25
$FFHSWHG 0DQXVFULSW 5HYLHZ PLFUR51$V DV PHGLDWRUV RI LQVHFW KRVWSDWKRJHQ LQWHUDFWLRQV DQG LPPXQLW\ 0D]KDU +XVVDLQ 6DVVDQ $VJDUL 3,, 6 '2, KWWSG[GRLRUJMMLQVSK\V 5HIHUHQFH ,3 7R DSSHDU LQ Journal of Insect Physiology 5HFHLYHG 'DWH -XQH 5HYLVHG 'DWH $XJXVW $FFHSWHG 'DWH $XJXVW 3OHDVH FLWH WKLV DUWLFOH DV +XVVDLQ 0 $VJDUL 6 PLFUR51$V DV PHGLDWRUV RI LQVHFW KRVWSDWKRJHQ LQWHUDFWLRQV DQG LPPXQLW\ Journal of Insect Physiology GRL KWWSG[GRLRUJMMLQVSK\V 7KLV LV D 3') ILOH RI DQ XQHGLWHG PDQXVFULSW WKDW KDV EHHQ DFFHSWHG IRU SXEOLFDWLRQ $V D VHUYLFH WR RXU FXVWRPHUV ZH DUH SURYLGLQJ WKLV HDUO\ YHUVLRQ RI WKH PDQXVFULSW 7KH PDQXVFULSW ZLOO XQGHUJR FRS\HGLWLQJ W\SHVHWWLQJ DQG UHYLHZ RI WKH UHVXOWLQJ SURRI EHIRUH LW LV SXEOLVKHG LQ LWV ILQDO IRUP 3OHDVH QRWH WKDW GXULQJ WKH SURGXFWLRQ SURFHVV HUURUV PD\ EH GLVFRYHUHG ZKLFK FRXOG DIIHFW WKH FRQWHQW DQG DOO OHJDO GLVFODLPHUV WKDW DSSO\ WR WKH MRXUQDO SHUWDLQ

microRNAs as mediators of insect host-pathogen interactions ...337628/UQ337628...4 52 the transcription factor STAT. Upon activation of the pathway, STAT is phosphorylated and 53 translocated

  • Upload
    others

  • View
    1

  • Download
    0

Embed Size (px)

Citation preview

  • Journal of Insect Physiology

    Journal of Insect Physiology

    http://dx.doi.org/10.1016/j.jinsphys.2014.08.003http://dx.doi.org/http://dx.doi.org/10.1016/j.jinsphys.2014.08.003

  • Review article

    microRNAs as mediators of insect host-pathogen interactions and

    immunity

    Mazhar Hussain and Sassan Asgari*

    Australian Infectious Disease Research Centre, School of Biological Sciences, The

    University of Queensland, Brisbane, QLD 4072, Australia

    * To whom correspondence should be addressed: Asgari, S, Australian Infectious Disease

    Research Centre, School of Biological Sciences, The University of Queensland, St Lucia

    QLD 4072, Tel: (617) 3365-2043; Fax: (617) 3365-1655; Email: [email protected].

  • 2

    Abstract 1

    Insects are the most successful group of animals on earth, owing this partly to their very effective 2

    immune responses to microbial invasion. These responses mainly include cellular and humoral 3

    responses as well as RNA interference (RNAi). Small non-coding RNAs (snRNAs) produced 4

    through RNAi are important molecules in the regulation of gene expression in almost all living 5

    organisms; contributing to important processes such as development, differentiation, immunity as 6

    well as host-microorganism interactions. The main snRNAs produced by the RNAi response 7

    include short interfering RNAs, microRNAs and piwi-interacting RNAs. In addition to the host 8

    snRNAs, some microorganisms encode snRNAs that affect the dynamics of host-pathogen 9

    interactions. In this review, we will discuss the latest developments in regards to the role of 10

    microRNA in insect host-pathogen interactions and provide some insights into this rapidly 11

    developing area of research. 12

    1. Introduction 13

    1.1. Unfolding the multiple layers of insect responses to microbial infection 14

    Multiple mechanisms of attack and counter-attack or host defences have evolved during the natural 15

    process of the host-pathogen arms race. Animals defend themselves against microorganisms in two 16

    major ways: innate and/or acquired immune systems. The main difference between the two systems 17

    is the involvement of germline encoded factors for recognizing and destroying pathogens in innate 18

    immunity, while in acquired immunity effector molecules are produced to recognize antigens 19

    leading to the development of immunological memory. In the insect world, the immune system is 20

    based on innate immune responses that can be further categorised into three main parts: (1) 21

    pathogen immobilization, (2) core immune signalling pathways (Toll, Imd and JAK-STAT) and (3) 22

    the RNAi pathway. Pathogen immobilization is a cellular response towards fungal and bacterial 23

    infections in the form of phagocytosis, engulfment, melanisation, coagulation and autophagy 24

    (Dushay, 2009; Jiravanichpaisal et al., 2006). These responses help in immobilizing pathogens, 25

    which are then destroyed extracellularly by antimicrobial peptides (AMPs) or eliminated 26

  • 3

    intracellularly. In response to systemic infection, AMPs are produced and secreted in the 27

    haemolymph that directly attack microbes (Bulet et al., 1999). 28

    Upon pathogen invasion, activation of immune signalling pathways usually ensues. In insects, these 29

    are mainly the Toll, Imd and JAK/STAT pathways. The Toll pathway is activated through an 30

    endogenous ligand namely the Nerve Growth Factor-related cytokine Spätzle (Spz) (Weber et al., 31

    2003). Binding of Spz triggers a conformational change in the Toll receptor which activates the 32

    downstream signalling pathway (Gangloff et al., 2008). The Toll-Spz interaction is the convergence 33

    point for the activation of three pathogen recognition pathways; one triggered by fungal or bacterial 34

    proteases, one induced by recognition of fungal cell wall and one activated by Lysine-type bacterial 35

    peptidoglycan (El Chamy et al., 2008; Gottar et al., 2006; Ligoxygakis et al., 2002). Following 36

    receptor activation, MyD88 proteins interact with Toll through their respective Toll/Interleukin-1 37

    receptor domains and recruit the Drosophila melanogaster IRAK homologue, the kinase Pelle that 38

    phosphorylates Cactus for degradation. Upon Cactus degradation, the NF-κB homologues, Dorsal 39

    or Dif, are freed to move to the nucleus and regulate hundreds of target genes (Tauszig-Delamasure 40

    et al., 2002; Towb et al., 2001). 41

    The Immune deficiency (Imd) pathway is activated by meso-diaminopimelic acid (DAP)-type 42

    bacterial peptidoglycans that form the cell wall of Gram-negative bacteria and some Gram-positive 43

    Bacilli. Pathogen recognition in Imd occurs through the transmembrane receptor PGRP-LC and the 44

    intracellular receptor PGRP-LE that are critical for recognizing extracellular bacteria. Flies deficient 45

    in both of these proteins are unable to induce AMPs in response to Gram-negative bacteria and are 46

    therefore highly susceptible to these infections (Kaneko et al., 2006). 47

    The JAK/STAT signalling pathway has been shown to be involved in insect immunity, in particular 48

    anti-viral responses and hemocyte development (Agaisse and Perrimon, 2004; Dostert et al., 2005); 49

    although the mechanism of activation of the pathway by viruses is not known. The signal 50

    transduction pathway consists of unpaired 1-3 ligands, the receptor domeless, the kinase JAK and 51

  • 4

    the transcription factor STAT. Upon activation of the pathway, STAT is phosphorylated and 52

    translocated into the nucleus, thus, activating several target genes. 53

    1.2. RNA interference an effective anti-viral defence against viruses 54

    RNA interference (RNAi) is another arm of the multilayered defence system in insects involved in 55

    gene regulation, genome protection and anti-viral responses. Short interfering RNAs (siRNAs), 56

    microRNAs (miRNAs) and piwi-interacting RNAs (piRNAs) are the three main types of small non-57

    coding RNAs (snRNAs) central to the RNAi response pathway. Recently, a connection between the 58

    JAK/STAT signalling transduction and the RNAi pathway has been established in mosquitoes as 59

    part of the anti-viral immunity through the secretion of the signalling molecule, vago (Paradkar et 60

    al., 2012). This suggests the effective communications between different pathways associated with 61

    the innate immune responses in insects. Likewise, involvement of several miRNAs generated by 62

    host cells or encoded by viral genomes in host-pathogen interactions have recently opened up new 63

    research avenues in insect-microbe interactions. As the field of RNAi is rapidly expanding, in this 64

    review article, we will mainly focus on the role of miRNAs in insect host-pathogen crosstalk and 65

    immunity with the majority of examples from viruses as more relevant literature is available. For 66

    other types of small RNAs and their role in anti-viral responses, readers are referred to recent 67

    reviews (e.g. Sabin and Cherry, 2013; Bronkhorst and van Rij, 2014). 68

    2. miRNA biogenesis at a glance 69

    Since the discovery of the first miRNA from Caenorhabditis elegans just over two decades ago 70

    (Lee et al., 1993) our understanding of miRNA biogenesis and its interaction with targets has 71

    tremendously increased and is continuously evolving. In addition to the canonical pathway of 72

    miRNA biogenesis, several non-canonical pathways have been described. Based on the canonical 73

    pathway, a miRNA gene is transcribed in the nucleus by RNA polymerase II producing the primary 74

    miRNA (pri-miRNA) transcript that could be variable in length but contains one or several stem-75

    loop structures (Fig. 1). A nuclear residing ribonuclease, Drosha, in association with Pasha (in 76

    insects) cleaves the stem-loop producing a ∼70 nt hairpin precursor miRNA (pre-miRNA) that is 77

  • 5

    translocated into the cytoplasm by Exportin-5. In the cytoplasm, the hairpin head is cleaved by 78

    another ribonuclease enzyme, Dicer, producing a short duplex RNA molecule. In insects, two dicer 79

    proteins have been identified: Dicer-1 that is mainly involved in the miRNA biogenesis, and Dicer-80

    2 that produces siRNAs. The miRNA duplex triggers the formation of the RNA induced silencing 81

    complex (RISC) in which an Argonaut (Ago) protein is the main component. Most miRNAs are 82

    loaded into RISC that includes Ago1 and siRNAs into Ago2 complexes; however, recent research 83

    indicates that miRNAs can also be loaded into Ago2 complexes (Ghildiyal et al., 2010; Rubio and 84

    Belles, 2013), especially as the insect ages (Abe et al., 2014). Once the miRNA duplex becomes 85

    associated with the RISC, one of the strands is degraded (referred to as miRNA* strand) and the 86

    other one forms the mature miR-RISC complex. In some instances, the miRNA* strand is retained 87

    and loaded into Ago2 and can be functional (Ghildiyal et al., 2010). Finally, the mature miRNA 88

    guides the complex to target mRNA(s) in a sequence specific manner, leading to post-89

    transcriptional gene regulation, or the complex might be transported into the nucleus where the 90

    miRNA normally binds to the promoter regions of target genes regulating gene expression at the 91

    transcriptional level. Notably, each miRNA may have multiple mRNA targets and each mRNA may 92

    be targeted by multiple miRNAs. 93

    3. Role of microRNAs as moderators of host-pathogen interactions 94

    While there is a lot known about the role of miRNAs in insect development, less is understood 95

    about their function in host-pathogen interactions. Nevertheless, accumulating evidence in recent 96

    years has suggested their effects on host-microorganisms interactions and cross-kingdom 97

    communications. For example, several cellular as well as viral miRNAs have been found to be 98

    involved in assisting or limiting virus replication (see below). What is commonly observed in 99

    various host-pathogen systems is the differential abundance of host miRNAs following an infection, 100

    which varies at different stages of infection and type of pathogen involved. Concurrently, a large 101

    number of host genes are found down- or upregulated that could be regulated by miRNAs or other 102

    types of snRNAs. These changes in miRNA/mRNA profiles could be part of the host anti-pathogen 103

  • 6

    response or controlled by pathogen manipulation, including miRNAs expressed by viruses. Below, 104

    we will examine several relevant examples of the role miRNAs play in mediating host-pathogen 105

    interactions. 106

    3.1. Host-virus interactions 107

    3.1.1. Impact of viral infection on the host miRNA profile 108

    In recent years, based on deep sequencing or microarray analysis, several publications have 109

    appeared reporting changes in the expression levels of cellular miRNAs in insects in response to 110

    pathogen infection, although in the majority of these cases functional analyses are still lacking. For 111

    example, the abundance of a large number of miRNAs was changed in Spodoptera frugiperda (Sf9) 112

    cells following infection with Autographa californica multiple nucleopolyhedrovirus (AcMNPV). 113

    These included several upregulated miRNAs (e.g. miR-184, miR-998 and miR-10) at 24 h post-114

    infection (hpi), with expression of some of those downregulated at 72 hpi (Mehrabadi et al., 2013). 115

    In addition, three of the novel S. frugiperda miRNAs identified in the study showed upregulation 116

    upon virus infection. Relevant to baculoviruses, Helicoverpa armigera single nucleopolyhedrovirus 117

    (HaSNPV) was shown to cause changes in the abundance of host miRNAs (Jayachandran et al., 118

    2013). Among those, miR-14 that positively regulates H. armigera ecdysone receptor (Ha-EcR) in 119

    this insect was downregulated upon infection providing an example of virus-medicated 120

    manipulation of host physiology. 121

    In Bombyx mori larvae infected with B. mori cytoplasmic polyhedrosis virus (BmCPV), 58 122

    miRNAs were found significantly up or downregulated in the midgut as compared with that of non-123

    infected larvae at 72 and 96 hpi (Wu et al., 2013). Based on gene ontology (GO) analysis, the 124

    putative target genes of six of the differentially expressed miRNAs (bmo-miR-275, miR-14, miR-125

    1a, N-50, N-46 and N-45) were associated with response to stimulus and immune system process. 126

    The authors speculated that significant increases in miR-275 levels following BmCPV infection 127

    could explain the stunted growth phenotype in virus-infected B. mori, since the miRNA has been 128

    shown to regulate development during metamorphosis (Liu et al., 2010). In another study, 129

  • 7

    microarray analysis showed differential expression of at least 20 miRNAs in a Heliothis zea fat 130

    body (HzFB) cell line upon infection with Heliothis virescens ascovirus (HvAV3e) with three of 131

    them (Hz-miR-24, -22, -7) significantly downregulated (Hussain and Asgari, 2010). 132

    Arthropod-borne viruses (arboviruses) have also been demonstrated to alter the host miRNA 133

    profile. For example, in Culex quinquefasciatus infected with West Nile virus (WNV) 134

    downregulation of miR-989 and upregulation of miR-92 was noted (Skalsky et al., 2010). Deep 135

    sequencing analysis of Aedes aegypti mosquitoes infected with dengue virus (DENV) at 2, 4 and 9 136

    days post-infection revealed differential abundance of 35 miRNAs. Among those miR-5119-5p, 137

    miR-34-3p, miR-87-5p and miR-988-5p were upregulated (Campbell et al., 2013). As more of these 138

    studies emerge, it will become possible to carry out comparative studies to find out which miRNAs 139

    are commonly involved in host-pathogen responses in different host insects and whether particular 140

    miRNAs could be specifically regulated in response to particular groups of pathogens (e.g. viruses, 141

    bacteria, etc). 142

    3.1.2. Cellular miRNAs target viral genes 143

    Given alteration of the abundance of host miRNAs following infection, it is likely that some of 144

    those could directly target viral genes. For instance, in H. zea HzFB cells, miR-24 downregulates 145

    DNA dependent RNA polymerase II RPC2 (DdRP; ORF64) and DdRP β subunits (DdRP; ORF82) 146

    of HvAV3e (Hussain and Asgari, 2010). Interestingly, this miRNA is selectively downregulated 147

    early in infection allowing transcription of viral genes required for virus replication suggesting that 148

    the virus may actively suppress host miRNAs that directly target its genes. Relevant to this, in B. 149

    mori, several targets of the cellular bmo-miR-8 were predicted in Bombyx mori 150

    nucleopolyhedrovirus (BmNPV) genes (see below). Inhibition of this miRNA led to 8-fold increase 151

    in viral titre in the fat bodies of infected larvae (Singh et al., 2012). 152

    A recent paper highlighted the possibility of utilizing cellular miRNAs as a tool to target virus 153

    genes as a control strategy (Zhang et al., 2014a). This work was conducted in B. mori against 154

    BmNPV using three cellular miRNAs; bmo-miR-279, bmo-miR-92b, and bmo-miR-2764. 155

  • 8

    Accordingly, the cellular pre-miRNA sequences were used as backbone to replace the natural 156

    miRNA duplex sequences with artificial siRNA duplex sequences targeting the viral lef-11 gene, 157

    which is involved in virus replication (Zhang et al., 2014a). For this, a virus inducible artificial 158

    miRNA (amiRNAs) expression cassette was constructed targeting the viral gene. Infection with the 159

    virus stimulates the baculovirus-induced promoter 39K included in the expression cassette leading 160

    to production of pre-amiRNA and eventually formation of siRNA duplexes. The mature amiRNAs 161

    successfully inhibited BmNPV proliferation by downregulating the lef-11 gene in silkworm BmN-162

    SWU1 cells. It will be interesting to see if the approach could be utilized in whole insects to inhibit 163

    BmNPV replication without compromising the insect’s fitness. 164

    Viruses might also subvert host miRNAs to their own benefit. For example, in American eastern 165

    equine encephalitis virus, binding of the mammalian host miR-142-3p to the 3’UTR of the virus 166

    genome in myeloid-lineage cells restricts replication of the virus in this particular cell type leading 167

    to minimal interferon induction and hence suppression of the host innate immune system (Trobaugh 168

    et al., 2013). Interestingly, this study showed that these same binding sites in the virus 3’UTR are 169

    also required for efficient infection of the mosquito vector Aedes taeniorhynchus. However, it is not 170

    known whether a mosquito miRNA might interact with the sequences facilitating its replication in 171

    the insect host similar to the mammalian host. 172

    3.1.3 Virus-encoded miRNAs can target host genes 173

    Recent investigations have revealed that insect virus miRNAs, similar to those reported from human 174

    viruses, might target and regulate host genes to establish infection and successfully replicate. 175

    However, the information is very limited in respect to insect viruses. The impact of the miRNA 176

    might be via regulation of a specific gene leading to the modification of a particular pathway or 177

    might cause a global reduction in the host miRNA depending on the target genes. For example, 178

    baculovirus encoded BmNPV-miR-1 downregulates the nuclear Exportin-5 cofactor Ran resulting 179

    in the blockage of the export of cellular pre-miRNAs into the cytoplasm generally affecting the 180

    abundance of cellular mature miRNAs in infected cells (Singh et al., 2012) (Table 1). This may 181

  • 9

    affect regulation of cellular as well as viral genes. For instance, predicted targets of four selected 182

    host miRNAs affected by downregulation of Ran via BmNPV-miR-1 were mostly related to 183

    immunity. In addition, targets of these miRNAs were also predicted in the viral genome. For 184

    example, bmo-miR-8, which is downregulated following infection, targets the immediate early-1 185

    (ie1) gene (Singh et al., 2012). Downregulation of the miRNA is obviously to the benefit of the 186

    virus. However, it remains to be elucidated how viral miRNAs are produced while the host miRNA 187

    biogenesis pathway is interrupted. 188

    In addition, a WNV (Kunjin strain, WNVKUN) encoded miRNA, KUN-miR-1, derived from the 189

    3’SL stem-loop located at the end of the 3’UTR of the virus in mosquito cells positively regulates 190

    the host GATA-binding protein 4 (GATA-4) (Hussain et al., 2012) (Table 1). RNAi-mediated 191

    silencing of GATA-4 led to a significant reduction in virus RNA replication, suggesting that 192

    GATA-4-induction via KUN-miR-1 plays an important role in virus replication. However, the 193

    mechanism remains to be explored. Members of the GATA family are transcription factors that 194

    regulate various biological processes in insects and vertebrates (Shapira et al., 2006). In particular, 195

    GATA-4 has been shown to be involved in regulation of egg development, vitellogenesis, and lipid 196

    metabolism in mosquitoes (Cheon et al., 2006; Park et al., 2006). 197

    3.1.4. Virus-encoded miRNAs target viral own genes 198

    Targeting virus genes by virus-encoded miRNAs appears to be common among different groups of 199

    viruses to regulate their own replication, enter latency or suppress host anti-viral responses. The 200

    first insect virus-encoded miRNA, HvAV-miR-1, was reported from the ascovirus HvAV-3e 201

    originating from the major capsid protein (MCP) gene of the virus (Hussain et al., 2008). The virus-202

    encoded miRNA was shown to autoregulate virus replication late in infection by downregulating 203

    the virus’s own DNA polymerase (Table 1). In a recent report, it was revealed that BmNPV 204

    expresses a miRNA, bmnpv-miR-3, early in infection, which targets and regulates expression of the 205

    viral DNA binding protein (P6.9) as well as other late genes (Singh et al., 2014) (Table 1). This may 206

    confer several benefits to the virus: (1) autoregulation of virus proliferation to avoid over 207

  • 10

    replication, (2) regulation of the expression levels of the late genes to suppress their expression 208

    early in infection, (3) modulation of the host immune response by regulating virus proliferation. In 209

    regards to the latter, the investigators showed that virus load correlates with expression levels of the 210

    AMP gloverin, which has been shown to be upregulated upon baculovirus infection and inactivates 211

    budded viruses (Moreno-Habel et al., 2012). 212

    A miRNA from the baculovirus AcMNPV, AcMNPV-miR-1, was found to be expressed from the 213

    negative strand of the viral genome and target ODV-E25 transcripts on the opposite strand with 214

    100% complementarity (Zhu et al., 2013) (Table 1). ODV-E25 is a late gene encoding a ∼25 kDa 215

    structural protein of occlusion derived virus (ODV) and also found on the budded virus (BV) 216

    envelope (Russell and Rohrmann, 1993; Wang et al., 2010). The protein is conserved in 217

    baculoviruses that infect lepidopterans and important in ODV formation and BV infectivity (Chen 218

    et al., 2012). AcMNPV-miR-1 was expressed late in infection starting from 12 hpi. Bioinformatics 219

    analysis identified eight potential targets for AcMNPV-miR-1 in the virus genome with ODV-E25 220

    showing 100% complementarity since it is expressed on the complementary strand of the miRNA. 221

    Validation experiments confirmed that AcMNPV-miR-1 causes cleavage of ODV-E25 mRNA and 222

    its higher expression levels coincided with reduced levels of the target mRNA levels. Notably, 223

    using AcMNPV-miR-1 mimic, it was found that the small RNA had no effect on budded virus 224

    production but affected their infectivity by reducing production of ODV-E25 (Chen et al., 2012). 225

    Furthermore, the mimic increased ODV formation, and inhibition of AcMNPV-miR-1 dramatically 226

    reduced the number of occluded ODVs. This implied that the miRNA promotes ODV formation. 227

    In another example, a functional viral small RNA (DENV-vsRNA-5) produced from a stem-loop of 228

    DENV-2 3’UTR was found in virus-infected Ae. aegypti mosquitoes and cell lines. The vsRNA was 229

    shown to target the virus genome at NS1 gene region (Hussain and Asgari, 2014) (Table 1). 230

    Synthetic RNA inhibitor or mimic of DENV-vsRNA-5 resulted in significant increases or decrease 231

    in viral replication, respectively. The autoregulation of virus replication may facilitate low but 232

    transmissible level of virus infection in the mosquito vector. 233

  • 11

    In a number of vertebrate herpesviruses, it has been shown that virus-encoded miRNAs are 234

    important in switching between lytic and latent phases (reviewed in Cullen, 2011). In insect viruses, 235

    Heliothis zea nudivirus (HzNV-1) expresses only one non-protein coding gene, the persistency-236

    associated gene 1 (pag1), during the latency phase (Wu et al., 2011) (Table 1). Two miRNAs 237

    encoded by pag1 were shown to downregulate the viral early gene hhi1, which allows the virus to 238

    enter the latent infection stage in cell culture. In the absence of the pag1 gene, the virus does not 239

    enter latency phase. It will be interesting to find out whether virus-encoded miRNAs play any role 240

    in latency or persistent infection observed in other insect viruses. 241

    3.2. Bacterial infection can modulate host miRNAs in insects 242

    A number of studies have revealed that similar to viruses bacterial infections can lead to changes in 243

    cellular miRNAs in different insect species, with information again being limited. For example, 244

    association of the endosymbiotic bacterium Wolbachia (wMelPop strain) with alterations in cellular 245

    miRNAs and miRNA pathway proteins in Ae. aegypti has been documented (Hussain et al., 2011; 246

    Hussain et al., 2013a; Mayoral et al., 2014; Osei-Amo et al., 2012; Zhang et al., 2013). Of note, this 247

    strain causes life-shortening in the mosquito as well as inhibit replication of several vector-borne 248

    pathogens especially RNA viruses such as DENV, WNV and Chikungunya virus (Moreira et al., 249

    2009a; Moreira et al., 2009b). Initially, microarray analysis revealed that a number of mosquito 250

    miRNAs were differentially expressed with two of them, aae-miR-2940 and -309a, showing high 251

    abundance in Wolbachia-infected mosquitoes as compared with non-infected ones (Hussain et al., 252

    2011). A target sequence of aae-miR-2940 was shown to reside in the 3’UTR of the mosquito’s 253

    metalloprotease m41 ftsh transcripts; however, the miRNA was shown to positively regulate the 254

    metalloprotease transcript levels. miRNA inhibition and silencing of the target gene revealed that 255

    the metalloprotease gene and aae-miR-2940 are important in the maintenance of Wolbachia density 256

    (Hussain et al., 2011). In a follow-up study, aae-miR-2940 was found to negatively regulate the 257

    transcript levels of its other target, the DNA methyltransferase (AaDnmt2) (Zhang et al., 2013). 258

    Overexpression of AaDnmt2 in Wolbachia-infected mosquito cells led to significant reductions in 259

  • 12

    Wolbachia density, but significantly enhanced replication of DENV. This suggested that 260

    downregulation of AaDnmt2 by aae-miR-2940 may contribute to inhibition of DENV replication in 261

    Wolbachia-infected mosquitoes. Notably, significant hypomethylation of Ae. aegypti genome was 262

    recognized in wMelPop infected mosquitoes (Ye et al., 2013). Given that AaDnmt2 is the only 263

    DNA methyltransferase gene in the mosquito’s genome, Wolbachia may epigenetically regulate 264

    expression of many host genes by suppression of AaDnmt2 via aae-miR-2940 induction. Further, 265

    the instances above provide an example of a miRNA regulating more than one gene in opposite 266

    ways, both benefiting the endosymbiotic bacterium. 267

    Function of another Wolbachia induced cellular miRNA, aae-miR-12, was confirmed in regulation 268

    of two target genes, DNA replication licensing factor MCM6 and monocarboxylate transporter 269

    MCT1, and shown to be important for the maintenance of Wolbachia density in mosquito cells 270

    (Osei-Amo et al., 2012). Recently, it was uncovered that Wolbachia blocks shuttling of Ago1 to the 271

    nucleus through upregulation of aae-miR-981 (Hussain et al., 2013a) and that leads to changes in 272

    the distribution of miRNAs in infected cells (Mayoral et al., 2014). Interestingly, this miRNA 273

    downregulates expression of importin β-4 that facilitates Ago1 distribution to the nucleus (Hussain 274

    et al., 2013a). 275

    A study on honeybees (Apis mellifera) showed activation of AMPs and regulation of a large number 276

    of immune-related genes in response to infection by the Gram-negative bacterium Serratia 277

    marcescens (Lourenco et al., 2013). Analysis of miRNAs from the infected bees also revealed 278

    differential expression of cellular miRNAs upon bacterial infection. Among the downregulated 279

    miRNAs were ame-bantam, miR-12, -34, -184, -278, -375, -989 and -1175, while some were 280

    upregulated such as let-7, miR-2, -13a, and -92a. Predicted interactions revealed that the miRNA 281

    could potentially regulate genes involved in Imd, Toll, JNK, and JAK/STAT pathways, regulation 282

    of AMPs and melanization. However, functional analysis and validation of these interactions 283

    remain to be investigated. 284

    285

  • 13

    286

    4. Role of miRNAs in regulation of insect immunity 287

    Despite many reports demonstrating the role of miRNAs in vertebrate immunity, the information 288

    pertinent to their role in insect immunity is scant and mostly based on bioinformatics predictions or 289

    correlating miRNA abundance with that of mRNAs. For example, Fullaondo et al. identified over 290

    70 miRNAs that could be involved in regulating immune related pathways such as Toll, Imd and 291

    JAK/STAT, melanization and JNK pathways using bioinformatics approaches (Fullaond and Lee, 292

    2012). Similarly, a computational approach was used to identify Anopheles gambiae miRNAs that 293

    may be involved in regulating immune responses. Two miRNAs, aga-miR-2304 and aga-miR-2390, 294

    were predicted to target the genes encoding suppressin and prophenoloxidase 295

    (Thirugnanasambantham et al., 2013). However, these predictions await experimental confirmation 296

    of miRNA-target interactions. In the flour beetle Tribolium castaneum, microbial challenge and 297

    stress led to alteration of the beetle’s miRNA profile (Freitak et al., 2012). Induction of immunity 298

    by injection of peptidoglycan into the beetles resulted in differential expression of 59 miRNAs from 299

    which 21 were upregulated. Similarly, changes in the abundance of Manduca sexta miRNAs were 300

    observed following immune challenge involved in pattern recognition, activation of 301

    prophenoloxidase, synthesis of AMPs and conserved immune signal transduction pathways (Zhang 302

    et al., 2014b). In An. gambiae, the possibility of involvement of miRNAs in defence against 303

    Plasmodium berghei invasion was demonstrated as Dicer-1 and Ago1 depletion in the mosquitoes 304

    led to increased sensitivity of the mosquitoes to P. berghei infection (Winter et al., 2007). However, 305

    no specific miRNA was identified responsible for the effect observed. 306

    One of the miRNAs whose role in insect immunity has experimentally been established is miR-8 307

    which negatively regulates expression of the AMPs, such as Drosomycin and Diptericin, in D. 308

    melanogaster (Choi and Hyun, 2012). This allows maintenance of low level of AMPs expression 309

    under non-infection conditions, enabling the homeostasis of immunity. However, in miR-8 null 310

    Drosophila, the transcript levels of the AMPs significantly increased; although it was not shown 311

  • 14

    how miR-8 regulates AMPs. It was predicted that the miRNA could potentially target the transcripts 312

    of Gram-negative binding protein 3 (GNBP3), an inducer of the Toll pathway upon fungal 313

    infection, and Pvf, a modulator of the JNK pathway. Negative regulation of AMPs by miR-8 was 314

    also demonstrated in the larvae of Plutella xylostella (Etebari and Asgari, 2013). This study further 315

    showed that miR-8 positively regulates the transcript levels of serpin-27, inhibiting activation of the 316

    Toll pathway and hence low level of production of AMPs when there is no infection/parasitization. 317

    Upon infection, miR-8 levels are reduced leading to a decrease in serpin-27 levels and consequently 318

    activation of the Toll pathway. 319

    In Aedes albopictus (C6/36) cells, the abundance of the mosquito-specific aae-miR-2940 was 320

    shown to be selectively reduced following WNVKUN infection, a flavivirus (Slonchak et al., 2014). 321

    As indicated above, aae-miR-2940 positively regulates the metalloprotease ftsh in mosquitoes 322

    (Hussain et al., 2011), and this protein was found to enhance WNVKUN replication (Slonchak et al., 323

    2014). Reduced aae-miR-2940 levels led to lower metalloprotease levels and consequently reduced 324

    virus titers. This suggested that reduction in the miRNA levels might be an anti-viral response to 325

    WNV infection. Similarly, reduced levels of aae-miR-2940 were detected in Ae. albopictus Singh’s 326

    cell line infected with Chikungunya virus, an alphavirus (Shrinet et al., 2014). These results suggest 327

    that reduced levels of aae-miR-2940 might be a common response to viral infection. 328

    In Ae. aegypti, two immune related genes, cactus and REL1, were found to be the target of a blood 329

    meal induced miRNA, aae-miR-375 (Hussain et al., 2013b). Interestingly, cactus, which is an 330

    inhibitor of the Toll pathway, was positively regulated by aae-miR-375, whereas REL1, which is an 331

    activator of AMPs, was downregulated by the same miRNA. The increase in cactus and decrease in 332

    REL1 was found to have a positive effect on DENV replication since AMPs have been shown to 333

    negatively affect DENV (Luplertlop et al., 2012; Xia et al., 2008). Coincidence of negative 334

    regulation of AMPs upon blood feeding may provide a fitness advantage for the virus. 335

    Finally, it is well-known that immune related genes evolve faster than non-immune genes due to 336

    constant evolutionarily pressures on hosts to neutralize pathogen novelties. Interestingly, when 337

  • 15

    analysing 50 genes involved in the Drosophila Toll and Imd signalling pathways, Han et al realized 338

    a strong negative correlation between the rate of evolution of the immune proteins and the number 339

    of their regulatory miRNAs, suggesting that genes regulated by more miRNAs evolve slower due to 340

    stronger functional constraints (Han et al., 2013). This suggests the role of miRNAs in the evolution 341

    of immune related genes in insects. 342

    5. Conclusions 343

    In the last decade, we have witnessed major developments in our understanding of the biogenesis 344

    and the function of miRNAs in various aspects of insect biology. Most of these efforts have focused 345

    on the role of miRNAs in development with the main focus on D. melanogaster. Recently, there 346

    have been a number of publications on the role of these small RNAs in host-pathogen interactions. 347

    As an initial step towards understanding their role in the interactions, the majority of investigators 348

    have focused on the effect of infection on the host miRNA profile with several being identified as 349

    differentially abundant miRNAs. It is of great interest to see these efforts to be followed up by 350

    empirical approaches to demonstrate the functional role of these differentially abundant miRNAs in 351

    host-pathogen interactions and insect immunity. In addition, a number of virus-encoded miRNAs 352

    have been characterized that either regulate virus replication or modulate host genes/miRNAs to 353

    facilitate their replication. 354

    Extending research on miRNAs beyond scientific curiosity and basic research to utilize them in 355

    applied insect biology with the aim to reduce crop damage by agricultural pests, inhibit mortality in 356

    beneficial insects and inhibit/limit transmission of vector-borne pathogens should be considered as 357

    future prospects. However, these efforts may encounter regulatory limitations when their 358

    applications are to extend to the field arena since genetic manipulations of insects or plants are most 359

    likely inevitable. Some attempts have already been made in this area with promising results either in 360

    cell culture or at the laboratory scale. miRNA research is certainly a very fast moving area of 361

    research and we shall see a strong progress in this field in the near future. 362

    363

  • 16

    364

    Acknowledgements 365

    The authors would like to acknowledge Dr Karyn Johnson from the University of Queensland for 366

    critically reading this review. The authors would like to acknowledge the Australian Research 367

    Council (DP110102112, DE120101512) and National Health and Medical Research (APP1062983, 368

    APP1027110) for providing funding that contributed to some of the findings reported in this review 369

    from our lab. 370

    References 371

    Abe, M., Naqvi, A., Hendriks, G., Feltzin, V., Zhu, Y., Grigoriev, A., Bonini, N., 2014. Impact of 372 age-associated increase in 2'-O-methylation of miRNAs on aging and neurodegeneration in 373 Drosophila. Genes Dev. 28, 44-57. 374

    Agaisse, H., Perrimon, N., 2004. The roles of JAK/STAT signaling in Drosophila immune 375 responses. Immunol. Rev. 198, 72-82. 376

    Bronkhorst, A.W., van Rij R.P., 2014. The long and short of antiviral defense: small RNA-based 377 immunity in insects. Curr. Opi. Virol. 7C:19-28. 378

    Bulet, P., Hetru, C., Dimarcq, J.L., Hoffmann, D., 1999. Antimicrobial peptides in insects; structure 379 and function. Dev. Comp. Immunol. 23, 329–344. 380

    Campbell, C.L., Harrison, T., Hess, A.M., Ebel, G.D., 2013. MicroRNA levels are modulated in 381 Aedes aegypti after exposure to Dengue-2. Insect Mol. Biol. 23, 132-139. 382

    Chen, L., Hu, X., Xiang, X., Yu, S., Yang, R., Wu, X., 2012. Autographa californica multiple 383 nucleopolyhedrovirus odv-e25 (Ac94) is required for budded virus infectivity and occlusion-384 derived virus formation. Arch. Virol. 157, 617-625. 385

    Cheon, H.M., Shin, S.W., Bian, G., Park, J.H., Raikhel, A.S., 2006. Regulation of lipid metabolism 386 genes, lipid carrier protein lipophorin, and its receptor during immune challenge in the 387 mosquito Aedes aegypti. J. Biol. Chem. 281, 8426-8435. 388

    Choi, I.K., Hyun, S., 2012. Conserved microRNA miR-8 in fat body regulates innate immune 389 homeostasis in Drosophila. Dev. Comp. Immunol. 37, 50-54. 390

    Cullen, B.R., 2011. Herpesvirus microRNAs: phenotypes and functions. Curr. Opin. Virol. 1, 211-391 215. 392

    Dostert, C., Jouanguy, E., Irving, P., Troxler, L., Galiana-Arnoux, D., Hetru, C., Hoffmann, J.A., 393 Imler, J.L., 2005. The Jak–STAT signaling pathway is required but not sufficient for the 394 antiviral response of Drosophila. Nat. Immunol. 6, 946–953. 395

    Dushay, M.S., 2009. Insect hemolymph clotting. Cell Mol Life Sci 66, 2643-2650. 396 El Chamy, L., Leclerc, V., Caldelari, I., Reichhart, J.M., 2008. Sensing of ‘danger signals’ and 397

    pathogen associated molecular patterns defines binary signaling pathways ’upstream’ of Toll. 398 Nat. Immunol. 9, 1165-1170. 399

    Etebari, K., Asgari, S., 2013. Conserved microRNA miR-8 blocks activation of the Toll pathway by 400 up-regulating Serpin 27 transcripts. RNA Biol. 10, 1356-1364. 401

    Freitak, D., Knorr, E., Vogel, H., Vilcinskas, A., 2012. Gender- and stressor-specific microRNA 402 expression in Tribolium castaneum. Biol. Lett. 8, 860-863. 403

    Fullaond, A., Lee, S.Y., 2012. Identification of putative miRNA involved in Drosophila 404 melanogaster immune response. Dev. Comp. Immunol. 36, 267-273. 405

  • 17

    Gangloff, M., Murali, A., Xiong, J., Arnot, C.J., Weber, A.N., Sandercock, A.M., Robinson, C.V., 406 Sarisky, R., Holzenburg, A., Kao, C., Gay, N.J., 2008. Structural insight into the mechanism of 407 activation of the Toll receptor by the dimeric ligand Spätzle. J. Biol. Chem. 283, 14629–14635. 408

    Ghildiyal, M., Xu, J., Seitz, H., Weng, Z., Zamore, P.D., 2010. Sorting of Drosophila small 409 silencing RNAs partitions microRNA* strands into the RNA interference pathway. RNA 16, 410 43-56. 411

    Gottar, M., Gobert, V., Matskevich, A.A., Reichhart, J.M., Wang, C., Butt, T.M., Belvin, M., 412 Hoffmann, J.A., Ferrandon, D., 2006. Dual detection of fungal infections in Drosophila via 413 recognition of glucans and sensing of virulence factors. Cell 127, 1425-1437. 414

    Han, M., Qin, S., Song, X., Li, Y., Jin, P., Chen, L., Ma, F., 2013. Evolutionary rate patterns of 415 genes involved in the Drosophila Toll and Imd signaling pathway. BMC Evol Biol 13, 245. 416

    Hussain, M., Asgari, S., 2010. Functional analysis of a cellular microRNA in insect host-ascovirus 417 interaction. J. Virol. 84, 612-620. 418

    Hussain, M., Asgari, S., 2014. MicroRNA-like viral small RNA from Dengue virus 2 autoregulates 419 its replication in mosquito cells. Proc. Natl. Acad. Sci. USA 111, 2746–2751. 420

    Hussain, M., Frentiu, F., Moreira, L., O’Neill, S., Asgari, S., 2011. Wolbachia utilizes host 421 microRNAs to manipulate host gene expression and facilitate colonization of the dengue vector 422 Aedes aegypti. Proc. Natl. Acad. Sci. USA 108, 9250-9255. 423

    Hussain, M., O'Neill, S.L., Asgari, S., 2013a. Wolbachia interferes with the intracellular 424 distribution of Argonaute 1 in the dengue vector Aedes aegypti by manipulating the host 425 microRNAs. RNA Biol. 10, 1868-1875. 426

    Hussain, M., Taft, R.J., Asgari, S., 2008. An insect virus-encoded microRNA regulates viral 427 replication. J. Virol. 82, 9164–9170. 428

    Hussain, M., Torres, S., Schnettler, E., Funk, A., Grundhoff, A., Pijlman, G.P., Khromykh, A.A., 429 Asgari, S., 2012. West Nile virus encodes a microRNA-like small RNA in the 3' untranslated 430 region which up-regulates GATA4 mRNA and facilitates virus replication in mosquito cells 431 Nucleic Acids Res. 40, 2210-2223. 432

    Hussain, M., Walker, T., O’Neill, S., Asgari, S., 2013b. Blood meal induced microRNA regulates 433 development and immune associated genes in the Dengue mosquito vector, Aedes aegypti. 434 Insect Biochem. Mol. Biol. 43, 146-152. 435

    Jayachandran, B., Hussain, M., Asgari, S., 2013. Regulation of Helicoverpa armigera ecdysone 436 receptor by miR-14 and its potential link to baculovirus infection. J. Invertebr. Pathol. 114, 437 151-157. 438

    Jiravanichpaisal, P., Lee, B.L., Soderhall, K., 2006. Cell-mediated immunity in arthropods: 439 hematopoiesis, coagulation, melanization and opsonization. Immunobiology 211, 213-236. 440

    Kaneko, T., Yano, T., Aggarwal, K., Lim, J.H., Ueda, K., Oshima, Y., Peach, C., Erturk-Hasdemir, 441 D., Goldman, W.E., Oh, B.H., Kurata, S., Silverman, N., 2006. PGRP-LC and PGRP-LE have 442 essential yet distinct functions in the Drosophila immune response to monomeric DAP-type 443 peptidoglycan. Nat. Immunol. 7, 715-723. 444

    Lee, R.C., Feinbaum, R.L., Ambros, V., 1993. The C. elegans heterochronic gene lin-4 encodes 445 small RNAs with antisense complimentarity to lin-14. Cell 75, 843-854. 446

    Ligoxygakis, P., Pelte, N., Hoffmann, J.A., Reichhart, J.M., 2002. Activation of Drosophila Toll 447 during fungal infection by a blood serine protease. Science 297, 114-116. 448

    Liu, S., Gao, S., Zhang, D., Yin, J., Xiang, Z., Xia, Q., 2010. MicroRNAs show diverse and 449 dynamic expression patterns in multiple tissues of Bombyx mori. BMC Genomics 11, 85. 450

    Lourenco, A.P., Guidugli-Lazzarini, K.R., Freitas, F.C.P., Bitondi, M.M.G., Simoes, Z.L.P., 2013. 451 Bacterial infection activates the immune system response and dysregulates microRNA 452 expression in honey bees. Insect Biochem. Mol. Biol. 43, 474-482. 453

    Luplertlop, N., Surasombatpattana, P., Patramool, S., Dumas, E., Wasinpiyamongkol, L., Saune, L., 454 Hamel, R., Bernard, E., Sereno, D., Thomas, F., Piquemal, D., Yssel, H., Briant, L., Misse, D., 455 2012. Induction of a peptide with activity against a broad spectrum of pathogens in the Aedes 456 aegypti salivary gland, following infection with dengue virus. PLoS Pathog. 7, e1001252. 457

  • 18

    Mayoral, J.G., Etebari, K., Hussain, M., Khromykh, A.A., Asgari, S., 2014. Wolbachia infection 458 modifies the profile, shuttling and structure of microRNAs in a mosquito cell line. PLoS ONE 459 9, e96107. 460

    Mehrabadi, M., Hussain, M., Asgari, S., 2013. MicroRNAome of Spodoptera frugiperda cells (Sf9) 461 and its alteration following baculovirus infection. J. Gen. Virol. 94, 1385–1397. 462

    Moreira, L.A., Iturbe-Ormaetxe, I., Jeffery, J.A., Lu, G.J., Pyke, A.T., Hedges, L.M., Rocha, B.C., 463 Hall-Mendelin, S., Day, A., Riegler, M., Hugo, L.E., Johnson, K.N., Kay, B.H., McGraw, E.A., 464 van den Hurk, A.F., Ryan, P.A., O'Neill, S.L., 2009a. A Wolbachia symbiont in Aedes aegypti 465 limits infection with Dengue, Chikungunya, and Plasmodium. Cell 139, 1268-1278. 466

    Moreira, L.A., Saig, E., Turley, A.P., Ribeiro, J.M.C., O'Neill, S.L., McGraw, E.A., 2009b. Human 467 probing behavior of Aedes aegypti when infected with a life-shortening strain of Wolbachia. 468 Plos Neglect. Trop. Dis. 3, e568. 469

    Moreno-Habel, D., Biglang-awa, I., Dulce, A., Luu, D., Garcia, P., Weers, P., Haas-Stapleton, E., 470 2012. Inactivation of the budded virus of Autographa californica M nucleopolyhedrovirus by 471 gloverin. J. Invertebr. Pathol. 110, 92-101. 472

    Osei-Amo, S., Hussain, M., O’Neill, S.L., Asgari, S., 2012. Wolbachia-induced aae-miR-12 473 miRNA negatively regulated the expression of MCT1 and MCM6 genes in Wolbachia-infected 474 mosquito cell line. PLoS ONE 7, e50049. 475

    Paradkar, P.N., Trinidad, L., Voysey, R., Duchemin, J.B., Walker, P.J., 2012. Secreted Vago 476 restricts West Nile virus infection in Culex mosquito cells by activating the Jak-STAT pathway. 477 Proc. Natl. Acad. Sci. USA 109, 18915–18920. 478

    Park, J.H., Attardo, G.M., Hansen, I.A., Raikhel, A.S., 2006. GATA factor translation is the final 479 downstream step in the amino acid/target-of-rapamycin-mediated vitellogenin gene expression 480 in the anautogenous mosquito Aedes aegypti. J. Biol. Chem. 281, 11167-11176. 481

    Rubio, M., Belles, X., 2013. Subtle roles of microRNAs let-7, miR-100 and miR-125 on wing 482 morphogenesis in hemimetabolan metamorphosis. J. Insect Physiol. 59, 1089-1094. 483

    Russell, R., Rohrmann, G., 1993. A 25-kDa protein is associated with the envelopes of occluded 484 baculovirus virions. Virology 195, 532–540. 485

    Sabin, L.R., Cherry, S., 2013. Small creatures use small RNAs to direct antiviral defenses. Eur. J. 486 Immunol. 43, 27-33. 487

    Shapira, M., Hamlin, B.J., Rong, J., Chen, K., Ronen, M., Tan, M.W., 2006. A conserved role for a 488 GATA transcription factor in regulating epithelial innate immune responses. Proc. Natl. Acad. 489 Sci. USA 103, 14086-14091. 490

    Shrinet, J., Jain, S., Jain, J., Bhatnagar, R., S, S., 2014. Next generation sequencing reveals 491 regulation of distinct Aedes microRNAs during chikungunya virus development. PLoS Negl. 492 Trop. Dis. 8, e2616. 493

    Singh, C.P., Singh, J., Nagaraju, J., 2012. A baculovirus-encoded microRNA (miRNA) suppresses 494 its host miRNA biogenesis by regulating the exportin-5 cofactor Ran. J. Virol. 86, 7867–7879. 495

    Singh, C.P., Singh, J., Nagaraju, J., 2014. bmnpv-miR-3 facilitates BmNPV infection 1 by 496 modulating the 2 expression of viral P6.9 and other late genes in Bombyx mori. Insect Biochem. 497 Mol. Biol. 49, 59-69. 498

    Skalsky, R.L., Vanlandingham, D.L., Scholle, F., Higgs, S., Cullen, B.R., 2010. Identification of 499 microRNAs expressed in two mosquito vectors, Aedes albopictus and Culex quinquefasciatus. 500 BMC Genomics 11, 119. 501

    Slonchak, A., Hussain, M., Morales, S., Asgari, S., Khromykh, A., 2014. Expression of mosquito 502 microRNA aae-miR-2940-5p is down-regulated in response to West Nile Virus infection to 503 restrict viral replication. J. Virol. DOI: 10.1128/JVI.00317-14. 504

    Tauszig-Delamasure, S., Bilak, H., Capovilla, M., Hoffmann, J.A., Imler, J.L., 2002. Drosophila 505 MyD88 is required for the response to fungal and Gram positive bacterial infections. Nat. 506 Immunol. 3, 91-97. 507

  • 19

    Thirugnanasambantham, K., Hairul-Islam, V., Saravanan, S., Subasri, S., Subastri, A., 2013. 508 Computational approach for identification of Anopheles gambiae miRNA involved in 509 modulation of host immune response. Appl. Biochem. Biotechnol. 170, 281-291. 510

    Towb, P., Bergmann, A., Wasserman, S.A., 2001. The protein kinase Pelle mediates feedback 511 regulation in the Drosophila Toll signaling pathway. Development 128, 4729–4736. 512

    Trobaugh, D., Gardner, C., Sun, C., Haddow, A., Wang, E., Chapnik, E., Mildner, A., Weaver, S., 513 Ryman, K., Klimstra, W., 2013. RNA viruses can hijack vertebrate microRNAs to suppress 514 innate immunity. Nature 506, 245-248. 515

    Wang, R., Deng, F., Hou, D., Zhao, Y., Guo, L., Wang, H., Hu, Z., 2010. Proteomics of the 516 Autographa californica nucleopolyhedrovirus budded virions. J. Virol. 84, 7233–7242. 517

    Weber, A.N., Tauszig-Delamasure, S., Hoffmann, J.A., Lelie`vre, E., Gascan, H., Ray, K.P., 518 Morse, M.A., Imler, J.L., Gay, N.J., 2003. Binding of the Drosophila cytokine Spätzle to Toll is 519 direct and establishes signaling. Nat. Immunol. 4, 794-800. 520

    Winter, F., Edaye, S., Hu¨ ttenhofer, A., Brunel, C., 2007. Anopheles gambiae miRNAs as actors of 521 defence reaction against Plasmodium invasion. Nucleic Acids Res. 35, 6953–6962. 522

    Wu, P., Han, S., Chen, T., Qin, G., Li, L., Guo, X., 2013. Involvement of microRNAs in infection 523 of silkworm with Bombyx mori cytoplasmic polyhedrosis virus (BmCPV). PLoS ONE 8, 524 e68209. 525

    Wu, Y.L., Wu, C.P., Liu, C.Y., Hsu, P.W., Wu, E.C., Chao, Y.C., 2011. A non-coding RNA of 526 insect HzNV-1 virus establishes latent viral infection through microRNA. Sci. Rep. 1, 60. 527

    Xia, T., O’Hara, A., Araujo, I., Barreto, J., Carvalho, E., Sapucaia, J.B., Ramos, J.C., Luz, E., 528 Pedroso, C., Manrique, M., Toomey, N.L., Brites, C., Dittmer, D.P., Harrington Jr., W.J., 2008. 529 EBV microRNAs in primary lymphomas and targeting of CXCL-11 by ebv-mir-BHRF1–3. 530 Cancer Res. 68, 1436–1442. 531

    Ye, Y.H., Woolfit, M., Huttley, G.A., Rancès, E., Caragata, E.P., Popovici, J., O'Neill, S.L., 532 McGraw, E.A., 2013. Infection with a virulent strain of Wolbachia disrupts genome wide-533 patterns of cytosine methylation in the mosquito Aedes aegypti. PLoS ONE 8, e66482. 534

    Zhang, G., Hussain, M., O’Neill, S., Asgari, S., 2013. Wolbachia uses a host microRNA to regulate 535 transcripts of a methyltransferase, contributing to dengue virus inhibition in Aedes aegypti. 536 Proc. Natl. Acad. Sci. USA 110, 10276-10281. 537

    Zhang, J., He, Q., Zhang, C.-D., Chen, X.-Y., Chen, X.-M., Dong, Z.-Q., Li, N., Kuang, X.-X., 538 Cao, M.-Y., Lu, C., Pan, M.-H., 2014a. Inhibition of BmNPV replication in silkworm cells 539 using inducible and regulated artificial microRNA precursors targeting the essential viral gene 540 lef-11. Antiviral Res. 104, 143-152. 541

    Zhang, X., Zheng, Y., Jagadeeswaran, G., Ren, R., Sunkar, R., Jiang, H., 2014b. Identification of 542 conserved and novel microRNAs in Manduca sexta and their possible roles in the expression 543 regulation of immunity-related genes. Insect Biochem. Mol. Biol. 47, 12-22. 544

    Zhu, M., Wang, J., Deng, R., Xiong, P., Liang, H., Wang, X., 2013. A microRNA encoded by 545 Autographa californica nucleopolyhedrovirus regulates expression of viral gene ODV-E25. J. 546 Virol. 87, 13029-13034. 547

  • 20

    Table 1. miRNAs encoded in viral genomes that target viral own or host transcripts. 548

    549 Virus miRNA and its

    origin Target in viral/host

    genome Effect on virus

    infection Reference

    DNA virus Heliothis virescens

    ascovirus (HvAV3e)

    HvAV-miR-1 is derived from MCP

    Downregulates DNA polymerase 1

    limits viral DNA replication during late infection

    (Hussain et al., 2008)

    Heliothis zea nudivirus (HzNV-1)

    hv-miR-246 and -2959 from pag-1 gene

    Both miRNAs target hhi1 transcript

    help virus to enter latent phase

    (Wu et al., 2011)

    Bombyx mori nucleopolyhedrovirus (BmNPV)

    BmNPV-miR-1 downregulates Exportin-5 cofactor Ran resulting in the blockage of pre-miRNA export into the cytoplasm

    (Singh et al., 2012)

    BmNPV

    bmnpv-miR-3 is derived from the plus strand early in infection

    targets 3′UTR of BmNPV DNA binding protein (P6.9) mRNA that transcribes from the opposite strand at the same locus

    helps virus to escape the early immune response from the host

    (Singh et al., 2014)

    Autographa californica nucleopolyhedrovirus (AcMNPV)

    AcMNPV-miR-1 is located on the negative strand of the core gene ODV-E25

    downregulates ODV-E25 mRNA on positive strand

    results in a reduction of infectious budded virions and accelerates the formation of occlusion-derived virions

    (Zhu et al., 2013)

    RNA virus Dengue virus-2

    (DENV-2) DENV-vsRNA-5 is produced from a stem-loop in the 3’UTR of DENV-2

    NS1 gene region Autoregulates viral replication late in infection

    (Hussain and Asgari, 2014)

    West Nile virus (Kunjin strain, WNVKUN)

    KUN-miR-1 is derived from the 3’SL located at the end of the 3’UTR

    Targets and induces transcript level of cellular GATA-4 transcription factor

    Enhances viral RNA replication

    (Hussain et al., 2012)

    550

  • 21

    Figure legend: 551

    Figure 1: Canonical pathway of miRNA biogenesis. miRNA genes are transcribed in the nucleus 552

    mostly by RNA polymerase II in the form of primary miRNA (pri-miRNA) that contains one or 553

    more stem-loop structures, and similar to mRNAs it is 5’capped and polyadenylated. The stem-loop 554

    is cleaved from pri-miRNA by Drosha in association with Pasha to form the precursor miRNA (pre-555

    miRNA which is then transported into the cytoplasm facilitated by Exportin-5/Ran. The hairpin 556

    head is removed either by Dicer-1 or Ago2 resulting in a miRNA duplex which initiates the 557

    formation of the RNA induced silencing complex (RISC). One of the strands in usually degraded 558

    (miRNA*) and the other strand guides the miR-RISC complex to target mRNA (posttranscriptional 559

    regulation) or transported into the nucleus where it binds to a promoter region and regulate gene 560

    transcription (transcriptional gene regulation, TGR). The miRNA* may also be loaded into one of 561

    the Ago proteins and perform a function by interacting with target sequences. The outcome of 562

    miRNA-target interaction may vary from mRNA degradation, translational repression to 563

    upregulation of target transcript levels. 564

    565

  • 566

    567

  • Graphical abstract

    AAAAA

    5’ cap

    Ago2 TRBP

    AAAA 5' cap pri-miRNA

    Drosha

    Pasha pre-miRNA

    Exportin-5

    miRNA duplex

    Dicer-1

    miR-RISC

    miR*-RISC

    TGR

    RNA Pol II

    AAAAA

    5’ cap

    nucleus

    cytoplasm

  • 24

    Highlights 572

    •••• MiRNAs are small non-coding RNAs that regulate gene expression. 573 •••• Following infection, miRNA profile of the host is altered either due to defence or manipulation by the 574

    pathogen. 575 •••• Virus-encoded miRNAs regulate host or viral genes facilitating establishment of infection. 576 577

    578