transgene expression Tumor cell lysis Antitumor immune response Viral genome transcription

miRNA decoy/sponge

miRNA restoration

Viral replication Transgene expression Tumor cell lysis Antitumor immune response

Current Opinion in Virology

Overview of miRNA regulated oncolytic viruses and viral delivery of miRNA therapeutics. Tissue-specific (TS) MREs can be built into nuclear or cytoplasmic replicating viruses to control expression of viral genes/genomes or therapeutic transgenes in normal cells expressing the respective TS-miRNAs, but allowing replication, lysis, and immune response induction in malignant cells. DNA and RNA viruses can be engineered to express artificial miRNA mimics to restore cellular homeostasis, target oncogenes, enhance OV destruction, or facilitate the antitumor immune response. OVs can also be engineered to express miRNA decoys or sponges that are recognized by specific miRNAs. Sequestration of miRNAs can allow expression of genes to enhance OVs or facilitate the immune response. miRNA sponges can also function to sequester oncomiRs. Encoding miRNA-based therapeutics in miRNA-targeted OVs can further prevent off-target toxicities.

www.sciencedirect.com

Current Opinion in Virology 2015, 13:40–48

46 Oncolytic viruses

This emphasizes the importance of trial and error when engineering miRNA-regulated OVs. This issue may be less of a concern in an immune-competent host when the virus can be cleared before evolution. The second concern is the potential for off-target effects on the host miRNA machinery. Saturation of the miRNA machinery could allow toxicity to reoccur or alter the cellular transcriptome and result in disease. Again, for redirecting viral tropism, using completely complementarity targets allowing rapid recycling of the miRNA, using different MREs, and determining the minimal number of miRNA targets needed will reduce the potential for miRNA saturation [61,64]. Introducing miRNA-based therapeutics into a host cell can similarly saturate the proteins involved in miRNA biogenesis (Figure 1) and create bottlenecks in the pathway if their expression levels are not calibrated. Overexpression of pre-miRNAs can result in accumulation of small RNAs within the nucleus by overloading the exportin proteins involved in translocation from the nucleus to the cytoplasm. Additionally, overexpression of miRNAs can cause a buildup of RNAinduced silencing complex (RISC) substrates by overloading the available RISC complexes and result in cellular toxicities. Therefore, miRNA biogenesis and pro-inflammatory responses in normal tissues need to be evaluated in clinically relevant animal models when analyzing the safety of OVs encoding miRNA-based therapeutics. Viruses have the ability to regulate miRNA biogenesis in infected cells and this may inhibit their ability to deliver miRNA-based therapeutics. Adenoviruses express a noncoding RNA, VA1, which saturates the function of exportin 5 preventing miRNA biogenesis. Poxviruses induce the degradation of all host miRNAs, but 20 O-methylated siRNAs are protected. These factors need to be taken into consideration when designing OV delivery vehicles and the reader is referred to [65] for an in-depth discussion on the potential constraints associated with virally delivered sRNAs.

Oncolytic virus delivery of miRNAtherapeutics Major efforts are focused on developing miRNA-based therapeutics to either restore or inhibit function of deregulated miRNAs and several have recently advanced into clinical testing [24,66]. These agents include synthetic miRNAs that can be processed by host cellular machinery and mimic normal miRNA function (miRNA mimics), or overexpressed RNAs encoding miRNA complementary target sites that compete for miRNA recognition (miRNA decoy/sponge). Studies suggest OVs can enhance delivery, duration, and efficacy of some miRNA-based therapeutics [67–69]. Lou et al. [70] developed an oncolytic Ad coexpressing miR-34a, a miRNA known to silence the antiapoptotic Bcl-2 gene, and the tumor-suppressor protein, IL24. This virus had significantly improved antitumor activity Current Opinion in Virology 2015, 13:40–48

against hepatocellular carcinoma xenografts compared to viruses expressing miR-34a or IL-24 alone. Moshiri et al. [71] modified the Ad-199T virus [42] and a recombinant AAV vector to express a decoy cassette that bound miR-221. This virus resulted in a reduction of miR-221 and an increase in miR-221 target protein, however, neither vector was analyzed in vivo. While a lot of focus has been on engineering DNA viruses for miRNA therapeutic delivery, it was recently discovered that RNA viruses, including IAV, Sindbis virus, and VSV, could also deliver miRNA mimics and decoys [53,72,73]. This is especially interesting as oncolytic RNA viruses are yielding promising results in clinical trials and by incorporating miRNAs that target proinflammatory cytokines or facilitate the antitumor immune response the therapeutic efficacy may be enhanced. While intriguing, caution must be taken when combining therapies as viral expression of miRNAs can cause toxicity in normal tissues if not properly controlled [65].

Conclusions miRNA-targeting can increase the therapeutic index of OVs and gene therapy by precisely detargeting expression in a species and cell-specific manner. As such an arsenal of OVs and therapeutic genes showing enhanced potency and reduced toxicity in preclinical models have been generated. As more tissue and cancer-specific miRNA signatures are identified, defining optimal configurations that can be tailored to individual patients becomes possible. It will also facilitate combination therapies that currently would prove far too toxic for patients. For example, it is now theoretically possible to generate an unattenuated OV encoding an immune stimulatory transgene, and an miRNA targeting an immune inhibitory protein with no off-target toxicities. If such viruses prove efficacious and safe in clinically relevant animal models, the future of cancer therapeutics will then include combinatorial therapy consisting of single-shot formulations (Figure 2).

Acknowledgments Al and Mary Agnes McQuinn, the Richard M. Schulze Family Foundation, and Mayo Clinic support studies in the Russell lab. We apologize to colleagues whose work was not cited due to space limitations.

References and recommended reading Papers of particular interest, published within the period of review, have been highlighted as:  of special interest  of outstanding interest 1.

Wightman B, Ha I, Ruvkun G: Posttranscriptional regulation of the heterochronic gene lin-14 by lin-4 mediates temporal pattern formation in C. elegans. Cell 1993, 75:855-862.

2.

Lee RC, Feinbaum RL, Ambros V: The C. elegans heterochronic gene lin-4 encodes small RNAs with antisense complementarity to lin-14. Cell 1993, 75:843-854.

3.

Ambros V: The functions of animal microRNAs. Nature 2004, 431:350-355.

4.

Bartel DP: MicroRNAs: genomics, biogenesis, mechanism, and function. Cell 2004, 116:281-297. www.sciencedirect.com

MicroRNAs and oncolytic viruses Ruiz and Russell 47

5.

Bartel DP: MicroRNAs: target recognition and regulatory functions. Cell 2009, 136:215-233.

6. Ha M, Kim VN: 1: Regulation of microRNA biogenesis. Nat Rev Mol Cell Biol 2014, 15:509-524.  Comprehensive review of microRNA biogenesis pathways: canonical and non-canonical. 7.

Li Z, Rana TM: Therapeutic targeting of microRNAs: current status and future challenges. Nat Rev Drug Discov 2014, 13:622-638.

8.

Griffiths-Jones S: The microRNA Registry. Nucleic Acids Res 2004, 32:D109-D111.

9.

Griffiths-Jones S, Grocock RJ, van Dongen S, Bateman A, Enright AJ: miRBase: microRNA sequences, targets and gene nomenclature. Nucleic Acids Res 2006, 34:D140-D144.

28. Papapetrou EP, Kovalovsky D, Beloeil L, Sant’angelo D, Sadelain M: Harnessing endogenous miR-181a to segregate transgenic antigen receptor expression in developing versus post-thymic T cells in murine hematopoietic chimeras. J Clin Invest 2009, 119:157-168. 29. Suzuki T, Sakurai F, Nakamura S, Kouyama E, Kawabata K, Kondoh M, Yagi K, Mizuguchi H: miR-122a-regulated expression of a suicide gene prevents hepatotoxicity without altering antitumor effects in suicide gene therapy. Mol Ther 2008, 16:1719-1726. 30. Geisler A, Jungmann A, Kurreck J, Poller W, Katus HA, Vetter R, Fechner H, Muller OJ: microRNA122-regulated transgene expression increases specificity of cardiac gene transfer upon intravenous delivery of AAV9 vectors. Gene Ther 2011, 18:199-209.

10. Griffiths-Jones S, Saini HK, van Dongen S, Enright AJ: miRBase: tools for microRNA genomics. Nucleic Acids Res 2008, 36:D154-D158.

31. Xie J, Xie Q, Zhang H, Ameres SL, Hung JH, Su Q, He R, Mu X, Seher Ahmed S, Park S et al.: MicroRNA-regulated, systemically delivered rAAV9: a step closer to CNS-restricted transgene expression. Mol Ther 2011, 19:526-535.

11. Kozomara A, Griffiths-Jones S: miRBase: integrating microRNA annotation and deep-sequencing data. Nucleic Acids Res 2011, 39:D152-D157.

32. Bo Y, Guo G, Yao W: MiRNA-mediated tumor specific delivery of TRAIL reduced glioma growth. J Neurooncol 2013, 112:27-37.

12. Kozomara A, Griffiths-Jones S: miRBase: annotating high confidence microRNAs using deep sequencing data. Nucleic Acids Res 2014, 42:D68-D73.

33. Liu J, Ma L, Li C, Zhang Z, Yang G, Zhang W: Tumor-targeting TRAIL expression mediated by miRNA response elements suppressed growth of uveal melanoma cells. Mol Oncol 2013, 7:1043-1055.

13. Harfe BD: MicroRNAs in vertebrate development. Curr Opin Genet Dev 2005, 15:410-415. 14. Lu J, Getz G, Miska EA, Alvarez-Saavedra E, Lamb J, Peck D, Sweet-Cordero A, Ebert BL, Mak RH, Ferrando AA et al.: MicroRNA expression profiles classify human cancers. Nature 2005, 435:834-838. 15. Jovanovic M, Hengartner MO: miRNAs and apoptosis: RNAs to die for. Oncogene 2006, 25:6176-6187. 16. Boehm M, Slack FJ: MicroRNA control of lifespan and metabolism. Cell Cycle 2006, 5:837-840. 17. Friedman RC, Farh KK, Burge CB, Bartel DP: Most mammalian mRNAs are conserved targets of microRNAs. Genome Res 2009, 19:92-105. 18. Li Y, Kowdley KV: MicroRNAs in common human diseases. Genomics Proteomics Bioinform 2012, 10:246-253. 19. Emde A, Hornstein E: miRNAs at the interface of cellular stress and disease. EMBO J 2014, 33:1428-1437. 20. Acunzo M, Romano G, Wernicke D, Croce CM: MicroRNA and cancer — a brief overview. Adv Biol Regul 2015, 57:1-9. 21. Iorio MV, Croce CM: microRNA involvement in human cancer. Carcinogenesis 2012, 33:1126-1133. 22. Russell SJ, Peng KW, Bell JC: Oncolytic virotherapy. Nat Biotechnol 2012, 30:658-670. 23. Walther W, Schlag PM: Current status of gene therapy for cancer. Curr Opin Oncol 2013, 25:659-664. 24. van Rooij E, Kauppinen S: Development of microRNA  therapeutics is coming of age. EMBO Mol Med 2014, 6:851-864. Comprehensive review of miRNA-based therapeutics currently under preclinical and clinical evaluation. 25. Brown BD, Venneri MA, Zingale A, Sergi Sergi L, Naldini L: Endogenous microRNA regulation suppresses transgene expression in hematopoietic lineages and enables stable gene transfer. Nat Med 2006, 12:585-591. 26. Brown BD, Gentner B, Cantore A, Colleoni S, Amendola M, Zingale A, Baccarini A, Lazzari G, Galli C, Naldini L: Endogenous microRNA can be broadly exploited to regulate transgene expression according to tissue, lineage and differentiation state. Nat Biotechnol 2007, 25:1457-1467. 27. Colin A, Faideau M, Dufour N, Auregan G, Hassig R, Andrieu T, Brouillet E, Hantraye P, Bonvento G, Deglon N: Engineered lentiviral vector targeting astrocytes in vivo. Glia 2009, 57:667-679. www.sciencedirect.com

34. Saydaminova K, Ye X, Wang H, Richter M, Ho M, Chen H, Xu N, Kim J-S, Papapetrou E, Holmes MC et al.: Efficient genome editing in hematopoietic stem cells with helper-dependent Ad5/35 vectors expressing site-specific endonucleases under microRNA regulation. Mol Therapy Methods Clin Dev 2015:1. 35. Kelly EJ, Russell SJ: MicroRNAs and the regulation of vector tropism. Mol Ther 2009, 17:409-416. 36. Kelly EJ, Hadac EM, Greiner S, Russell SJ: Engineering microRNA responsiveness to decrease virus pathogenicity. Nat Med 2008, 14:1278-1283. 37. Su CQ, Sham J, Xue HB, Wang XH, Chua D, Cui ZF, Peng LH, Li LF, Jiang LH, Wu MC, Qian QJ: Potent antitumoral efficacy of a novel replicative adenovirus CNHK300 targeting telomerasepositive cancer cells. J Cancer Res Clin Oncol 2004, 130:591-603. 38. Zhang Z, Zhang X, Newman K, Liu X, Seth P: MicroRNA regulation of oncolytic adenovirus 6 for selective treatment of castration-resistant prostate cancer. Mol Cancer Ther 2012, 11:2410-2418. 39. Luo J, Xia Q, Zhang R, Lv C, Zhang W, Wang Y, Cui Q, Liu L, Cai R, Qian C: Treatment of cancer with a novel dual-targeted conditionally replicative adenovirus armed with mda-7/IL-24 gene. Clin Cancer Res 2008, 14:2450-2457. 40. He X, Liu J, Yang C, Su C, Zhou C, Zhang Q, Li L, Wu H, Liu X, Wu M, Qian Q: 5/35 fiber-modified conditionally replicative adenovirus armed with p53 shows increased tumorsuppressing capacity to breast cancer cells. Hum Gene Ther 2011, 22:283-292. 41. Yao W, Guo G, Zhang Q, Fan L, Wu N, Bo Y: The application of multiple miRNA response elements enables oncolytic adenoviruses to possess specificity to glioma cells. Virology 2014, 458–459:69-82. 42. Callegari E, Elamin BK, D’Abundo L, Falzoni S, Donvito G, Moshiri F, Milazzo M, Altavilla G, Giacomelli L, Fornari F et al.: Antitumor activity of a miR-199-dependent oncolytic adenovirus. PLoS One 2013, 8:e73964. 43. Lee CY, Rennie PS, Jia WW: MicroRNA regulation of oncolytic herpes simplex virus-1 for selective killing of prostate cancer cells. Clin Cancer Res 2009, 15:5126-5135. 44. Mazzacurati L, Marzulli M, Reinhart B, Miyagawa Y, Uchida H, Goins WF, Li A, Kaur B, Caligiuri M, Cripe T et al.: Use of miRNA  response sequences to block off-target replication and increase the safety of an unattenuated, glioblastoma-targeted oncolytic HSV. Mol Ther 2015, 23:99-107. Current Opinion in Virology 2015, 13:40–48

48 Oncolytic viruses

Combined targeting strategies to improve the safety of an unattenuated oncolytic virus. 45. Barnes D, Kunitomi M, Vignuzzi M, Saksela K, Andino R: Harnessing endogenous miRNAs to control virus tissue tropism as a strategy for developing attenuated virus vaccines. Cell Host Microbe 2008, 4:239-248.

61. Brown BD, Naldini L: Exploiting and antagonizing microRNA regulation for therapeutic and experimental applications. Nat Rev Genet 2009, 10:578-585. 62. Pham AM, Langlois RA, TenOever BR: Replication in cells of hematopoietic origin is necessary for Dengue virus dissemination. PLoS Pathog 2012, 8:e1002465.

46. Edge RE, Falls TJ, Brown CW, Lichty BD, Atkins H, Bell JC: A let-7 MicroRNA-sensitive vesicular stomatitis virus demonstrates tumor-specific replication. Mol Ther 2008, 16:1437-1443.

63. Kelly EJ, Hadac EM, Cullen BR, Russell SJ: MicroRNA antagonism of the picornaviral life cycle: alternative mechanisms of interference. PLoS Pathog 2010, 6:e1000820.

47. Hikichi M, Kidokoro M, Haraguchi T, Iba H, Shida H, Tahara H, Nakamura T: MicroRNA regulation of glycoprotein B5R in oncolytic vaccinia virus reduces viral pathogenicity without impairing its antitumor efficacy. Mol Ther 2011, 19:1107-1115.

64. Gentner B, Schira G, Giustacchini A, Amendola M, Brown BD, Ponzoni M, Naldini L: Stable knockdown of microRNA in vivo by lentiviral vectors. Nat Methods 2009, 6:63-66.

48. Sugio K, Sakurai F, Katayama K, Tashiro K, Matsui H, Kawabata K, Kawase A, Iwaki M, Hayakawa T, Fujiwara T, Mizuguchi H: Enhanced safety profiles of the telomerase-specific replication-competent adenovirus by incorporation of normal cell-specific microRNA-targeted sequences. Clin Cancer Res 2011, 17:2807-2818. 49. Baertsch MA, Leber MF, Bossow S, Singh M, Engeland CE, Albert J, Grossardt C, Jager D, von Kalle C, Ungerechts G:  MicroRNA-mediated multi-tissue detargeting of oncolytic measles virus. Cancer Gene Ther 2014. Ability to modify multiple-tissue tropisms of a clinically relevant oncolytic virus through microRNA targeting. 50. Bofill-De Ros X, Gironella M, Fillat C: miR-148a- and miR-216aregulated oncolytic adenoviruses targeting pancreatic tumors attenuate tissue damage without perturbation of miRNA activity. Mol Ther 2014, 22:1665-1677. 51. Fu X, Rivera A, Tao L, De Geest B, Zhang X: Construction of an oncolytic herpes simplex virus that precisely targets hepatocellular carcinoma cells. Mol Ther 2012, 20:339-346. 52. Kelly EJ, Nace R, Barber GN, Russell SJ: Attenuation of vesicular stomatitis virus encephalitis through microRNA targeting. J Virol 2010, 84:1550-1562. 53. Varble A, Chua MA, Perez JT, Manicassamy B, Garcia-Sastre A, tenOever BR: Engineered RNA viral synthesis of microRNAs. Proc Natl Acad Sci U S A 2010, 107:11519-11524. 54. Perez JT, Pham AM, Lorini MH, Chua MA, Steel J, tenOever BR: MicroRNA-mediated species-specific attenuation of influenza A virus. Nat Biotechnol 2009, 27:572-576. 55. Langlois RA, Albrecht RA, Kimble B, Sutton T, Shapiro JS, Finch C, Angel M, Chua MA, Gonzalez-Reiche AS, Xu K et al.: MicroRNAbased strategy to mitigate the risk of gain-of-function influenza studies. Nat Biotechnol 2013, 31:844-847.

65. tenOever BR: RNA viruses and the host microRNA machinery.  Nat Rev Microbiol 2013, 11:169-180. In-depth review of the interactions between miRNAs, RNA viruses and the ability of RNA viruses to deliver miRNA-based therapeutics. 66. Bouchie A: First microRNA mimic enters clinic. Nat Biotechnol 2013, 31:577. 67. Kota J, Chivukula RR, O’Donnell KA, Wentzel EA, Montgomery CL, Hwang HW, Chang TC, Vivekanandan P, Torbenson M, Clark KR et al.: Therapeutic microRNA delivery suppresses tumorigenesis in a murine liver cancer model. Cell 2009, 137:1005-1017. 68. Trang P, Medina PP, Wiggins JF, Ruffino L, Kelnar K, Omotola M, Homer R, Brown D, Bader AG, Weidhaas JB, Slack FJ: Regression of murine lung tumors by the let-7 microRNA. Oncogene 2010, 29:1580-1587. 69. Miyazaki Y, Adachi H, Katsuno M, Minamiyama M, Jiang YM, Huang Z, Doi H, Matsumoto S, Kondo N, Iida M et al.: Viral delivery of miR-196a ameliorates the SBMA phenotype via the silencing of CELF2. Nat Med 2012, 18:1136-1141. 70. Lou W, Chen Q, Ma L, Liu J, Yang Z, Shen J, Cui Y, Bian XW,  Qian C: Oncolytic adenovirus co-expressing miRNA-34a and IL-24 induces superior antitumor activity in experimental tumor model. J Mol Med (Berl) 2013, 91:715-725. Combination of oncolytic virotherapy and a miRNA therapy currently in clinical trials. 71. Moshiri F, Callegari E, D’Abundo L, Corra F, Lupini L, Sabbioni S, Negrini M: Inhibiting the oncogenic mir-221 by microRNA sponge: toward microRNA-based therapeutics for hepatocellular carcinoma. Gastroenterol Hepatol Bed Bench 2014, 7:43-54. 72. Shapiro JS, Varble A, Pham AM, Tenoever BR: Noncanonical cytoplasmic processing of viral microRNAs. RNA 2010, 16:2068-2074.

56. Chua MA, Schmid S, Perez JT, Langlois RA, Tenoever BR: Influenza A virus utilizes suboptimal splicing to coordinate the timing of infection. Cell Rep 2013, 3:23-29.

73. Langlois RA, Shapiro JS, Pham AM, tenOever BR: In vivo delivery of cytoplasmic RNA virus-derived miRNAs. Mol Ther 2012,  20:367-375. Analysis of the mechanisms associated with pre-miRNAs expressed from cytoplasmic viruses.

57. Leber MF, Bossow S, Leonard VH, Zaoui K, Grossardt C, Frenzke M, Miest T, Sawall S, Cattaneo R, von Kalle C, Ungerechts G: MicroRNA-sensitive oncolytic measles viruses for cancer-specific vector tropism. Mol Ther 2011, 19:10971106.

74. Guo Z, Maki M, Ding R, Yang Y, Zhang B, Xiong L: Genome-wide survey of tissue-specific microRNA and transcription factor regulatory networks in 12 tissues. Sci Rep 2014, 4:5150.

58. Russell SJ, Federspiel MJ, Peng KW, Tong C, Dingli D, Morice WG,  Lowe V, O’Connor MK, Kyle RA, Leung N et al.: Remission of disseminated cancer after systemic oncolytic virotherapy. Mayo Clin Proc 2014, 89:926-933. First clinical demonstration of systemic oncolytic virotherapy for treatment of disseminated cancer.

76. Gentner B, Naldini L: Exploiting microRNA regulation for genetic engineering. Tissue Antigens 2012, 80:393-403.

59. Kueberuwa G, Cawood R, Tedcastle A, Seymour LW: Tissuespecific attenuation of oncolytic sindbis virus without compromised genetic stability. Hum Gene Ther Methods 2014, 25:154-165. 60. Yarbrough ML, Zhang K, Sakthivel R, Forst CV, Posner BA, Barber GN, White MA, Fontoura BM: Primate-specific miR-5763p sets host defense signalling threshold. Nat Commun 2014, 5:4963.

Current Opinion in Virology 2015, 13:40–48

75. Sakurai F, Katayama K, Mizuguchi H: MicroRNA-regulated transgene expression systems for gene therapy and virotherapy. Front Biosci (Landmark Ed) 2011, 16:2389-2401.

77. Di Leva G, Croce CM: The role of microRNAs in the tumorigenesis of ovarian cancer. Front Oncol 2013, 3:153. 78. Di Leva G, Garofalo M, Croce CM: MicroRNAs in cancer. Annu  Rev Pathol 2014, 9:287-314. Thorough review of miRNAs and the development of cancer following miRNA dysregulation. 79. Skalsky RL, Cullen BR: Reduced expression of brain-enriched microRNAs in glioblastomas permits targeted regulation of a cell death gene. PLoS One 2011, 6:e24248.

www.sciencedirect.com

MicroRNAs and oncolytic viruses.

MicroRNAs regulate gene expression in mammalian cells and often exhibit tissue-specific expression patterns. Incorporation of microRNA target sequence...
1MB Sizes 1 Downloads 12 Views