Next Article in Journal
Loss of Actin-Based Motility Impairs Ectromelia Virus Release In Vitro but Is Not Critical to Spread In Vivo
Previous Article in Journal
Molecular Requirements for Self-Interaction of the Respiratory Syncytial Virus Matrix Protein in Living Mammalian Cells
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Are microRNAs Important Players in HIV-1 Infection? An Update

1
Center for AIDS Health Disparities Research, Tennessee Center for AIDS Research; Meharry Medical College, Nashville, TN 37208, USA
2
Department of Biochemistry and Cancer Biology, Meharry Medical College, Nashville, TN 37208, USA
3
Department of Graduate Studies and Research, Meharry Medical College, Nashville, TN 37208, USA
*
Author to whom correspondence should be addressed.
Submission received: 7 January 2018 / Revised: 21 February 2018 / Accepted: 25 February 2018 / Published: 3 March 2018

Abstract

:
HIV-1 has already claimed over 35 million human lives globally. No curative treatments are currently available, and the only treatment option for over 36 million people currently living with HIV/AIDS are antiretroviral drugs that disrupt the function of virus-encoded proteins. However, such virus-targeted therapeutic strategies are constrained by the ability of the virus to develop drug-resistance. Despite major advances in HIV/AIDS research over the years, substantial knowledge gaps exist in many aspects of HIV-1 replication, especially its interaction with the host. Hence, understanding the mechanistic details of virus–host interactions may lead to novel therapeutic strategies for the prevention and/or management of HIV/AIDS. Notably, unprecedented progress in deciphering host gene silencing processes mediated by several classes of cellular small non-coding RNAs (sncRNA) presents a promising and timely opportunity for developing non-traditional antiviral therapeutic strategies. Cellular microRNAs (miRNA) belong to one such important class of sncRNAs that regulate protein synthesis. Evidence is mounting that cellular miRNAs play important roles in viral replication, either usurped by the virus to promote its replication or employed by the host to control viral infection by directly targeting the viral genome or by targeting cellular proteins required for productive virus replication. In this review, we summarize the findings to date on the role of miRNAs in HIV-1 biology.

1. Introduction

The human immunodeficiency virus 1 (HIV-1) belongs to the Lentivirus subfamily of retroviruses and causes the acquired immunodeficiency syndrome (AIDS). HIV-1 primarily targets and replicates in the cluster of differentiation 4+ (CD4+) T-lymphocytes and cells of the monocyte/macrophage lineage. HIV-1 pathogenesis progresses through three major stages: acute, chronic, and AIDS [1]. During the acute infection phase that typically lasts one to two months, the virus replicates rapidly and, by triggering multiple cell death pathways, causes loss of the CD4+ T-cells [2]. The subsequent chronic infection phase is marked by generalized immune activation and low levels of virus replication, but is accompanied by a sustained loss of CD4+ T-cells [3]. It might be several years (10 years on average) before the infection progresses to AIDS, which is characterized by low CD4+ T cell counts and the development of opportunistic infections and tumors.
HIV-1 infections have reached pandemic proportions since an ancestral form of the virus entered the human population over a century ago through zoonotic transmission from a simian immunodeficiency virus (SIV)-infected chimpanzee in central Africa [4,5,6]. The ongoing global HIV pandemic has already claimed over 35 million lives, with Sub-Saharan Africa, home to nearly 70% of HIV-1 infected people, continuing to bear the brunt. While an effective anti-HIV vaccine remains elusive [7,8,9], an estimated 1.8 million people were infected with HIV-1 in 2016 and millions continue to die each year (one million deaths in 2016) from HIV/AIDS-related illnesses worldwide. The only recourse for the more than 36 million people currently infected with HIV-1 are antiretroviral drugs (ARV) that inhibit various stages of the virus life cycle [10]. However, drug-resistant HIV-1 strains rapidly emerge due to mutations in the viral genome that largely ensue from a combination of high replication rate and low-fidelity of the viral enzyme that copies the viral RNA genome into a double-stranded complementary DNA (cDNA) and, potentially, due to low-fidelity of the host DNA-dependent RNA polymerase II that transcribes the integrated viral DNA genome [11,12]. Hence, a combination antiretroviral therapy (cART) comprising of a cocktail of potent ARVs is widely used for treatment to minimize the emergence of drug resistance [10]. The resultant reduction in viral load, restoration or maintenance of immunocompetence, prevention of or delay in the onset of AIDS, and improved quality of life, have all together significantly reduced HIV/AIDS-associated mortality and morbidity rates [13]. Nevertheless, persistence of stable viral reservoirs in resting memory CD4+ T-cells (HIV-1 latency) has precluded continued suppression of viral load to undetectable levels upon cessation of cART (functional cure) or eradication of the virus (sterilizing cure) [14,15,16,17,18,19,20,21,22,23,24]. The prevalence of drug-resistant HIV-1 strains [25,26,27,28] and the enduring challenge posed by the latent viral reservoirs necessitates urgent development of novel ARVs that target unexploited stages of HIV-1 life cycle [29].
Both viral and host factors determine the clinical course of HIV-1 infection [1,30,31,32,33,34,35,36,37]. Recent progress in identifying and determining the role of host factors and cellular processes in virus biogenesis and transmission underscore the prospect of novel and promising therapeutic strategies for the prevention and/or management of, and possibly cure for, HIV/AIDS [38,39]. Notably, the cellular gene silencing process termed RNA interference (RNAi), which is mediated by small non-coding RNAs (sncRNAs) and can modulate gene expression at the posttranscriptional level, is increasingly being recognized as a prime platform for developing and delivering non-traditional antiviral therapeutic strategies.
In this review, we start with a primer on the HIV-1 replication cycle in target cells, followed by a concise overview of the nature and role of cellular sncRNAs in host gene expression, and an outline of the current understanding of the miRNA biogenesis. We then discuss the reported roles of cellular miRNAs as cofactors or inhibitors of HIV-1 replication, all the while drawing upon, to provide a broader context and the current status of this rapidly evolving field, topical findings from studies on other viruses. We also address the ongoing debate on whether HIV-1 encodes miRNAs and, if so, their functional relevance. We note here that several reviews discussing the interplay between viruses and cellular miRNA machinery have been published in the recent past [40,41,42,43,44,45,46,47,48,49].

2. HIV-1 Replication

Three decades of intense research on HIV/AIDS has considerably improved our understanding of virus infection and transmission [23]. Basic research on HIV/AIDS has made inroads into several aspects of the replication strategy employed by the virus [50]. A relatively small virus (~100–120 nm), the spherical HIV-1 particles contain two copies of non-covalently linked positive-sense single-stranded RNA genome of ~10 kilobases (Kb) size [51,52]. This viral genomic RNA (vRNA), harboring several functional secondary structures, contains nine genes encoding 15 canonical proteins [53,54]. As an obligate intracellular parasite, HIV-1 relies on host cell machinery and proteins for its productive replication in infected cells. Remarkably, over 2000 host cellular proteins, termed host-dependency factors (HDFs), have been implicated in HIV-1 replication cycle [55,56]. The virus also encounters, and in most cases, overcomes the restriction imposed by the host-encoded intrinsic and innate immune factors [34,36,57,58]. To gain entry into the target cell, HIV-1 uses the viral membrane-anchored envelope glycoprotein (gp120) to first bind to the CD4 receptor and, then, the C-C chemokine receptor type 5 (CCR5) or C-X-C chemokine receptor type 4 (CXCR4) chemokine co-receptor on the target cell surface [59]. These virus–cell interactions promote fusion of their respective membranes thereby allowing entry and cytoplasmic release of the viral “core”—a cone-shaped protein shell composed of polymer of the viral capsid protein (CA) [60,61,62,63,64]. The viral core harbors the viral RNA genome, virally-encoded enzymes and proteins, and host proteins; the viral enzymes reverse transcriptase (RT) and integrase (IN) are essential for mediating the next steps in viral replication. The viral RT, functioning as a component of the nucleoprotein complex termed reverse transcription complex (RTC), synthesizes a double-stranded cDNA (vDNA) copy of the vRNA [11]. This reverse transcription process is intricately synchronized with the disassembly of the CA shell of the viral core (uncoating), a highly ordered, but poorly understood, process that regulates the subsequent early post-entry steps of viral replication including nuclear entry and integration of the vDNA [62,65,66]. The viral IN, functioning as a component of another nucleoprotein complex termed pre-integration complex (PIC), traffics the vDNA into the nucleus and enables its integration into the host chromosome [67,68,69]. The integrated vDNA (provirus) may then, be transcribed or remain silent as latent provirus [22,70]. The cellular RNA polymerase II transcribes the proviral DNA into viral mRNAs, a population of which undergoes further processing including splicing. The host cell machinery mediates the nuclear export of the differentially spliced viral mRNAs whereas virus-encoded proteins facilitate the transport of the unspliced full-length vRNA to the cytoplasm [71]. The viral mRNAs are translated into the respective viral proteins competent for structural (Gag polyprotein (Pr55Gag), Gag-Pro-Pol polyprotein (Pr160GagPol), envelope glycoprotein polyprotein (gp160)), accessory (Nef, Vif, Vpu, Vpr) and regulatory (Tat, Rev) roles. Subsequently, viral constituents necessary for the generation of progeny virions including the Gag, Gag-Pro-Pol, envelope glycoproteins, and the vRNA traffic to specific intracellular membrane niches, where the Gag orchestrates their assembly into immature virus-like particles (VLPs) [72,73,74,75,76]. The cellular endosomal sorting complexes required for transport (ESCRT) machinery mediates the subsequent budding and pinching of the assembled VLPs into the extracellular milieu and the concomitant PR-mediated maturation process yields infectious progeny virus (virions) [72,77,78,79]. During the assembly and budding process, several host RNAs and proteins from the producer cell are incorporated into the virus particle [80,81,82]. Finally, transmission of the progeny virions leads to the establishment of new infections [83,84].

3. Regulation of Cellular Gene Expression by Small Non-Coding RNAs

Gene expression is regulated at transcriptional, post-transcriptional, and post-translational stages. Research in the past two decades has uncovered RNAi, a key and sophisticated cellular pathway in eukaryotes that mediates gene silencing via the action of sequence-specific sncRNAs [85,86]. Three major classes of sncRNAs that mediate RNAi in animals include the small interfering RNAs (siRNAs), microRNAs (miRNAs), and Piwi-interacting RNAs (piRNAs) [87,88,89,90,91]. Additional types of sncRNAs possessing diverse biological roles continue to be discovered. RNAi-mediated gene silencing can be executed at the level of transcription—termed transcriptional gene silencing (TGS), or post-transcription—termed post-transcriptional gene silencing (PTGS). TGS involves stable repression of transcription via DNA methylation or histone modification and has generally been recognized as a cellular gene regulatory mechanism in lower eukaryotes and plants. RNAi-mediated PTGS involves sncRNAs acting in concert with the Argonaute (AGO) family proteins to form the multi-protein RNA-induced silencing complexes (RISCs), which then cleave the transcript or inhibit translation [92].
The siRNAs and miRNAs are generated by the action of cellular RNase III enzymes and are posttranscriptional regulators of gene expression [93,94,95,96,97]. However, they exhibit structural, physiological and functional differences, and their target mRNAs face different fates [98]. The siRNAs are ~21-nucleotides-long double-stranded (ds) sncRNA molecules derived by sequential exonucleolytic processing of longer, linear, completely complementary dsRNA precursors by the RNAse III enzyme Dicer. One strand of the siRNA, termed as the “guide strand”, then incorporates into and guides the RISC to the target mRNA displaying perfect sequence similarity. The RISC then endonucleolytically cleaves and degrades the target mRNA. This sequence complementarity-dependent activity of siRNAs ensures knockdown of specific genes. The siRNAs derived from endogenous or exogenous dsRNAs have been categorically shown to silence host genes or viral genes in plants and invertebrates, both of which lack any antibody or cell-mediated immunity. However, their existence and function in mammals, especially in somatic cells, remains ill-defined and controversial. In mammalian cells, siRNAs typically arise from exogenous dsRNAs, which typically are viral replication intermediates, imperfect RNA hairpins in the viral genome, or experimentally introduced substrates. Unlike the siRNAs, mature functional miRNAs are derived from longer, incompletely complimentary stemloop dsRNA precursors termed primary miRNA (pri-miRNA), which are either encoded by cellular genes located primarily in the intergenic regions or present in the introns intervening the protein-coding genes. Downstream sequential processing steps involving multiple proteins or protein complexes yield the ~22-nucleotides-long single-stranded mature miRNAs (Figure 1) (for detailed discussion on miRNA biogenesis, please see Section 4). The mature miRNAs harbor a “seed region”, defined by nucleotides 2–8, that recognizes and binds perfectly complementary sequences in the target mRNAs; binding beyond this seed region is generally based on partial complementarity. The bound miRNAs regulate the expression of target mRNAs at the post-transcriptional level via translational repression and/or mRNA deadenylation or decay. The miRNAs exhibit partial complementarity to target mRNAs and hence may regulate hundreds of target genes. Further, miRNAs have been shown to occur and function exclusively in eukaryotes—plants and animals. In contrast to the siRNAs and the miRNAs, the ~24–30 nucleotides-long single-stranded piRNAs are generated from single-stranded RNA precursors independent of any RNase III enzyme activity. The piRNAs are specifically expressed in germ line cells of vertebrates and invertebrates, wherein they bind PIWI subfamily members of the AGO family proteins and the resulting RNA-protein complexes silence, primarily, the transposons [89,99,100].

4. Cellular microRNAs

miRNAs modulate diverse cellular processes including development, differentiation, metabolism, proliferation, tumorigenesis, immune response, and cell death [101,102]. Many miRNA genes are conserved from C. elegans to humans. To date, the human genome has been reported to encode anywhere from 500 (MirGeneDB, v1.1 [103]) to over 1880 miRNAs (miRBase, Release 21 [104]), randomly distributed in all chromosomes except for the Y chromosome. Interestingly, a recent study of 13 different human tissue samples revealed the presence of 3707 novel miRNAs, of which over half were human-specific [105]. However, the functional relevance of the majority of these miRNA annotations is currently unknown. Nevertheless, miRNAs have been projected to regulate the expression of up to 60% of mammalian [106,107,108] or 90% of human protein-encoding genes [109]. Because miRNAs are only partially complementary to their target mRNAs, each miRNA is estimated to be capable of regulating the expression of up to 200 genes [110,111]. Conversely, binding sites for multiple miRNAs have been reported to be present on individual target mRNAs [112]. In the canonical miRNA-processing pathway, specific transcription factors stimulate the transcription of the primary miRNA (pri-miRNA) transcripts from miRNA-encoding genes located within the introns or 3′-untranslated region (UTR) of coding or non-coding mRNAs, primarily by RNA polymerase II [113,114]. However, RNA polymerase III catalyzes the transcription of pri-miRNAs corresponding to certain miRNAs, including those encoded within Alu repeat sequences and by some viruses. A large percentage of human miRNAs (~40%) are cotranscribed as miRNA clusters, i.e., a single pri-miRNA transcript harbors multiple miRNA sequences (Figure 1). The pri-miRNAs harbor stem loop (hairpin) structures and undergo sequential processing, first in the nucleus and then in the cytoplasm. In the nucleus of mammalian cells, a heterotrimeric “microprocessor” complex, containing one copy of the RNase III endonuclease Drosha and two copies of its cofactor the ds-RNA binding protein DiGeorge syndrome critical region gene 8 (DGCR8), cleaves the pri-miRNA ~11 bp from the basal junction and ~22 bp from the apical junction [115,116,117,118,119,120]. This releases on average ~70-nucleotides-long (may range from ~60–110 nucleotides) hairpin precursors known as precursor miRNAs (pre-miRNAs) that harbor a two-nucleotide 3′ overhang. The pre-miRNAs are then exported by a protein complex containing Exportin 5 and RanGTP into the cytoplasm [121,122,123]. In the cytoplasm of mammalian cells, the RNase-III family enzyme Dicer, in concert with HIV-transactivating response RNA-binding protein 2 (TRBP2) and protein activator of the interferon-induced protein kinase (PACT), recognizes the pre-miRNA’s 2 nucleotide 3′ overhang and cleaves off the loop structure ~22 nucleotides from the Drosha cleavage site [124,125,126,127,128,129,130,131,132]. The released asymmetric ~22-nucleotides-long miRNA duplexes harbor a 5′ monophosphate and a two-nt 3′ overhang on both strands-features that help them get loaded, aided by heat shock cognate 70 kDa (Hsc70)/heat shock protein 90 (Hsp90) chaperones and their co-chaperones, onto the AGO protein complex [133,134,135,136]. The miRNA duplexes are then unwound and, in a strand selection process based on 5′ nucleotide composition and thermodynamic stability, the concomitant dissociation or degradation of one of the strands (passenger strand) finally yields the mature miRNA (guide strand); the ensuing effector nucleoprotein complexes are called the miRNA-induced silencing complexes (miRISC) [137,138,139,140,141,142]. A seed region comprising of nucleotides 2–8 located in the 5′ region of the guide miRNA strand determines its ability to recognize the target mRNA harboring complementary sequences [143,144,145,146,147]. This biologically active mature miRNA guides the miRISC complex to bind the target mRNA, typically in the 3′UTR but potentially in the coding region [148] and in the 5′UTR [149] and cause translational suppression or mRNA deadenylation or mRNA degradation [88,150,151,152,153,154,155,156]. Recent research has also identified miRNA biosynthetic pathways in mammalian cells that deviate from the canonical pathway, thereby adding multiple layers of complexity to the miRNA-mediated gene regulation. Such alternative miRNA biosynthesis pathways may be Drosha/DGCR8-independent (e.g., Mirtrons), Dicer-independent (e.g., miR-451), and both Drosha/DGCR8- and Dicer-independent (e.g., Agotrons) [157,158,159,160,161,162,163,164].

5. Cellular miRNAs as Cofactors in HIV-1 Replication

Viruses generally encode minimal genetic information, and hence depend on host cellular factors and pathways to complete productive replication. Studies utilizing genomic, transcriptomic, and proteomic approaches have identified hundreds, and in some studies, thousands of host cell factors involved in different stages of virus replication. Interestingly, striking overlaps exist between the cellular pathways usurped by the virus and those regulated by the miRNAs. Accordingly, cellular miRNAs have been reported to affect virus infections in a positive or negative manner.
Cellular miRNAs can positively impact virus replication directly by associating with the virus or viral components to counter the antiviral pathways. A striking example, and a first report involving an RNA virus, was the demonstration that the liver-specific miR-122 binds directly to the 5′-UTR of the hepatitis C virus (HCV) RNA without affecting the viral mRNA stability or translation. Interestingly, the virus replication was reduced when the miR-122 binding to HCV RNA was blocked, thus indicating that miR-122 promoted HCV replication [165]. Subsequently, miR-122, in concert with AGO2, was shown to protect the bound HCV RNA from host mRNA decay machinery [166,167]. Interestingly, miR-122 was recently shown to primarily enhance the viral RNA synthesis by competing with the poly(rC)-binding protein (PCBP2), thus steering the viral genome away from translation [168]; this, albeit indirectly, reinforces the canonical role of miRNAs as translational repressors. The recognition of miR-122 as an essential cofactor for HCV replication has resulted in the development, and clinical trial, of an miR-122-targeting antimiR. In an example of a miRNA targeting the 3′-UTR of RNA viruses, the bovine viral diarrhea virus (BVDV) and classical swine fever virus (CSFV) were shown to rely on the binding of cellular miR-17 family to the viral 3′-UTR towards promoting viral RNA stability and translation [169]. Cellular miRNAs may also be sequestered by viruses, thereby protecting cellular cofactors from miRNA-mediated downregulation. Interestingly, both miR-122 and miR-17 has been shown to act as “miRNA sponges”, thus altering the host transcriptome [169,170]. Further, the hepatitis B virus (HBV) RNA harbors six miR-122-binding sites that act as sponges of miR-122; the ensuing sequestration and downregulation of miR-122 was proposed to steer them away from their inhibitory action on cellular proteins that aid HBV replication [171,172]. Notably, to date no cellular miRNA has been demonstrated to promote HIV-1 replication by directly associating with the vRNA.
Cellular miRNAs can also positively impact virus replication in an indirect manner, as evidenced by the induction of such miRNAs by some viruses. In an example of an RNA virus employing cellular miRNA to selectively disrupt host machinery, the enterovirus EV71 was shown to upregulate cellular miR-141, which binds the 3′UTR of the eukaryotic translation initiation factor 4E (eIF4E) mRNA and downregulates its expression, thereby facilitating virus-mediated switch from cellular to viral translation [173]. The eastern equine encephalitis virus (EEEV) has reportedly evolved an intriguing replication strategy wherein binding of the cellular miR-142-3p to the 3′UTR of the viral RNA and the ensuing suppression of viral replication was proposed to aid the virus avoid triggering the host interferon (IFN) system [174]. Interestingly, the Epstein-Barr virus (EBV) appears to have multiple options in cellular miRNAs towards establishing and maintaining latent infections. The EBV induces expression of several cellular miRNAs including miR-155 [175] and let-7a [176]; while miR-155 suppresses the expression of cellular proteins mediating cell cycle arrest and apoptosis, thus ensuring EBV latency, the let-7a downregulates Dicer and, as a consequence, its positive effect on EBV reactivation. Conversely, the cellular miRNAs miR-200b and miR-429 were shown to down-regulate the expression of two cellular proteins, zinc finger E-box binding homeobox 1 (ZEB1) and , zinc finger E-box binding homeobox 2 (ZEB2), known to play critical roles in the maintenance of EBV latency, thereby relieving the block on virus reactivation from latency and production of infectious virus [177].
Chiang et al. [178] recently reported the first cellular miRNA that appears to promote HIV-1 replication in a CD4+ T-cell line. Activation of CD4+ T-cells is a prerequisite for productive HIV-1 infection and entails substantial transcriptional changes, including alterations in cellular miRNA expression. Interestingly, CD4+ T-cell activation has been shown to diminish cellular levels of several anti-HIV miRNAs that directly target the viral RNA [179,180] or inhibit cellular cofactors [181,182,183]. So, Chiang et al. [178] hypothesized that miRNAs upregulated upon CD4+ T-cell activation might promote HIV-1 replication. Accordingly, when compared to the resting cells, miR-132 levels were significantly upregulated in phytohaemagglutinin (PHA)-activated CD4+ T-cells isolated from healthy donors and peripheral blood mononuclear cells (PBMC) from HIV-1 patients. Conversely, overexpression of exogenous miR-132 promoted HIV-1 replication in Jurkat T cells and reactivated HIV-1 in a latently infected Jurkat cells. Interestingly, miR-132 has previously been reported to target and down-regulate a cellular transcriptional regulatory protein called Methyl-CpG-binding protein 2 (MeCP2) [184,185]. In accordance with the miR-132 overexpression data, MeCP2 knockdown augmented HIV-1 production. However, MeCP2 knockdown did not reactivate latent HIV-1, which suggests that miR-132 modulates the biogenesis of HIV-1 and reactivation of latent HIV-1 via two mechanistically different routes. Predating the report by Chiang et al. [178], two cellular miRNAs- miR-34a and miR-217, were reported to bind the 3′UTR, and repress the translation, of the mRNA of silent mating type information regulation 2 homolog 1 (SIRT1) [186,187]. The SIRT1 is a cellular protein deacetylase known to deacetylate the nuclear factor kappa B (NF-kB) and inhibit its positive effect in Tat-induced transactivation of HIV-1 long terminal repeat (LTR). Therefore, the miRNA-induced perturbation of SIRT1′s deacetylation activity resulted in increased acetylation of NF-kB and, consequently, enhanced transactivation of HIV-1 LTR. However, these studies investigated the effect of miRNAs on the transactivation of HIV-1 LTR in the absence of virus infection by using TZM-bl cells (containing integrated HIV-1 LTR-driven Luciferase reporter) or Magi cells (containing integrated HIV-1 LTR-driven beta-galactosidase reporter), or SupT1 cells transfected with LTR-Luc construct.
Additional cellular miRNAs enabling HIV-1 infection by downregulating the levels of host antiviral proteins have been reported. The cyclin-dependent kinase inhibitor protein p21 is known to disrupt HIV-1 gene expression by blocking the transcriptional elongation factor cyclin-dependent kinase 9 (CDK9), and the potassium channel protein TWIK-related acid-sensitive potassium channel 1 (TASK1) impairs the viral protein Vpu-mediated enhancement in HIV-1 release [188,189]. Farberov et al. [190] reported that the downregulation of p21 and TASK1 by the cellular miRNAs Let-7c and miR-124a or miR-34a-5p, respectively, led to increased HIV-1 replication in the JLTRG-R5 (Jurkat-derived cells containing integrated HIV-1 LTR-driven GFP reporter) and HeLa-CCR5 cells. Another member of the cellular miR-34 family, the miR-34c-5p was also recently shown to promote HIV-1 infection in CD4+ T-cells [191]. The expression of miR-34c-5p was upregulated during the T cell receptor (TCR) signaling-mediated activation of naïve CD4+ T-cells, which altered the expression levels of multiple genes involved in HIV-1 replication. Interestingly, the enhancement of HIV-1 replication in T-cells overexpressing miR-34c-5p and the downregulation of endogenous miR-34c-5p upon HIV-1 infection led the authors to speculate that the latter is an antiviral host response.

6. Cellular miRNAs as Anti-Viral Factors in HIV-1 Replication

Cellular miRNAs may also negatively modulate virus replication. This is unsurprising because siRNAs, which inhibit mRNA expression by comparable mechanism, have been shown to effectively impair virus replication. In fungi, plants, and invertebrates, siRNA-mediated RNAi functions as an innate defense response against virus infections. The dsRNA intermediates generated during virus replication, especially that of RNA viruses, are processed by DICER into virus siRNAs (vsiRNAs) that direct the RISC machinery to silence the corresponding viral RNA. Invariably, viruses have evolved to equip themselves with evasion strategies to counter this host-mediated defense response. For instance, viruses are known to generate viral variants that have acquired escape mutations in the RNAi target sequence, and the ensuing sequence noncomplementarity or modified local RNA secondary structure might impair siRNA recognition and binding to the viral target sequence. A more prevalent strategy employed by plant viruses is to encode viral suppressors of RNAi (VSRs) that, besides playing crucial roles in the virus replication, block certain steps of the RNAi pathway via different mechanisms. For example, VSRs might bind to long dsRNA duplexes to block siRNA synthesis or to siRNA duplexes to preclude the RISC formation; they might also directly target the constituents of the RISC machinery or the host enzymes catalyzing DNA methylation.
Though some mammalian viruses have been reported to encode VSR proteins, including influenza A virus (IAV) NS1, Ebola virus (EBOV) VP35, and human enterovirus 71 (HEV71) 3A, the existence of an mammalian RNAi pathway possessing antiviral potential, especially in somatic cells, remains a contentious area of research [192,193,194,195]; the lack of or low levels of vsiRNAs in virus-infected mammalian cells [196,197] and the lack of enhanced virus infection in Dicer-deficient cells [198,199] are often cited in argument against the existence of mammalian antiviral RNAi. Adding to the complexity, reciprocal inhibition between the RNAi machinery and the interferon response pathways involving interferon-stimulated genes (ISGs) has been reported. Seo et al. [200] demonstrated that RNAi represses the expression of ISGs in uninfected cells to avert cytotoxicity, but virus infection and the ensuing antiviral responses relieve the inhibitory effect on the expression of ISGs. This, the authors proposed, indicates that interferons exclusively mediate the cellular response to virus infection. Backes et al. [201] came to similar conclusion after comparing the titers of recombinant viruses capable of disrupting IFN-mediated response or RISC-associated silencing; while impaired IFN response yielded higher virus titers, inactivation of small RNA-mediated silencing in the presence or absence of IFN response lowered the virus titers. These findings led the authors to propose that mammalian response to virus infection is not only independent of RNAi pathway.
However, infection of IFN-nonresponsive mouse embryonic stem cells with encephalomyocarditis virus (EMCV) or Nodamura virus (NoV) [202], and infection of differentiated cells by a NoV mutant lacking its VSR B2, both induced the production of siRNAs [203]. Despite the lack of evidence showing that these siRNAs were functional products of the canonical RNAi machinery and the use of stem cells not naturally targeted by EMCV or NoV, these studies raised the possibility that the prior failed attempts at demonstrating antiviral RNAi in mammals could simplty reflect the masking or inhibitory effect by the IFN pathway and/or by the virus-encoded VSRs. Accordingly, Maillard et al. [204] demonstrated that mammalian somatic cells with impaired IFN pathway were capable of mounting antiviral RNAi, and Li et al. [205] reported that infection by a mutant IAV lacking the VSR NS1, but not by the WT virus, led to generation of vsiRNAs in a Dicer-dependent manner. More recently, a mutant HEV71 lacking the VSR 3A was shown to be replication-defective and to induce vsiRNA production and RISC-loading in human somatic cells and mice [206]; the partial rescue of the replication defect in Dicer-deficient cells, independent of IFN pathway, was cited as supportive evidence for the functional nature of these vsiRNAs. While these findings imply that the constraints hampering the validation of mammalian antiviral RNAi might be a reflection of the highly potent VSRs effectively blocking the generation of vsiRNAs, which negates any potentiating effect due to Dicer loss or inactivity, many critical questions remain unanswered. Schuster et al. [207] reported that inactivation of the interferon response pathways in differentiated cells by depleting the cytoplasmic sensors like RIG-1 and MDA5 still failed to uncover vsiRNAs and RNAi-mediated antiviral activity.
The first evidence that host cells might employ a miRNA in antiviral defense was provided by Lecellier et al. [208] who showed that human miR-32 restricts the accumulation of the retrovirus primate foamy virus type 1 (PFV-1) in human cells by targeting viral RNA transcripts. Interestingly, the PFV-1 appears to, apparently taking a page from the playbook used by plant viruses that encode VSRs, preempt this host response via the viral protein Tas that globally represses the miRNA biogenesis pathways including the synthesis of miR-32. Thus, an intricate interplay between the cellular miR-32 expression and virus-encoded Tas protein was proposed to determine the success of PFV-1 infection of target cells. Interestingly, miRNAs have also been proposed to function as IFN-inducible antiviral effectors. The human IFNβ was reported to induce the expression of cellular miRNAs miR-196, miR-296, miR-351, miR-431, and miR-448, which displayed seed-sequence complementarity with HCV RNA genome and downregulated virus replication; conversely, the IFNβ was shown to downregulate the HCV cofactor miR-122, thereby causing a reduction in HCV replication [209]. Taking advantage of the availability of Dicer1-deficient mice, Otsuka et al. [210] determined that cellular miRNAs miR-24 and miR-93 display antiviral activity against vesicular stomatitis virus (VSV) by targeting the viral genes encoding large protein (L protein) and phosphoprotein (P protein). The enterovirus EV71 infection induces the expression of miR-296-5p, which targets the coding region of the viral RNA and inhibits viral replication [211]; interestingly, the virus acquired mutations that enabled it to evade miR-296-5p-mediated suppression. More recently, the IFN-induced miR-342-5p was shown to negatively regulate the cholesterol biosynthesis pathway, which is relied upon by several viruses for replication, thus impairing the replication of pathogenic viruses including Influenza A and cytomegalovirus [212].
As in the case of siRNAs, co-evolution with cellular miRNAs in target cells potentially provides the viruses with necessary selection pressure to develop mechanisms to outmaneuver the antiviral activity of these miRNAs. However, despite exposure to multitude of miRNAs in the target cell, viruses have not generally been observed to alter cellular miRNA levels in infected cells. While poxvirus infection coincides with cellular miRNA degradation [213], HIV-1 and influenza A do not appear to antagonize cellular miRNA function [214,215]. Conversely, viruses might simply exploit one defining functional attribute of the miRNAs to evade surveillance. The miRNAs are evolutionarily conserved and, unlike the siRNAs, display partial homology to the target mRNA sequences. However, exact complementarity to the “seed” sequence comprising nucleotides 2–8 of the miRNAs is critical for mRNA target recognition and binding. Hence, even single nucleotide changes acquired by the target mRNAs, including the viral RNAs, may potentially make them immune to any inhibitory action by miRNAs [216]. A striking example for such escape from inhibitory miRNA has been reported in the case of the human T-lymphotropic virus type 1 (HTLV-I), another retrovirus that has recently shown to be controlled by a cellular miRNA. The miR-28-3p was shown to target and prevent the translation of HTLV-1 RNA [217], whereas miR-28-5p, reported to target the HIV-1 RNA, has no binding site on HTLV-1. Interestingly, the ATK-1 strain of HTLV-1 harbors a single-nucleotide polymorphism within the miR-28-3p binding site, and, thus, is resistant to the miR-28-3p. HIV-1 has also been reported to evolve viral quasispecies harboring silent codon changes in order to evade RNAi targeting of the critical viral ORFs [218].
Computational target prediction programs have come to represent one of the favored steps in decoding the function of a miRNA, which largely is defined by that of its target genes. Using a combination of four different miRNA target prediction software, Hariharan et al. [219] identified five sites in the HIV-1 genomic RNA that are potential targets of miRNAs. Interestingly, all these miRNA target sites mapped to the viral RNA regions encoding viral accessory proteins: the nef gene harbored target sites for miRNAs miR-29a and miR-29b, the vpr for miR-149, the vif for miR-324-5p, and vpu for miR-378 (Figure 2 and Table 1). In a follow-up study, the authors demonstrated that miR-29a disrupted the expression of viral nef protein and impaired HIV-1 replication in Jurkat cells [220]. Houzet et al. [221], used an in silico approach to identify target sites for 20 cellular miRNAs in HIV-1 RNA, and then experimentally tested the ability of the candidate miRNAs to suppress HIV-1 replication in transfected cells. This led to the narrowing down of the list to five miRNAs: miR-133b, miR-138-5p, miR-326, miR-149-5p, and miR-92a-3p, which all reduced HIV-1 replication by over 40%. Based on experiments that tested the effect of increasing the complementarity between miR-326 and HIV-1 mRNA, the authors contended that the potency of miR-326-mediated repression of HIV-1 replication qualifies for a functionally relevant miRNA–vRNA interaction.
Another widely used strategy to investigate the interplay between miRNAs and HIV-1 has been to analyze changes in cellular miRNA levels in response to HIV-1 infection or downstream host responses. Any such HIV-specific miRNA expression signature may negatively affect HIV-1 replication directly by targeting the virus or indirectly by modulating cellular cofactors or antiviral factors. Yeung et al. [226] employed a high throughput microarray method to quantify the miRNA expression levels in HIV-1-transfected HeLa cells. They reported that, while no cellular miRNAs were upregulated, the HIV-1 infection was accompanied by the downregulation of 312 host miRNAs. RT-PCR analysis of a selective group of four miRNAs from this list—miR-93, miR-148b, miR-221 ad miR-16, supported the microarray data. However, this study did not clarify the mechanistic basis for such extensive downregulation of cellular miRNAs by the virus, i.e., whether all or many of the downregulated cellular miRNAs hold anti-HIV potential. Employing a similar microarray-based approach, Triboulet et al. [181] showed that, while 11 cellular miRNAs (including miR-122, miR-297, miR-370, and miR-373) were upregulated, just two cellular miRNAs, miR-17-5p and miR-20a were downregulated upon HIV-1 infection in the Jurkat cells. Though, neither miR-17-5p nor miR-20a could directly target the viral RNA, both miRNAs had four binding sites located on the 3′-UTR of histone acetyltransferase P300/CBP-associated factor (PCAF), an important cellular cofactor for the viral Tat function. Accordingly, exogenous overexpression of these miRNAs in PBMCs caused substantial depletion of cellular PCAF levels accompanied by impaired HIV-1 replication, whereas blocking the function of the endogenous versions of these miRNAs by antisense inhibitors enhanced PCAF protein expression levels and HIV-1 production. Intriguingly, Whisnant et al. [222] failed to detect any significant alteration in the expression levels of miR-17-5p and miR-20a in HIV-1-infected primary CD4+ PBMCs. Though many potential reasons may underlie such opposing observations, this particular discrepancy further supports the notion that findings from studies that utilize monoculture cell lines optimized for in vitro assays may not necessarily reflect the in vivo situation. For instance, Triboulet et al. [181] used the immortalized Jurkat T-cell line and assayed for the miRNA expression levels 21 days post infection by microarray analysis and, at the earliest time point of, seven days post infection by RT-PCR. In contrast, Whisnant et al. (2013) [222] tested the primary CD4+ PBMCs 3 days post infection, because, as they reason in their report, the HIV-1-infected T cells in vivo have a half-life of only around 24 h [227]. By comparing the miRNA expression profiles of cells productively or latently infected by HIV-1, Wang et al. [224] identified two cellular miRNAs- miR-196b and miR-1290, that were relatively upregulated in latently infected cells. These two miRNAs were also shown to directly target the 3′UTR of HIV-1 transcripts, while overexpression and knockdown experiments conducted using lab cell lines and CD4+ T-cells from HIV-1-infected patients receiving cART revealed the potential of these miRNAs to modulate HIV-1 replication, as evidenced by changes in virus production and infectivity. Based on these results, the authors proposed that miR-196b and miR-1290 may contribute to HIV-1 latency.
A potential link between miRNA expression levels and the recalcitrance of resting memory CD4+ T-cells to HIV-1 infection has also been reported. Compared to the activated CD4+ T-cells, resting memory CD4+ T-cells displayed upregulation of five miRNAs—miR-28, miR-125b, miR-150, miR-223, and miR-382, which negatively targeted the 3′-ends of HIV-1 mRNAs [179]. Accordingly, in resting CD4+ T-cells that were infected with HIV-1 infectious clones or were isolated from cART-experienced HIV-1 patients, knockdown of these miRNAs by antisense inhibitors enhanced both viral protein production and virion production. Addtionally, over-expression and knock-down studies showed that miR-125b expression negatively correlated with HIV-1 infection in a T cell line [228]. Besides the miRNAs that appear to modulate HIV-1 infection by directly targeting the viral RNAs, certain miRNAs reportedly play an indirect role by regulating the expression of cellular cofactors in resting CD4+ T-cells. The protein levels of the cellular cofactor cyclin T1, which plays critical roles in the viral Tat-mediated transactivation of HIV-1 LTR-driven gene expression, were upregulated independent of the transcript levels upon activation of the resting CD4+ T-cells [183]. Interestingly, this upregulation of cyclin T1 was accompanied by a significant downregulation of a group of miRNAs that included miR-27b, miR-29b, miR-150, and miR-223 in activated CD4+ T-cells. The authors verified this observation by overexpressing or depleting the miRNAs, which decreased or increased the cyclin T1 protein levels, respectively. However, only the miR-27b was shown to directly regulate the cyclin T1 expression, whereas miR-29b, miR-150, and miR-223 were all proposed to indirectly affect the cyclin T1 levels.
Monocytes are refractory to productive HIV-1 replication, but become progressively susceptible upon differentiation into monocyte-derived macrophages (MDM) or monocyte-derived dendritic cells (MDDC). Though several mechanisms have been proposed to account for this post-entry restriction to HIV-1, as in the case of resting memory CD4+ T-cells, a potential links between miRNA expression levels and the recalcitrance of monocytes to HIV-1 infection have been reported. Wang et al. [223] reported that monocytes expressed elevated levels of the cellular miRNAs miR-125b-5p, miR-28-5p, miR-150-5p, miR-223-3p, and miR-382-5p, all previously shown by Huang et al. [179] to impair HIV-1 replication in resting CD4+ T-cells. Because knockdown of these miRNAs in monocytes enhanced HIV-1 infection, and miRNA overexpression in MDMs inhibited HIV-1 replication, the authors proposed that the expression levels of these miRNAs determine the susceptibility of monocytes or MDMs to HIV-1.
Interestingly, as in the case of resting CD4+ T-cells, cyclin T1 expression was shown to be upregulated upon differentiation of monocytes into MDMs, which was accompanied by robust downregulation of the cellular miR-198 [182]. The miR-198 was shown to directly target the 3′UTR of cyclin T1, and miR-198 overexpression or knockdown in monocytes undergoing differentiation decreased or increased the cyclin T1 protein levels, respectively, with concomitant effects on HIV-1 replication. Expression of another cellular protein called purine-rich element binding protein α (Pur-α), known to promote tat-mediated transactivation, has also been shown to be diminished in the monocytes compared to the MDDCs. This downregulation of Pur-α in monocytes was attributed to the robust expression of a group of miRNAs that included miR-15a, miR-15b, miR-16, miR-20a, miR-106b, and miR-93, which all directly targeted the 3′UTR of Pur-α [229]. While inhibition of these miRNAs in monocytes, via chemically synthesized miRNA inhibitors, caused only a modest enhancement of HIV-1 infection, suppression of endogenous Pur-α expression in MDDCs caused substantial reduction in virus infection. Further, the repression of Pur-α by multiple miRNAs highlights the cooperativity principle that has been proposed to ensure effective target repression [230]. The diminished expression of another cellular cofactor called Vpr(HIV-1)-binding protein (VprBP) in monocytes compared to the MDDCs has also been attributed to the direct targeting of VprBP 3′UTR by the cellular miRNA miR-1236 [231]. Knockdown of miR-1236 in monocytes or overexpression in MDDCs increased or decreased the VprBP protein levels, respectively, with concomitant effects on HIV-1 replication.
The MDMs, though generally susceptible to HIV-1 infection, are known to become relatively recalcitrant to HIV-1 upon stimulation with the Toll-like receptor 4 (TLR4) agonist polyinosine-polycytidylic acid (poly(I:C)) or Toll-like receptor 4 (TLR4) agonist lipopolysaccharide (LPC). Swaminathan et al. [232] demonstrated that this TLR3/4 stimulation-induced suppression of HIV-1 infection in MDMs could be attributed to enhanced expression of the cellular miR-155, which interfered with the post-entry and pre-integration events in HIV-1 replication, as evidenced from accumulation of late RT products and diminished proviral integration and 2-LTR circles. Interestingly, poly(I:C) stimulation or miR-155 overexpression led to repression of several HDFs implicated in trafficking and nuclear import of PICs, including ADAM10, TNPO3, Nup153, and LEDGF/p75, all of which were shown to be directly targeted at their 3′UTR by the miR-155. The cellular miR-155 has also been shown to reinforce HIV-1 latency by targeting the 3′UTR of the E3 ubiquitin ligase TRIM32 transcript and repressing its expression, thereby blocking its promotion of reactivation of latent HIV-1 [233]. The recent report by Lodge et al. [234] adds two cellular miRNAs, miR-221 and miR-22, to the repertoire of miRNAs acting as effectors of antiviral host response in MDMs. The miR-221 and miR-22 were both upregulated in MDMs, and they were shown to target the 3′UTR of CD4 transcript and suppress both the mRNA and protein levels, thereby impairing HIV-1 entry into MDMs.
Astrocyte is another cell type reported to be non-productively infected by HIV-1, in a CD4-independent manner. Recently, elevated levels of the anti-HIV-1 restriction factor sterile alpha motif and histidine/aspartic acid domain-containing protein 1 (SAMHD1) was reported to contribute to the restriction of virus replication in cultured astrocytes [235]. An earlier report by Jin et al. [236] had demonstrated the downregulation of SAMHD1 levels by miR-181a in a monocytic cell line (THP-1) and T-cell line (Jurkat). Interestingly, the SAMHD1 mRNA was shown to harbor binding sites for two cellular miRNAs—miR-181a and miR-155, both of which were poorly expressed in astrocytes compared to the microglial cells [235]. Accordingly, exogenous overexpression of these two miRNAs led to reduction in SAMHD1 levels and, consequently, enhanced HIV-1 replication in vitro. Intriguingly, and in a notable demonstration of a potential disconnect between miRNA-mediated effects observed in vitro and in vivo, Witwer et al. [237] reported a lack of correlation between the in vivo levels of SAMHD1 and miR-181a or miR-155 in a well-characterized SIV model of HIV-1.
Nathans et al. [180] pursued an integrative approach wherein they performed miRNA target prediction analysis, which revealed that the HIV-1 3′-UTR harbored target binding sites for at least 11 cellular miRNAs, and then followed it up by testing the data by miRNA microarray analysis, which showed that the miR-29a was highly expressed in HIV-infected T-cells. The miR-29a target site was highly conserved in various HIV-1 subtypes and was shown to directly target and prevent the expression of the viral mRNA. Investigations on underlying mechanistic basis revealed that miR-29a promotes the sequestration or association of the viral mRNAs with P-bodies and RISC proteins. Interestingly, physical and functional interactions between the HIV-1 mRNA and the RISC complex effector proteins have been reported by Chable-Bessia et al. [238]. The authors showed that some components of the RISC complex, including GW182, associate with HIV-1 mRNA and block its association with polysomes, thereby impairing viral gene expression. Attempts have also been made to dissect HIV-1 interaction with host miRNAs in a wider array of clinically relevant target cells. Sun et al. [225] performed a comprehensive analysis of any perturbations to the host miRNA expression levels during acute HIV-1 infection of a variety of target cells including CD4+CD8 PBMCs, CEM, and Jurkat. The authors neither observed any HIV-induced global miRNA deregulation nor could they recapitulate the findings from some miRNA profiling studies reported earlier [181,226]. Interestingly, while these earlier studies had used small RNA-enriched samples for miRNA profiling analysis, Sun et al. [225] relied on total RNA samples instead. Even though downregulation of miR-29 family members were observed [180,220], the authors contended that any suppressive effect by the miRNA is attenuated by the secondary structure of the viral LTR target sequence. A recent report by Adoro et al. [239] provides an alternative conceptual framework to decode the anti-HIV activity of miR-29. In experiments targeted towards deciphering the mechanistic details underlying the IL-21-mediated inhibition of HIV-1 replication, the authors showed that IL-21 promotes miR-29 biogenesis in a STAT3-dependent fashion and reverses the HIV-induced downregulation of miR-29.
Intriguingly, besides the cellular miRNAs, the components of the cellular machinery involved in the biogenesis and execution of the miRNA-mediated gene silencing have also been shown to suppress HIV-1 replication [240]. In a key study, Triboulet et al. [181] reported that knockdown of the endogenous Dicer resulted in enhanced HIV-1 replication and virus production. Nathans et al. [180] reported that depleting Dicer or Drosha in the producer cell enhanced HIV-1 production, whereas depleting either of those two enzymes in the target cell increased viral infectivity. However, in a recent study, Bogerd et al. [199] demonstrated that the replication of many viruses including HIV-1 was not enhanced in human cell lines that lack a functional Dicer enzyme and therefore deficient in the generation of miRNAs and siRNAs; these findings were recently corroborated by Eckenfelder et al. [241]. Direct interaction of miRISC protein components with the HIV-1 RNAs, in a miRNA-independent manner, has also been reported [180,238,241,242]. In HIV-1-infected CD4+ T-cells, the AGO2 protein was shown to directly and preferentially bind, in a miRNA- and Dicer-independent manner, near splice donor sites in the unspliced viral mRNA [241]. This AGO2-viral mRNA interaction impaired the production of spliced HIV-1 mRNAs, thereby implicating the AGO2 protein in HIV-1 mRNA splicing events. Interestingly, Vongrad et al. [243] reported an absence of such interaction between the HIV-1 RNA and AGO2 in monocyte-derived macrophages (MDM), suggesting that HIV-1 is not targeted by the canonical RNAi pathway in MDMs.
Inhibition of HIV-1 replication by cellular miRNAs via non-canonical (i.e., gene silencing-independent) pathways has been reported. Overexpression of miR-146a or miR-888—two exogenous miRNAs neither present in the 293T cells nor possess any known target sites on HIV-1 mRNAs—impaired HIV-1 production but did not alter intracellular Gag protein expression, thereby implicating an RNAi-independent mechanism [244]. Super resolution microscopy analyses revealed that the overexpressed miRNA competes with viral RNA for Gag binding, and the aberrantly assembled RNA-protein complexes at the PM were endocytosed in a dynamin-dependent manner into large intracellular vacuoles. Interestingly, coexpressing target mRNAs of these exogenous miRNAs relieved the inhibitory effect on Gag assembly and virus production. Further, knockdown of AGO2 induced a phenotype comparable to that caused by the exogenous miRNAs. These observations suggest that endogenous miRNAs when unbound to miRISC might bind Gag disrupting its assembly and virus production. Importantly, knockdown of Dicer or Drosha reduced the numbers of Gag-enriched intracellular compartments and promoted robust virus particle production. Taken together, these findings provide a potential mechanistic framework for explaining the inhibitory effects exerted by endogenous Dicer and Drosha on HIV-1 replication. Importantly, these findings also provide a cautionary note on the physiological relevance and interpretation of data obtained from miRNA overexpression experiments to test potential antiviral effect, i.e., any gene silencing-independent effects of candidate cellular miRNAs must be accounted for before conferring functional relevance to miRNA-mediated antiviral activity.

7. Does HIV-1 Encode miRNAs?

Like their hosts, increasing numbers of viruses are being reported to encode miRNAs. Majority of the viral miRNAs discovered so far are encoded by DNA viruses that replicate and assemble in the host cell nucleus, including Herpesviruses, polyomaviruses, and adenoviruses (miRBase, Release 21). Notably, over 90% of these miRNAs are conserved between related species and are encoded by Herpesviridae family that include Epstein–Barr virus (EBV), Kaposi’s sarcoma-associated herpesvirus (KSHV), Herpes simplex virus (HSV), and Human cytomegalovirus (HCMV). These viral miRNAs have been shown to regulate viral and host gene expression that, by modulating metabolic processes and/or the host immune response, generally promote viral replication and, in certain cases, viral latency.
The evidence that RNA viruses encode miRNAs is highly controversial [222,245,246,247,248,249]. The reservations against the feasibility of RNA viruses encoding miRNAs mostly invoke the distinct subcellular locations of the virus replication cycle and miRNA biogenesis pathway. In addition, potential cleavage of the viral genomic RNA in cis by the miRNA processing machinery poses a tangible bottleneck limiting productive virus replication. For example, the first step of the canonical miRNA biogenesis occurs in nucleus, wherein Drosha-mediated cleavage of pri-miRNA generates the pre-miRNA. Because majority of RNA viruses complete their replication in the cytoplasm, this potentially precludes them from accessing the nuclear-localized miRNA biogenesis machinery to generate their own miRNAs. Additionally, the endonucleolytic cleavage of precursor transcripts during miRNA-generating pathways might adversely target the viral genomic RNA or mRNAs. Nevertheless, to date, three retroviruses—bovine leukemia virus (BLV), simian foamy virus of the African green monkey (SFVagm), and the bovine foamy virus (BFV)—have been shown to circumvent the canonical nuclear Drosha-mediated RNA processing step. These viruses harbor Pol III promoters in their genome and accordingly employ a non-canonical strategy wherein RNA polymerase III, instead of RNA Pol II, are usurped to transcribe viral sub-genomic transcripts harboring pre-miRNA hairpins, which are then processed into mature miRNAs in Drosha-independent manner [250,251,252,253,254]. It should be noted that these studies were preceded by an important report showing that the murine herpesvirus 68 encodes several miRNAs that are products of transcription by Pol III [255]. The BLV’s reliance on Pol III to express miRNAs has been proposed to maintain its Pol II-transcribed coding genes silent and thereby evade the cellular immune surveillance. While BLV replication proceeds despite weak viral transcription levels [256,257], robustly generated BLV miRNAs reportedly constitute up to 40% of total cellular miRNAs [251]. Further, at least one of the BLV-encoded miRNAs appears to share common traits with miRNAs encoded by some DNA viruses. Comparable to the KSHV- and MDV-encoded miRNAs that function as viral analogs of the cellular oncomir miR-155, the BLV-encoded miR-B4 and the cellular oncomir miR-29a share an identical seed sequence, and both contribute to oncogenesis by targeting the same cellular mRNAs. The SFVagm reportedly encodes six miRNAs transcribed from its LTR region by Pol III. The seed sequences of two of these miRNAs, SFVagm-miR-S4-3p and SFVagm-miR-S6-3p, are homologous to the cellular miR-155 and innate immunity suppressor miR-132, respectively. In the case of BFV, it employs an unusual strategy wherein Pol III transcribes a single ~122-nucleotides-long pri-miRNA, which is then cleaved into two pre-miRNAs that are further processed to yield three different mature miRNAs. The first retroviral miRNA identified as one generated via the canonical miRNA biogenesis pathway (using Pol II) was transcribed from the viral E-element/exogenous virus-specific region (XSR) of the retrovirus avian leucosis virus subgroup J (ALV-J) [258]. However, the biological function of this E/XSR-miRNA in the ALV-J-associated myeloid leukosis is yet to be deciphered. It also remains to be seen if retroviruses encode miRNA to modulate the host machinery for productive replication and transmission.
The notion that HIV-1 may encode miRNAs stemmed from the recognition that the viral genome harbored secondary structures including stem-loops that mirror the substrates of the cellular miRNA biogenesis machinery. Earliest investigations along that line on HIV-encoded candidate miRNAs relied on computational approaches. Bennasser et al. [259] reported that stem-loop structures located in five regions of the HIV-1 genomic RNA—trans-activation response (TAR) element, Gag-CA, Gag-Pol frameshift, Nef, and 3′-LTR—are potential miRNA-yielding hotspots encoding up to 10 miRNAs. Couturier and Root-Bernstein [260] probed for sequence complementarity between the proviral genome and the mRNA of target cells, which led to the prediction that HIV-1 might encode miRNAs capable of blocking the production of CD4, CD28 and certain interleukins. Omoto et al. [261] reported that HIV-1 encodes miR-N367 from a 70-nucleotide-long structure in the Nef/LTR overlapping region. The miR-N367 was subsequently shown to target the negative response element of the LTR U3 region and inhibit the LTR’s promoter activity [262]. However, miR-N367 was not picked up in screens for HIV-encoded miRNAs in infected HeLa cells [255] and in chronically-infected T-cells [245]. Another reportedly HIV-encoded miRNA miR-H1 is derived from an 81-nucleotide stem-loop structure located downstream of the two NF-kB-binding sites in the viral LTR, and was shown to down-regulate the expression of the apoptosis antagonizing transcription factor (AATF) by cleaving its 12th exon [263]. Intriguingly, Lamers et al. [264] reported substantial sequence variability of both the pre-miRNA and mature miR-H1 in over 1000 HIV-1 sequences derived from post-mortem tissue samples of several HIV-1 patients. The Dicer enzyme was reported to bind and cleave the HIV-1 TAR stem-loop structure, and the resulting ~22-nucleotides-long small RNA species, termed TAR-derived miRNA were proposed to promote viral latency by chromatin remodeling of the LTR [265] and to preclude death of HIV-infected cells by downregulating the proapoptotic cellular genes excision repair cross complementation group 1 (ERCC1) and immediate early response 3 (IER3) [266]. Ouellet et al. [267] reported that asymmetrical processing of the HIV-1 TAR element by Dicer yielded two additional miRNAs, namely miR-TAR-5p and miR-TAR-3p, in HIV-infected cell lines and primary CD4+ T-cells. In a follow-up study, the TAR miRNAs were shown to target host genes including caspase 8 (CASP8), Ikaros family Zinc Finger 3 (Aiolos), Ikaros family Zinc Finger 1 (IKZF1), and nucleophosmin 1 (NPM), thereby preventing apoptosis of HIV-infected cells and promoting cell survival and viral replication [268]. Interestingly, exosomes from HIV-infected cells and patient sera harbored TAR RNAs (precursor of TAR miRNAs), and these exosomes down-regulated BCL2 Like 11 (Bim) and CDK9 protein levels and thereby apoptosis in recipient cells [269]. The RT region of the HIV-1 genome was shown by Zhang et al. [270] to encode a miRNA named miR-H3 that targets the TATA box in the viral 5′-LTR promoter. While miR-H3 expression upregulated viral RNA transcription and protein production, mutations in the miR-H3 encoding region impaired HIV-1 replication.
Next-generation sequencing technologies are increasingly being used to validate the reported HIV-1 miRNAs and to address the contentious question of whether HIV-1 is capable of encoding miRNAs. Schopman et al. [271] used sequencing by oligonucleotide ligation and detection (SOLiD) technology-based deep sequencing approach to screen for small RNAs in HIV-infected SupT1 cells. Approximately one percent of the total small RNAs identified mapped to the HIV-1 RNA genome. A majority of these ~18–21-nucleotides-long candidate viral-encoded miRNAs aligned with the viral positive sense RNA, while a small segment corresponded to the viral negative sense RNA. In contrast, a deep sequencing-based analysis using two different HIV-infected cell lines (TZM-bl and CD8166) and two types of primary human cells (CD4+ PBMCs and macrophages) did not uncover any HIV-encoded miRNAs [253]. These authors contend that the candidate HIV-specific small RNAs fail to satisfy three criteria for authentic viral miRNAs. First, all other viral miRNAs reported to date are generally derived from fewer locations in their respective genome. However, the HIV-specific small RNAs recovered by Whisnant et al. [253] correspond to the entire proviral genome. Further, unlike in the case of other viral miRNAs, only a minor population (<~1%) of the small RNAs in HIV-infected cells aligned with the viral genome. Second, miRNAs typically range in size from ~20–24 nucleotides in length. However, the small RNAs that align with the HIV-1 genome exhibited a random size distribution with a majority <~20 nucleotides in length. Third, the authors contend that the TAR stem, source of several miRNAs reported, lacks several key defining characteristics of canonical pri-miRNA stem loops. The TAR stem is only ~24 bp long, the terminal loop is just six nucleotides, and it is located at the 5′-end of viral mRNAs. In contrast, the canonical pri-miRNA stem is ~33 bp long, the terminal loop is over 10 nucleotides, and both the 5′- and 3′-ends of the stem is flanked by unstructured RNA sequences. Further, comparing cellular miRNA expression profiles of HIV-infected CD4+ PBMCs 72 h post infection, a time point well beyond the reported half-life (~24 h) of virus-infected T-cells in vivo, with that of uninfected control cells revealed no difference (greater than two-fold). To determine whether HIV-1 has successfully evolved out potentially detrimental miRNA-binding sites on its genome, photoactivatable ribonucleoside-enhanced cross-linking and immunoprecipitation (PAR-CLIP) technique was used to determine miRNA-directed RISC-binding sites on viral genome in infected non-target cell types. Unexpectedly, despite making up 10–50% of the total mRNA population in these infected cells, only ~0.3% of the viral transcripts harbored RISC-binding sites. Among these, only a few cellular miRNAs including miR-423, miR-301a, and miR-155 bound viral transcripts and reduced their expression. The authors proposed that evolutionary constraints posed by cellular miRNAs with antiviral properties likely forced the viral RNA to adopt extensive secondary structures that mask miRNA-binding sites. Considering all their findings, the authors concluded that HIV-1 neither encodes miRNAs nor strongly influences cellular miRNA expression, at least early after infection. Vongrad et al. [243] came to similar conclusions when they failed to detect incorporation of HIV-1-derived sncRNAs or any viral target sequences into the AGO2-RISC complex.

8. Concluding Remarks and Outstanding Questions

Extraordinary progress in deciphering the biology of cellular miRNAs, since their discovery 25 years back [272], offers a great opportunity to understand the role of these sncRNAs in virus-host interactions. Accumulating evidence suggest that cellular miRNAs play important roles in viral life cycle by promoting or inhibiting virus replication. Accordingly, cellular miRNAs that can directly inhibit HIV-1 replication by targeting the vRNA and/or viral mRNAs, and cellular miRNAs that can indirectly affect HIV-1 infection by targeting HDFs have been reported (Table 2). However, the efficacy and the physiological relevance of any such endogenous cellular miRNA-mediated effects on HIV-1 replication, and the ability of HIV-1 genome to encode miRNAs remain contentious, albeit enriching and important, areas of ongoing research [273]. Several outstanding questions need special attention to better understand the role of miRNAs in HIV-1 infection.
Are the copy numbers of cellular anti-HIV miRNAs in target cells sufficient to exert a consequential effect on the HIV-1 replication? At least 100 copies of miRNA is reportedly required to effectively downregulate a cellular target RNA [274,275]. This ratio is likely to be far greater when the target is vRNA or viral mRNA. A single HIV-infected target cell is estimated to produce over 10,000 progeny virus particles within one to two days [276], which would require the generation of at least 20,000 vRNA molecules for encapsidation purpose alone, not taking into account the additional number of viral mRNAs necessary for the synthesis of viral proteins (Figure 3). Indeed, analysis by Whisnant et al. [222] revealed the presence of around 20,000 viral transcripts per HIV-1-infected T-cell. Further, miRISC binding to HIV-1 transcripts were reported to be 100-fold less efficient than miRNA binding to cellular mRNAs in HIV-1-infected cells [222]. Interestingly, engineering of artificial miRNA target sites into the HIV-1 RNA led to up to 40-fold increase in the packaging of the cognate miRNAs into the virus particle and inhibited virion infectivity [277]. Further, miRISC binding to HIV-1 transcripts were reported to be 100-fold less efficient than miRISC binding to cellular mRNAs in HIV-1-infected cells [222]. Interestingly, engineering of artificial miRNA target sites into the HIV-1 RNA led to ~40-fold increase in the packaging of the cognate miRNAs into the virus particle and inhibited virion infectivity [276]. Compared to the investigations on determining whether HIV-1 infection alters cellular miRNA signatures, information on the levels of the cellular miRNAs reported to alter HIV-1 replication is very limited. A potential contributing factor is the prevalent practice of reporting miRNA levels from qPCR-based assays as fold-change. Instead, assessing and reporting miRNA induction based on miRNA copy numbers might preclude inadvertent implication of physiologically and/or pathologically irrelevant miRNAs in HIV-1 infection. Further, due to partial complementarity of the miRNA: mRNA interaction, a cellular miRNA typically downregulates the production of a target protein by less than two-fold [278,279]. Hence, even if the expression level of an endogenous candidate anti-HIV miRNA reaches such threshold, and not considering the turnover rate of that miRNA [280], it remains unclear whether the projected 50% reduction in HIV-1 gene expression could negatively impact the virus replication and/or pathogenesis in vivo, in any consequential manner. Expression of endogenous anti-HIV miRNAs sufficient to achieve a complete repression of the viral RNAs has generally not been demonstrated in infected target cells. Also, the kinetics of the miRNA-mediated activity appears not optimal for acute response to virus infection [281].
Do HIV-1 RNAs/mRNAs function as molecular sponges of human miRNAs, thereby relieving translational repression of specific cellular cofactors? Virus replication in target cell produces vRNAs and viral mRNAs for encapsidation and/or generation of viral proteins. The “miRNA sponges” are exogenously-expressed artificial RNA transcripts harboring binding sites for specific target miRNAs, and the resulting competitive inhibition (sponging off) of the miRNA can lead to the derepression of its target cellular mRNAs [282,283]. The occurrence of endogenous miRNA sponges (competitive endogenous RNA, ceRNA) and their ability to effectively derepress the target mRNAs in vivo is an equally exciting and contentious area of ongoing research [284,285,286]. However, this has led to the proposal that viral RNAs, harboring shared miRNA binding sites with cellular mRNAs, can compete and sequester the cognate cellular miRNAs, thereby selectively altering the stability and translational efficiency of the target cellular mRNAs in the infected cells. It has been speculated that viruses can employ such “miRNA sponge” effect to counter cellular miRNAs towards establishing a productive infection and promote viral pathogenesis. Indeed, the Herpesvirus saimiri (HSV)-derived non-coding transcripts called H. saimiri U-rich RNAs (HSURs) were reported to interact with and downregulate the cellular miR-27, thereby altering the expression of its target genes [287]. Similar virus-directed miRNA sponge effects have been reported to occur in infections by other viruses including HBV [172,288], murine cytomegalovirus (MCMV) [289,290], HCMV [291], HCV [170], and human papillomaviruses (HPV). However, HIV-1 has not yet been demonstrated to employ its transcripts as miRNA sponges to modulate the expression of host genes.
Does HIV-1 adapt and evolve resistance to anti-viral miRNAs? Drug-resistant mutant HIV-1 strains continue to pose a significant challenge to the cART and management of HIV/AIDS. So, one would presume that HIV-1 would eventually acquire mutations in its vRNA genome to evade the sequence-specific repression by anti-HIV cellular miRNAs. RNAi-mediated degradation of vRNA genome has been employed to inhibit replication of many viruses including HIV-1; specifically, the tat, rev and gag coding regions of HIV-1 RNA have been targeted by RNAi to inhibit virus replication [292,293,294,295,296,297,298,299,300,301,302,303,304]. Despite the RNAi exhibiting potent antiviral activity towards HIV-1, the virus has been shown to eventually escape the RNAi-mediated repression by acquiring mutations within the siRNA-target sequence [297,299,300,304]. Indeed, HIV-1 has been shown even to counter combinatorial RNAi approaches by acquiring mutations in multiple siRNA-target sites [304,305].
Is undue reliance on miRNA target prediction programs contributing to reporting of miRNA-mRNA interactions that are probably physiologically irrelevant in the context of HIV-1 replication? Existing computational methods used to predict miRNA targets generally rely on sequence similarity, including perfect Watson–Crick base pairing between the seed sequence of the miRNA and the target mRNA [107]. However, the occurrence of a complementary seed sequence in the target mRNA does not by default translate into effective miRNA binding [143,306]. Accordingly, the false positive rates of the popular prediction algorithms range between 20–50% [278,307]. Conversely, the inability of the miRNA target prediction tools to recognize non-canonical interactions between miRNAs and their mRNA targets [308] contribute to equally non-productive false negatives. Furthermore, the local 3D structure on the target mRNA, especially in the case of highly structured HIV-1 RNA [53], can significantly impact the accessibility of the miRNA-binding site to the miRISC and thus the outcome of the miRNA-mRNA interaction [144,309,310]. In other words, even the best available prediction algorithms may not be modeling what really happens inside an HIV-infected target cell.
Collectively, information on any functional interplay between HIV-1 and cellular miRNA pathways provides a significant and timely opportunity to reveal new insights into virus–host interactions governing the virus replication and viral pathogenesis. The potential discoveries hold great promise for developing novel anti-HIV-1 therapies [46]. However, such promises can only be materialized by systematically addressing the key knowledge gaps highlighted above.

Acknowledgments

This work was partly supported by National Institutes of Health (NIH) Grants DA024558, DA30896, DA033892, DA021471, AI22960 and MD007586 (to Chandravanu Dash) and DA037779 (to Jui Pandhare). The work was also supported by the RCMI Grant G12MD007586, the Vanderbilt CTSA Grant UL1RR024975, the Meharry Translational Research Center (MeTRC) CTSA grant (U54 RR026140 from NCRR/NIH, the U54 Grant MD007593 from NIMHD/NIH, and the Tennessee Center for AIDS Research (P30 AI110527).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Swanstrom, R.; Coffin, J. HIV-1 pathogenesis: The virus. Cold Spring Harb. Perspect. Med. 2012, 2, a007443. [Google Scholar] [CrossRef] [PubMed]
  2. Walker, B.; McMichael, A. The T-cell response to HIV. Cold Spring Harb. Perspect. Med. 2012, 2, a007054. [Google Scholar] [CrossRef] [PubMed]
  3. Overbaugh, J.; Morris, L. The Antibody Response against HIV-1. Cold Spring Harb. Perspect. Med. 2012, 2, a007039. [Google Scholar] [CrossRef] [PubMed]
  4. Sharp, P.M.; Hahn, B.H. Origins of HIV and the AIDS pandemic. Cold Spring Harb. Perspect. Med. 2011, 1, a006841. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Faria, N.R.; Rambaut, A.; Suchard, M.A.; Baele, G.; Bedford, T.; Ward, M.J.; Tatem, A.J.; Sousa, J.D.; Arinaminpathy, N.; Pepin, J.; et al. HIV epidemiology. The early spread and epidemic ignition of HIV-1 in human populations. Science 2014, 346, 56–61. [Google Scholar] [CrossRef] [PubMed]
  6. Worobey, M.; Watts, T.D.; McKay, R.A.; Suchard, M.A.; Granade, T.; Teuwen, D.E.; Koblin, B.A.; Heneine, W.; Lemey, P.; Jaffe, H.W. 1970s and ‘Patient 0’ HIV-1 genomes illuminate early HIV/AIDS history in North America. Nature 2016, 539, 98–101. [Google Scholar] [CrossRef] [PubMed]
  7. Picker, L.J.; Hansen, S.G.; Lifson, J.D. New paradigms for HIV/AIDS vaccine development. Annu. Rev. Med. 2012, 63, 95–111. [Google Scholar] [CrossRef] [PubMed]
  8. Stephenson, K.E.; D’Couto, H.T.; Barouch, D.H. New concepts in HIV-1 vaccine development. Curr. Opin. Immunol. 2016, 41, 39–46. [Google Scholar] [CrossRef] [PubMed]
  9. Haynes, B.F.; Burton, D.R. Developing an HIV vaccine. Science 2017, 355, 1129–1130. [Google Scholar] [CrossRef] [PubMed]
  10. Arts, E.J.; Hazuda, D.J. HIV-1 antiretroviral drug therapy. Cold Spring Harb. Perspect. Med. 2012, 2, a007161. [Google Scholar] [CrossRef] [PubMed]
  11. Hu, W.S.; Hughes, S.H. HIV-1 reverse transcription. Cold Spring Harb. Perspect. Med. 2012, 2, a006882. [Google Scholar] [CrossRef] [PubMed]
  12. Cuevas, J.M.; Geller, R.; Garijo, R.; Lopez-Aldeguer, J.; Sanjuan, R. Extremely High Mutation Rate of HIV-1 In Vivo. PLoS Biol. 2015, 13, e1002251. [Google Scholar] [CrossRef] [PubMed]
  13. Laskey, S.B.; Siliciano, R.F. A mechanistic theory to explain the efficacy of antiretroviral therapy. Nat. Rev. Microbiol. 2014, 12, 772–780. [Google Scholar] [CrossRef] [PubMed]
  14. Finzi, D.; Hermankova, M.; Pierson, T.; Carruth, L.M.; Buck, C.; Chaisson, R.E.; Quinn, T.C.; Chadwick, K.; Margolick, J.; Brookmeyer, R.; et al. Identification of a reservoir for HIV-1 in patients on highly active antiretroviral therapy. Science 1997, 278, 1295–1300. [Google Scholar] [CrossRef] [PubMed]
  15. Chun, T.W.; Stuyver, L.; Mizell, S.B.; Ehler, L.A.; Mican, J.A.; Baseler, M.; Lloyd, A.L.; Nowak, M.A.; Fauci, A.S. Presence of an inducible HIV-1 latent reservoir during highly active antiretroviral therapy. Proc. Natl. Acad. Sci. USA 1997, 94, 13193–13197. [Google Scholar] [CrossRef] [PubMed]
  16. Finzi, D.; Blankson, J.; Siliciano, J.D.; Margolick, J.B.; Chadwick, K.; Pierson, T.; Smith, K.; Lisziewicz, J.; Lori, F.; Flexner, C.; et al. Latent infection of CD4+ T cells provides a mechanism for lifelong persistence of HIV-1, even in patients on effective combination therapy. Nat. Med. 1999, 5, 512–517. [Google Scholar] [CrossRef] [PubMed]
  17. Davey, R.T., Jr.; Bhat, N.; Yoder, C.; Chun, T.W.; Metcalf, J.A.; Dewar, R.; Natarajan, V.; Lempicki, R.A.; Adelsberger, J.W.; Miller, K.D.; et al. HIV-1 and T cell dynamics after interruption of highly active antiretroviral therapy (HAART) in patients with a history of sustained viral suppression. Proc. Natl. Acad. Sci. USA 1999, 96, 15109–15114. [Google Scholar] [CrossRef] [PubMed]
  18. Pierson, T.; McArthur, J.; Siliciano, R.F. Reservoirs for HIV-1: Mechanisms for viral persistence in the presence of antiviral immune responses and antiretroviral therapy. Annu. Rev. Immunol. 2000, 18, 665–708. [Google Scholar] [CrossRef] [PubMed]
  19. Blankson, J.N.; Persaud, D.; Siliciano, R.F. The challenge of viral reservoirs in HIV-1 infection. Annu. Rev. Med. 2002, 53, 557–593. [Google Scholar] [CrossRef] [PubMed]
  20. Siliciano, J.D.; Kajdas, J.; Finzi, D.; Quinn, T.C.; Chadwick, K.; Margolick, J.B.; Kovacs, C.; Gange, S.J.; Siliciano, R.F. Long-term follow-up studies confirm the stability of the latent reservoir for HIV-1 in resting CD4+ T cells. Nat. Med. 2003, 9, 727–728. [Google Scholar] [CrossRef] [PubMed]
  21. Trono, D.; Van Lint, C.; Rouzioux, C.; Verdin, E.; Barre-Sinoussi, F.; Chun, T.W.; Chomont, N. HIV persistence and the prospect of long-term drug-free remissions for HIV-infected individuals. Science 2010, 329, 174–180. [Google Scholar] [CrossRef] [PubMed]
  22. Siliciano, R.F.; Greene, W.C. HIV latency. Cold Spring Harb. Perspect. Med. 2011, 1, a007096. [Google Scholar] [CrossRef] [PubMed]
  23. Barre-Sinoussi, F.; Ross, A.L.; Delfraissy, J.F. Past, present and future: 30 years of HIV research. Nat. Rev. Microbiol. 2013, 11, 877–883. [Google Scholar] [CrossRef] [PubMed]
  24. Churchill, M.J.; Deeks, S.G.; Margolis, D.M.; Siliciano, R.F.; Swanstrom, R. HIV reservoirs: What, where and how to target them. Nat. Rev. Microbiol. 2016, 14, 55–60. [Google Scholar] [CrossRef] [PubMed]
  25. Gupta, R.K.; Jordan, M.R.; Sultan, B.J.; Hill, A.; Davis, D.H.; Gregson, J.; Sawyer, A.W.; Hamers, R.L.; Ndembi, N.; Pillay, D.; et al. Global trends in antiretroviral resistance in treatment-naive individuals with HIV after rollout of antiretroviral treatment in resource-limited settings: A global collaborative study and meta-regression analysis. Lancet 2012, 380, 1250–1258. [Google Scholar] [CrossRef]
  26. Stadeli, K.M.; Richman, D.D. Rates of emergence of HIV drug resistance in resource-limited settings: A systematic review. Antivir. Ther. 2013, 18, 115–123. [Google Scholar] [CrossRef] [PubMed]
  27. Clutter, D.S.; Jordan, M.R.; Bertagnolio, S.; Shafer, R.W. HIV-1 drug resistance and resistance testing. Infect Genet. Evol. 2016, 46, 292–307. [Google Scholar] [CrossRef] [PubMed]
  28. Gupta, R.K.; Gregson, J.; Parkin, N.; Haile-Selassie, H.; Tanuri, A.; Andrade Forero, L.; Kaleebu, P.; Watera, C.; Aghokeng, A.; Mutenda, N.; et al. HIV-1 drug resistance before initiation or re-initiation of first-line antiretroviral therapy in low-income and middle-income countries: A systematic review and meta-regression analysis. Lancet Infect. Dis. 2017, 18, 346–355. [Google Scholar] [CrossRef]
  29. Archin, N.M.; Sung, J.M.; Garrido, C.; Soriano-Sarabia, N.; Margolis, D.M. Eradicating HIV-1 infection: Seeking to clear a persistent pathogen. Nat. Rev. Microbiol. 2014, 12, 750–764. [Google Scholar] [CrossRef] [PubMed]
  30. O’Connell, K.A.; Bailey, J.R.; Blankson, J.N. Elucidating the elite: Mechanisms of control in HIV-1 infection. Trends Pharmacol. Sci. 2009, 30, 631–637. [Google Scholar] [CrossRef] [PubMed]
  31. Coffin, J.; Swanstrom, R. HIV pathogenesis: Dynamics and genetics of viral populations and infected cells. Cold Spring Harb. Perspect. Med. 2013, 3, a012526. [Google Scholar] [CrossRef] [PubMed]
  32. Lackner, A.A.; Lederman, M.M.; Rodriguez, B. HIV pathogenesis: The host. Cold Spring Harb. Perspect. Med. 2012, 2, a007005. [Google Scholar] [CrossRef] [PubMed]
  33. Telenti, A.; Johnson, W.E. Host genes important to HIV replication and evolution. Cold Spring Harb. Perspect. Med. 2012, 2, a007203. [Google Scholar] [CrossRef] [PubMed]
  34. Carrington, M.; Alter, G. Innate immune control of HIV. Cold Spring Harb. Perspect. Med. 2012, 2, a007070. [Google Scholar] [CrossRef] [PubMed]
  35. Fraser, C.; Lythgoe, K.; Leventhal, G.E.; Shirreff, G.; Hollingsworth, T.D.; Alizon, S.; Bonhoeffer, S. Virulence and pathogenesis of HIV-1 infection: An evolutionary perspective. Science 2014, 343, 1243727. [Google Scholar] [CrossRef] [PubMed]
  36. Simon, V.; Bloch, N.; Landau, N.R. Intrinsic host restrictions to HIV-1 and mechanisms of viral escape. Nat. Immunol. 2015, 16, 546–553. [Google Scholar] [CrossRef] [PubMed]
  37. Boritz, E.A.; Darko, S.; Swaszek, L.; Wolf, G.; Wells, D.; Wu, X.; Henry, A.R.; Laboune, F.; Hu, J.; Ambrozak, D.; et al. Multiple Origins of Virus Persistence during Natural Control of HIV Infection. Cell 2016, 166, 1004–1015. [Google Scholar] [CrossRef] [PubMed]
  38. Hoxie, J.A.; June, C.H. Novel cell and gene therapies for HIV. Cold Spring Harb. Perspect. Med. 2012, 2, a007179. [Google Scholar] [CrossRef] [PubMed]
  39. Haynes, B.F.; Shaw, G.M.; Korber, B.; Kelsoe, G.; Sodroski, J.; Hahn, B.H.; Borrow, P.; McMichael, A.J. HIV-Host Interactions: Implications for Vaccine Design. Cell Host Microbe 2016, 19, 292–303. [Google Scholar] [CrossRef] [PubMed]
  40. Skalsky, R.L.; Cullen, B.R. Viruses, microRNAs, and host interactions. Annu. Rev. Microbiol. 2010, 64, 123–141. [Google Scholar] [CrossRef] [PubMed]
  41. Cullen, B.R. Viruses and microRNAs: RISCy interactions with serious consequences. Genes Dev. 2011, 25, 1881–1894. [Google Scholar] [CrossRef] [PubMed]
  42. Chiang, K.; Rice, A.P. Mini ways to stop a virus: MicroRNAs and HIV-1 replication. Future Virol. 2011, 6, 209–221. [Google Scholar] [CrossRef] [PubMed]
  43. Klase, Z.; Houzet, L.; Jeang, K.T. MicroRNAs and HIV-1: Complex interactions. J. Biol. Chem. 2012, 287, 40884–40890. [Google Scholar] [CrossRef] [PubMed]
  44. Swaminathan, S.; Murray, D.D.; Kelleher, A.D. miRNAs and HIV: Unforeseen determinants of host-pathogen interaction. Immunol. Rev. 2013, 254, 265–280. [Google Scholar] [CrossRef] [PubMed]
  45. Swaminathan, G.; Navas-Martin, S.; Martin-Garcia, J. MicroRNAs and HIV-1 infection: Antiviral activities and beyond. J. Mol. Biol. 2014, 426, 1178–1197. [Google Scholar] [CrossRef] [PubMed]
  46. tenOever, B.R. RNA viruses and the host microRNA machinery. Nat. Rev. Microbiol. 2013, 11, 169–180. [Google Scholar] [CrossRef] [PubMed]
  47. Guo, Y.E.; Steitz, J.A. Virus meets host microRNA: The destroyer, the booster, the hijacker. Mol. Cell. Biol. 2014, 34, 3780–3787. [Google Scholar] [CrossRef] [PubMed]
  48. Trobaugh, D.W.; Klimstra, W.B. MicroRNA Regulation of RNA Virus Replication and Pathogenesis. Trends Mol. Med. 2017, 23, 80–93. [Google Scholar] [CrossRef] [PubMed]
  49. Harwig, A.; Das, A.T.; Berkhout, B. HIV-1 RNAs: Sense and antisense, large mRNAs and small siRNAs and miRNAs. Curr. Opin. HIV AIDS 2015, 10, 103–109. [Google Scholar] [CrossRef] [PubMed]
  50. Freed, E.O.; Martin, M.A. Human immundeficiency viruses: Replication. In Fields Virology, 6th ed.; Lippincott, Williams and Wilkins: Philadelphia, PA, USA, 2013. [Google Scholar]
  51. Engelman, A.; Cherepanov, P. The structural biology of HIV-1: Mechanistic and therapeutic insights. Nat. Rev. Microbiol. 2012, 10, 279–290. [Google Scholar] [CrossRef] [PubMed]
  52. Coffin, J.M. Retroviruses; Cold Spring Harbor Laboratory Press: Cold Spring Harbor, NY, USA, 1997. [Google Scholar]
  53. Watts, J.M.; Dang, K.K.; Gorelick, R.J.; Leonard, C.W.; Bess, J.W., Jr.; Swanstrom, R.; Burch, C.L.; Weeks, K.M. Architecture and secondary structure of an entire HIV-1 RNA genome. Nature 2009, 460, 711–716. [Google Scholar] [CrossRef] [PubMed]
  54. Ott, D.E. HIV-1 biology at the protein level. In HIV-1 Proteomics: From Discovery to Clinical Application; Springer-Verlag: New York, NY, USA, 2016. [Google Scholar]
  55. Bushman, F.D.; Malani, N.; Fernandes, J.; D’Orso, I.; Cagney, G.; Diamond, T.L.; Zhou, H.; Hazuda, D.J.; Espeseth, A.S.; Konig, R.; et al. Host cell factors in HIV replication: Meta-analysis of genome-wide studies. PLoS Pathog. 2009, 5, e1000437. [Google Scholar] [CrossRef] [PubMed]
  56. Yamashita, M.; Engelman, A.N. Capsid-Dependent Host Factors in HIV-1 Infection. Trends Microbiol. 2017, 25, 741–755. [Google Scholar] [CrossRef] [PubMed]
  57. Malim, M.H.; Bieniasz, P.D. HIV Restriction Factors and Mechanisms of Evasion. Cold Spring Harb. Perspect. Med. 2012, 2, a006940. [Google Scholar] [CrossRef] [PubMed]
  58. Harris, R.S.; Hultquist, J.F.; Evans, D.T. The restriction factors of human immunodeficiency virus. J. Biol. Chem. 2012, 287, 40875–40883. [Google Scholar] [CrossRef] [PubMed]
  59. Wilen, C.B.; Tilton, J.C.; Doms, R.W. HIV: Cell binding and entry. Cold Spring Harb. Perspect. Med. 2012, 2, a006866. [Google Scholar] [CrossRef] [PubMed]
  60. Zhao, G.; Perilla, J.R.; Yufenyuy, E.L.; Meng, X.; Chen, B.; Ning, J.; Ahn, J.; Gronenborn, A.M.; Schulten, K.; Aiken, C.; et al. Mature HIV-1 capsid structure by cryo-electron microscopy and all-atom molecular dynamics. Nature 2013, 497, 643–646. [Google Scholar] [CrossRef] [PubMed]
  61. Gres, A.T.; Kirby, K.A.; KewalRamani, V.N.; Tanner, J.J.; Pornillos, O.; Sarafianos, S.G. STRUCTURAL VIROLOGY. X-ray crystal structures of native HIV-1 capsid protein reveal conformational variability. Science 2015, 349, 99–103. [Google Scholar] [CrossRef] [PubMed]
  62. Campbell, E.M.; Hope, T.J. HIV-1 capsid: The multifaceted key player in HIV-1 infection. Nat. Rev. Microbiol. 2015, 13, 471–483. [Google Scholar] [CrossRef] [PubMed]
  63. Perilla, J.R.; Gronenborn, A.M. Molecular Architecture of the Retroviral Capsid. Trends Biochem. Sci. 2016, 41, 410–420. [Google Scholar] [CrossRef] [PubMed]
  64. Mattei, S.; Glass, B.; Hagen, W.J.; Krausslich, H.G.; Briggs, J.A. The structure and flexibility of conical HIV-1 capsids determined within intact virions. Science 2016, 354, 1434–1437. [Google Scholar] [CrossRef] [PubMed]
  65. Fassati, A. Multiple roles of the capsid protein in the early steps of HIV-1 infection. Virus Res. 2012, 170, 15–24. [Google Scholar] [CrossRef] [PubMed]
  66. Ambrose, Z.; Aiken, C. HIV-1 uncoating: Connection to nuclear entry and regulation by host proteins. Virology 2014, 454–455, 371–379. [Google Scholar] [CrossRef] [PubMed]
  67. Lee, K.; Ambrose, Z.; Martin, T.D.; Oztop, I.; Mulky, A.; Julias, J.G.; Vandegraaff, N.; Baumann, J.G.; Wang, R.; Yuen, W.; et al. Flexible use of nuclear import pathways by HIV-1. Cell Host Microbe 2010, 7, 221–233. [Google Scholar] [CrossRef] [PubMed]
  68. Craigie, R.; Bushman, F.D. HIV DNA integration. Cold Spring Harb. Perspect. Med. 2012, 2, a006890. [Google Scholar] [CrossRef] [PubMed]
  69. Matreyek, K.A.; Engelman, A. Viral and cellular requirements for the nuclear entry of retroviral preintegration nucleoprotein complexes. Viruses 2013, 5, 2483–2511. [Google Scholar] [CrossRef] [PubMed]
  70. Ruelas, D.S.; Greene, W.C. An integrated overview of HIV-1 latency. Cell 2013, 155, 519–529. [Google Scholar] [CrossRef] [PubMed]
  71. Karn, J.; Stoltzfus, C.M. Transcriptional and posttranscriptional regulation of HIV-1 gene expression. Cold Spring Harb. Perspect. Med. 2012, 2, a006916. [Google Scholar] [CrossRef] [PubMed]
  72. Freed, E.O. HIV-1 assembly, release and maturation. Nat. Rev. Microbiol. 2015, 13, 484–496. [Google Scholar] [CrossRef] [PubMed]
  73. Briggs, J.A.; Krausslich, H.G. The molecular architecture of HIV. J. Mol. Biol. 2011, 410, 491–500. [Google Scholar] [CrossRef] [PubMed]
  74. Ganser-Pornillos, B.K.; Yeager, M.; Pornillos, O. Assembly and architecture of HIV. Adv. Exp. Med. Biol. 2012, 726, 441–465. [Google Scholar] [PubMed]
  75. Schur, F.K.; Hagen, W.J.; Rumlova, M.; Ruml, T.; Muller, B.; Krausslich, H.G.; Briggs, J.A. Structure of the immature HIV-1 capsid in intact virus particles at 8.8 A resolution. Nature 2015, 517, 505–508. [Google Scholar] [CrossRef] [PubMed]
  76. Sundquist, W.I.; Krausslich, H.G. HIV-1 assembly, budding, and maturation. Cold Spring Harb. Perspect. Med. 2012, 2, a006924. [Google Scholar] [CrossRef] [PubMed]
  77. Usami, Y.; Popov, S.; Popova, E.; Inoue, M.; Weissenhorn, W.; Göttlinger, H.G. The ESCRT pathway and HIV-1 budding. Biochem. Soc. Trans. 2009, 37, 181–184. [Google Scholar] [CrossRef] [PubMed]
  78. Votteler, J.; Sundquist, W.I. Virus budding and the ESCRT pathway. Cell Host Microbe 2013, 14, 232–241. [Google Scholar] [CrossRef] [PubMed]
  79. Mattei, S.; Schur, F.K.; Briggs, J.A. Retrovirus maturation-an extraordinary structural transformation. Curr. Opin. Virol. 2016, 18, 27–35. [Google Scholar] [CrossRef] [PubMed]
  80. Eckwahl, M.J.; Arnion, H.; Kharytonchyk, S.; Zang, T.; Bieniasz, P.D.; Telesnitsky, A.; Wolin, S.L. Analysis of the human immunodeficiency virus-1 RNA packageome. RNA 2016, 22, 1228–1238. [Google Scholar] [CrossRef] [PubMed]
  81. Chertova, E.; Chertov, O.; Coren, L.V.; Roser, J.D.; Trubey, C.M.; Bess, J.W., Jr.; Sowder, R.C., II; Barsov, E.; Hood, B.L.; Fisher, R.J.; et al. Proteomic and biochemical analysis of purified human immunodeficiency virus type 1 produced from infected monocyte-derived macrophages. J. Virol. 2006, 80, 9039–9052. [Google Scholar] [CrossRef] [PubMed]
  82. Ott, D.E. Cellular proteins detected in HIV-1. Rev. Med. Virol. 2008, 18, 159–175. [Google Scholar] [CrossRef] [PubMed]
  83. Shaw, G.M.; Hunter, E. HIV transmission. Cold Spring Harb. Perspect. Med. 2012, 2, a006965. [Google Scholar] [CrossRef] [PubMed]
  84. Fackler, O.T.; Murooka, T.T.; Imle, A.; Mempel, T.R. Adding new dimensions: Towards an integrative understanding of HIV-1 spread. Nat. Rev. Microbiol. 2014, 12, 563–574. [Google Scholar] [CrossRef] [PubMed]
  85. Shabalina, S.A.; Koonin, E.V. Origins and evolution of eukaryotic RNA interference. Trends Ecol. Evol. 2008, 23, 578–587. [Google Scholar] [CrossRef] [PubMed]
  86. Wilson, R.C.; Doudna, J.A. Molecular mechanisms of RNA interference. Annu. Rev. Biophys. 2013, 42, 217–239. [Google Scholar] [CrossRef] [PubMed]
  87. Ghildiyal, M.; Zamore, P.D. Small silencing RNAs: An expanding universe. Nat. Rev. Genet. 2009, 10, 94–108. [Google Scholar] [CrossRef] [PubMed]
  88. Huntzinger, E.; Izaurralde, E. Gene silencing by microRNAs: Contributions of translational repression and mRNA decay. Nat. Rev. Genet. 2011, 12, 99–110. [Google Scholar] [CrossRef] [PubMed]
  89. Ishizu, H.; Siomi, H.; Siomi, M.C. Biology of PIWI-interacting RNAs: New insights into biogenesis and function inside and outside of germlines. Genes Dev. 2012, 26, 2361–2373. [Google Scholar] [CrossRef] [PubMed]
  90. Rana, T.M. Illuminating the silence: Understanding the structure and function of small RNAs. Nat. Rev. Mol. Cell Biol. 2007, 8, 23–36. [Google Scholar] [CrossRef] [PubMed]
  91. Steinkraus, B.R.; Toegel, M.; Fulga, T.A. Tiny giants of gene regulation: Experimental strategies for microRNA functional studies. Wiley Interdiscip. Rev. Dev. Biol. 2016, 5, 311–362. [Google Scholar] [CrossRef] [PubMed]
  92. Hutvagner, G.; Simard, M.J. Argonaute proteins: Key players in RNA silencing. Nat. Rev. Mol. Cell Biol. 2008, 9, 22–32. [Google Scholar] [CrossRef] [PubMed]
  93. Meister, G.; Landthaler, M.; Patkaniowska, A.; Dorsett, Y.; Teng, G.; Tuschl, T. Human Argonaute2 mediates RNA cleavage targeted by miRNAs and siRNAs. Mol. Cell 2004, 15, 185–197. [Google Scholar] [CrossRef] [PubMed]
  94. Doench, J.G.; Petersen, C.P.; Sharp, P.A. siRNAs can function as miRNAs. Genes Dev. 2003, 17, 438–442. [Google Scholar] [CrossRef] [PubMed]
  95. Zeng, Y.; Yi, R.; Cullen, B.R. MicroRNAs and small interfering RNAs can inhibit mRNA expression by similar mechanisms. Proc. Natl. Acad. Sci. USA 2003, 100, 9779–9784. [Google Scholar] [CrossRef] [PubMed]
  96. Jinek, M.; Doudna, J.A. A three-dimensional view of the molecular machinery of RNA interference. Nature 2009, 457, 405–412. [Google Scholar] [CrossRef] [PubMed]
  97. Carthew, R.W.; Sontheimer, E.J. Origins and Mechanisms of miRNAs and siRNAs. Cell 2009, 136, 642–655. [Google Scholar] [CrossRef] [PubMed]
  98. Tang, G. siRNA and miRNA: An insight into RISCs. Trends Biochem. Sci. 2005, 30, 106–114. [Google Scholar] [CrossRef] [PubMed]
  99. Klattenhoff, C.; Theurkauf, W. Biogenesis and germline functions of piRNAs. Development 2008, 135, 3–9. [Google Scholar] [CrossRef] [PubMed]
  100. Malone, C.D.; Hannon, G.J. Small RNAs as guardians of the genome. Cell 2009, 136, 656–668. [Google Scholar] [CrossRef] [PubMed]
  101. Bartel, D.P. MicroRNAs: Genomics, biogenesis, mechanism, and function. Cell 2004, 116, 281–297. [Google Scholar] [CrossRef]
  102. He, L.; Hannon, G.J. MicroRNAs: Small RNAs with a big role in gene regulation. Nat. Rev. Genet. 2004, 5, 522–531. [Google Scholar] [CrossRef] [PubMed]
  103. Fromm, B.; Billipp, T.; Peck, L.E.; Johansen, M.; Tarver, J.E.; King, B.L.; Newcomb, J.M.; Sempere, L.F.; Flatmark, K.; Hovig, E.; et al. A Uniform System for the Annotation of Vertebrate microRNA Genes and the Evolution of the Human microRNAome. Annu. Rev. Genet. 2015, 49, 213–242. [Google Scholar] [CrossRef] [PubMed]
  104. Kozomara, A.; Griffiths-Jones, S. miRBase: Annotating high confidence microRNAs using deep sequencing data. Nucleic Acids Res. 2014, 42, D68–D73. [Google Scholar] [CrossRef] [PubMed]
  105. Londin, E.; Loher, P.; Telonis, A.G.; Quann, K.; Clark, P.; Jing, Y.; Hatzimichael, E.; Kirino, Y.; Honda, S.; Lally, M.; et al. Analysis of 13 cell types reveals evidence for the expression of numerous novel primate- and tissue-specific microRNAs. Proc. Natl. Acad. Sci. USA 2015, 112, E1106–E1115. [Google Scholar] [CrossRef] [PubMed]
  106. Lewis, B.P.; Burge, C.B.; Bartel, D.P. Conserved seed pairing, often flanked by adenosines, indicates that thousands of human genes are microRNA targets. Cell 2005, 120, 15–20. [Google Scholar] [CrossRef] [PubMed]
  107. Bartel, D.P. MicroRNAs: Target recognition and regulatory functions. Cell 2009, 136, 215–233. [Google Scholar] [CrossRef] [PubMed]
  108. Friedman, R.C.; Farh, K.K.; Burge, C.B.; Bartel, D.P. Most mammalian mRNAs are conserved targets of microRNAs. Genome Res. 2009, 19, 92–105. [Google Scholar] [CrossRef] [PubMed]
  109. Miranda, K.C.; Huynh, T.; Tay, Y.; Ang, Y.S.; Tam, W.L.; Thomson, A.M.; Lim, B.; Rigoutsos, I. A pattern-based method for the identification of MicroRNA binding sites and their corresponding heteroduplexes. Cell 2006, 126, 1203–1217. [Google Scholar] [CrossRef] [PubMed]
  110. Lim, L.P.; Lau, N.C.; Garrett-Engele, P.; Grimson, A.; Schelter, J.M.; Castle, J.; Bartel, D.P.; Linsley, P.S.; Johnson, J.M. Microarray analysis shows that some microRNAs downregulate large numbers of target mRNAs. Nature 2005, 433, 769–773. [Google Scholar] [CrossRef] [PubMed]
  111. Griffiths-Jones, S.; Grocock, R.J.; van Dongen, S.; Bateman, A.; Enright, A.J. miRBase: MicroRNA sequences, targets and gene nomenclature. Nucleic Acids Res. 2006, 34, D140–D144. [Google Scholar] [CrossRef] [PubMed]
  112. Lewis, B.P.; Shih, I.H.; Jones-Rhoades, M.W.; Bartel, D.P.; Burge, C.B. Prediction of mammalian microRNA targets. Cell 2003, 115, 787–798. [Google Scholar] [CrossRef]
  113. Lee, Y.; Kim, M.; Han, J.; Yeom, K.H.; Lee, S.; Baek, S.H.; Kim, V.N. MicroRNA genes are transcribed by RNA polymerase II. EMBO J. 2004, 23, 4051–4060. [Google Scholar] [CrossRef] [PubMed]
  114. Ha, M.; Kim, V.N. Regulation of microRNA biogenesis. Nat. Rev. Mol. Cell Biol. 2014, 15, 509–524. [Google Scholar] [CrossRef] [PubMed]
  115. Lee, Y.; Ahn, C.; Han, J.; Choi, H.; Kim, J.; Yim, J.; Lee, J.; Provost, P.; Radmark, O.; Kim, S.; et al. The nuclear RNase III Drosha initiates microRNA processing. Nature 2003, 425, 415–419. [Google Scholar] [CrossRef] [PubMed]
  116. Denli, A.M.; Tops, B.B.; Plasterk, R.H.; Ketting, R.F.; Hannon, G.J. Processing of primary microRNAs by the Microprocessor complex. Nature 2004, 432, 231–235. [Google Scholar] [CrossRef] [PubMed]
  117. Han, J.; Lee, Y.; Yeom, K.H.; Nam, J.W.; Heo, I.; Rhee, J.K.; Sohn, S.Y.; Cho, Y.; Zhang, B.T.; Kim, V.N. Molecular basis for the recognition of primary microRNAs by the Drosha-DGCR8 complex. Cell 2006, 125, 887–901. [Google Scholar] [CrossRef] [PubMed]
  118. Nguyen, T.A.; Jo, M.H.; Choi, Y.G.; Park, J.; Kwon, S.C.; Hohng, S.; Kim, V.N.; Woo, J.S. Functional Anatomy of the Human Microprocessor. Cell 2015, 161, 1374–1387. [Google Scholar] [CrossRef] [PubMed]
  119. Kwon, S.C.; Nguyen, T.A.; Choi, Y.G.; Jo, M.H.; Hohng, S.; Kim, V.N.; Woo, J.S. Structure of Human DROSHA. Cell 2016, 164, 81–90. [Google Scholar] [CrossRef] [PubMed]
  120. Auyeung, V.C.; Ulitsky, I.; McGeary, S.E.; Bartel, D.P. Beyond secondary structure: Primary-sequence determinants license pri-miRNA hairpins for processing. Cell 2013, 152, 844–858. [Google Scholar] [CrossRef] [PubMed]
  121. Yi, R.; Qin, Y.; Macara, I.G.; Cullen, B.R. Exportin-5 mediates the nuclear export of pre-microRNAs and short hairpin RNAs. Genes Dev. 2003, 17, 3011–3016. [Google Scholar] [CrossRef] [PubMed]
  122. Lund, E.; Guttinger, S.; Calado, A.; Dahlberg, J.E.; Kutay, U. Nuclear export of microRNA precursors. Science 2004, 303, 95–98. [Google Scholar] [CrossRef] [PubMed]
  123. Bohnsack, M.T.; Czaplinski, K.; Gorlich, D. Exportin 5 is a RanGTP-dependent dsRNA-binding protein that mediates nuclear export of pre-miRNAs. RNA 2004, 10, 185–191. [Google Scholar] [CrossRef] [PubMed]
  124. Chendrimada, T.P.; Gregory, R.I.; Kumaraswamy, E.; Norman, J.; Cooch, N.; Nishikura, K.; Shiekhattar, R. TRBP recruits the Dicer complex to Ago2 for microRNA processing and gene silencing. Nature 2005, 436, 740–744. [Google Scholar] [CrossRef] [PubMed]
  125. Wilson, R.C.; Tambe, A.; Kidwell, M.A.; Noland, C.L.; Schneider, C.P.; Doudna, J.A. Dicer-TRBP complex formation ensures accurate mammalian microRNA biogenesis. Mol. Cell 2015, 57, 397–407. [Google Scholar] [CrossRef] [PubMed]
  126. Kim, Y.; Yeo, J.; Lee, J.H.; Cho, J.; Seo, D.; Kim, J.S.; Kim, V.N. Deletion of human tarbp2 reveals cellular microRNA targets and cell-cycle function of TRBP. Cell Rep. 2014, 9, 1061–1074. [Google Scholar] [CrossRef] [PubMed]
  127. Macrae, I.J.; Zhou, K.; Li, F.; Repic, A.; Brooks, A.N.; Cande, W.Z.; Adams, P.D.; Doudna, J.A. Structural basis for double-stranded RNA processing by Dicer. Science 2006, 311, 195–198. [Google Scholar] [CrossRef] [PubMed]
  128. MacRae, I.J.; Zhou, K.; Doudna, J.A. Structural determinants of RNA recognition and cleavage by Dicer. Nat. Struct. Mol. Biol. 2007, 14, 934–940. [Google Scholar] [CrossRef] [PubMed]
  129. Park, J.E.; Heo, I.; Tian, Y.; Simanshu, D.K.; Chang, H.; Jee, D.; Patel, D.J.; Kim, V.N. Dicer recognizes the 5′ end of RNA for efficient and accurate processing. Nature 2011, 475, 201–205. [Google Scholar] [CrossRef] [PubMed]
  130. Lau, P.W.; Guiley, K.Z.; De, N.; Potter, C.S.; Carragher, B.; MacRae, I.J. The molecular architecture of human Dicer. Nat. Struct. Mol. Biol. 2012, 19, 436–440. [Google Scholar] [CrossRef] [PubMed]
  131. Lee, Y.; Hur, I.; Park, S.Y.; Kim, Y.K.; Suh, M.R.; Kim, V.N. The role of PACT in the RNA silencing pathway. EMBO J. 2006, 25, 522–532. [Google Scholar] [CrossRef] [PubMed]
  132. Lee, H.Y.; Zhou, K.; Smith, A.M.; Noland, C.L.; Doudna, J.A. Differential roles of human Dicer-binding proteins TRBP and PACT in small RNA processing. Nucleic Acids Res. 2013, 41, 6568–6576. [Google Scholar] [CrossRef] [PubMed]
  133. Czech, B.; Hannon, G.J. Small RNA sorting: Matchmaking for Argonautes. Nat. Rev. Genet. 2011, 12, 19–31. [Google Scholar] [CrossRef] [PubMed]
  134. Iwasaki, S.; Kobayashi, M.; Yoda, M.; Sakaguchi, Y.; Katsuma, S.; Suzuki, T.; Tomari, Y. Hsc70/Hsp90 chaperone machinery mediates ATP-dependent RISC loading of small RNA duplexes. Mol. Cell 2010, 39, 292–299. [Google Scholar] [CrossRef] [PubMed]
  135. Meister, G. Argonaute proteins: Functional insights and emerging roles. Nat. Rev. Genet. 2013, 14, 447–459. [Google Scholar] [CrossRef] [PubMed]
  136. Nakanishi, K. Anatomy of RISC: How do small RNAs and chaperones activate Argonaute proteins? Wiley Interdiscip. Rev. RNA 2016, 7, 637–660. [Google Scholar] [CrossRef] [PubMed]
  137. Kwak, P.B.; Tomari, Y. The N domain of Argonaute drives duplex unwinding during RISC assembly. Nat. Struct. Mol. Biol. 2012, 19, 145–151. [Google Scholar] [CrossRef] [PubMed]
  138. Noland, C.L.; Doudna, J.A. Multiple sensors ensure guide strand selection in human RNAi pathways. RNA 2013, 19, 639–648. [Google Scholar] [CrossRef] [PubMed]
  139. Khvorova, A.; Reynolds, A.; Jayasena, S.D. Functional siRNAs and miRNAs exhibit strand bias. Cell 2003, 115, 209–216. [Google Scholar] [CrossRef]
  140. Matranga, C.; Tomari, Y.; Shin, C.; Bartel, D.P.; Zamore, P.D. Passenger-strand cleavage facilitates assembly of siRNA into Ago2-containing RNAi enzyme complexes. Cell 2005, 123, 607–620. [Google Scholar] [CrossRef] [PubMed]
  141. Rand, T.A.; Petersen, S.; Du, F.; Wang, X. Argonaute2 cleaves the anti-guide strand of siRNA during RISC activation. Cell 2005, 123, 621–629. [Google Scholar] [CrossRef] [PubMed]
  142. Gregory, R.I.; Chendrimada, T.P.; Cooch, N.; Shiekhattar, R. Human RISC couples microRNA biogenesis and posttranscriptional gene silencing. Cell 2005, 123, 631–640. [Google Scholar] [CrossRef] [PubMed]
  143. Grimson, A.; Farh, K.K.; Johnston, W.K.; Garrett-Engele, P.; Lim, L.P.; Bartel, D.P. MicroRNA targeting specificity in mammals: Determinants beyond seed pairing. Mol. Cell 2007, 27, 91–105. [Google Scholar] [CrossRef] [PubMed]
  144. Kertesz, M.; Iovino, N.; Unnerstall, U.; Gaul, U.; Segal, E. The role of site accessibility in microRNA target recognition. Nat. Genet. 2007, 39, 1278–1284. [Google Scholar] [CrossRef] [PubMed]
  145. Nam, J.W.; Rissland, O.S.; Koppstein, D.; Abreu-Goodger, C.; Jan, C.H.; Agarwal, V.; Yildirim, M.A.; Rodriguez, A.; Bartel, D.P. Global analyses of the effect of different cellular contexts on microRNA targeting. Mol. Cell 2014, 53, 1031–1043. [Google Scholar] [CrossRef] [PubMed]
  146. Schirle, N.T.; Sheu-Gruttadauria, J.; MacRae, I.J. Structural basis for microRNA targeting. Science 2014, 346, 608–613. [Google Scholar] [CrossRef] [PubMed]
  147. Chandradoss, S.D.; Schirle, N.T.; Szczepaniak, M.; MacRae, I.J.; Joo, C. A Dynamic Search Process Underlies MicroRNA Targeting. Cell 2015, 162, 96–107. [Google Scholar] [CrossRef] [PubMed]
  148. Hausser, J.; Syed, A.P.; Bilen, B.; Zavolan, M. Analysis of CDS-located miRNA target sites suggests that they can effectively inhibit translation. Genome Res. 2013, 23, 604–615. [Google Scholar] [CrossRef] [PubMed]
  149. Lytle, J.R.; Yario, T.A.; Steitz, J.A. Target mRNAs are repressed as efficiently by microRNA-binding sites in the 5′UTR as in the 3′UTR. Proc. Natl. Acad. Sci. USA 2007, 104, 9667–9672. [Google Scholar] [CrossRef] [PubMed]
  150. Guo, H.; Ingolia, N.T.; Weissman, J.S.; Bartel, D.P. Mammalian microRNAs predominantly act to decrease target mRNA levels. Nature 2010, 466, 835–840. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  151. Djuranovic, S.; Nahvi, A.; Green, R. miRNA-mediated gene silencing by translational repression followed by mRNA deadenylation and decay. Science 2012, 336, 237–240. [Google Scholar] [CrossRef] [PubMed]
  152. Meijer, H.A.; Kong, Y.W.; Lu, W.T.; Wilczynska, A.; Spriggs, R.V.; Robinson, S.W.; Godfrey, J.D.; Willis, A.E.; Bushell, M. Translational repression and eIF4A2 activity are critical for microRNA-mediated gene regulation. Science 2013, 340, 82–85. [Google Scholar] [CrossRef] [PubMed]
  153. Eichhorn, S.W.; Guo, H.; McGeary, S.E.; Rodriguez-Mias, R.A.; Shin, C.; Baek, D.; Hsu, S.H.; Ghoshal, K.; Villen, J.; Bartel, D.P. mRNA destabilization is the dominant effect of mammalian microRNAs by the time substantial repression ensues. Mol. Cell 2014, 56, 104–115. [Google Scholar] [CrossRef] [PubMed]
  154. Agarwal, V.; Bell, G.W.; Nam, J.W.; Bartel, D.P. Predicting effective microRNA target sites in mammalian mRNAs. eLife 2015, 4. [Google Scholar] [CrossRef] [PubMed]
  155. Jonas, S.; Izaurralde, E. Towards a molecular understanding of microRNA-mediated gene silencing. Nat. Rev. Genet. 2015, 16, 421–433. [Google Scholar] [CrossRef] [PubMed]
  156. Iwakawa, H.O.; Tomari, Y. The Functions of MicroRNAs: MRNA Decay and Translational Repression. Trends Cell Biol. 2015, 25, 651–665. [Google Scholar] [CrossRef] [PubMed]
  157. Berezikov, E.; Chung, W.J.; Willis, J.; Cuppen, E.; Lai, E.C. Mammalian mirtron genes. Mol. Cell 2007, 28, 328–336. [Google Scholar] [CrossRef] [PubMed]
  158. Ruby, J.G.; Jan, C.H.; Bartel, D.P. Intronic microRNA precursors that bypass Drosha processing. Nature 2007, 448, 83–86. [Google Scholar] [CrossRef] [PubMed]
  159. Okamura, K.; Hagen, J.W.; Duan, H.; Tyler, D.M.; Lai, E.C. The mirtron pathway generates microRNA-class regulatory RNAs in Drosophila. Cell 2007, 130, 89–100. [Google Scholar] [CrossRef] [PubMed]
  160. Yang, J.S.; Maurin, T.; Robine, N.; Rasmussen, K.D.; Jeffrey, K.L.; Chandwani, R.; Papapetrou, E.P.; Sadelain, M.; O'Carroll, D.; Lai, E.C. Conserved vertebrate mir-451 provides a platform for Dicer-independent, Ago2-mediated microRNA biogenesis. Proc. Natl. Acad. Sci. USA 2010, 107, 15163–15168. [Google Scholar] [CrossRef] [PubMed]
  161. Cifuentes, D.; Xue, H.; Taylor, D.W.; Patnode, H.; Mishima, Y.; Cheloufi, S.; Ma, E.; Mane, S.; Hannon, G.J.; Lawson, N.D.; et al. A novel miRNA processing pathway independent of Dicer requires Argonaute2 catalytic activity. Science 2010, 328, 1694–1698. [Google Scholar] [CrossRef] [PubMed]
  162. Cheloufi, S.; Dos Santos, C.O.; Chong, M.M.; Hannon, G.J. A dicer-independent miRNA biogenesis pathway that requires Ago catalysis. Nature 2010, 465, 584–589. [Google Scholar] [CrossRef] [PubMed]
  163. Hansen, T.B.; Veno, M.T.; Jensen, T.I.; Schaefer, A.; Damgaard, C.K.; Kjems, J. Argonaute-associated short introns are a novel class of gene regulators. Nat. Commun. 2016, 7, 11538. [Google Scholar] [CrossRef] [PubMed]
  164. Daugaard, I.; Hansen, T.B. Biogenesis and Function of Ago-Associated RNAs. Trends Genet. 2017, 33, 208–219. [Google Scholar] [CrossRef] [PubMed]
  165. Jopling, C.L.; Yi, M.; Lancaster, A.M.; Lemon, S.M.; Sarnow, P. Modulation of hepatitis C virus RNA abundance by a liver-specific MicroRNA. Science 2005, 309, 1577–1581. [Google Scholar] [CrossRef] [PubMed]
  166. Shimakami, T.; Yamane, D.; Jangra, R.K.; Kempf, B.J.; Spaniel, C.; Barton, D.J.; Lemon, S.M. Stabilization of hepatitis C virus RNA by an Ago2-miR-122 complex. Proc. Natl. Acad. Sci. USA 2012, 109, 941–946. [Google Scholar] [CrossRef] [PubMed]
  167. Li, Y.; Masaki, T.; Yamane, D.; McGivern, D.R.; Lemon, S.M. Competing and noncompeting activities of miR-122 and the 5′ exonuclease Xrn1 in regulation of hepatitis C virus replication. Proc. Natl. Acad. Sci. USA 2013, 110, 1881–1886. [Google Scholar] [CrossRef] [PubMed]
  168. Masaki, T.; Arend, K.C.; Li, Y.; Yamane, D.; McGivern, D.R.; Kato, T.; Wakita, T.; Moorman, N.J.; Lemon, S.M. miR-122 stimulates hepatitis C virus RNA synthesis by altering the balance of viral RNAs engaged in replication versus translation. Cell Host Microbe 2015, 17, 217–228. [Google Scholar] [CrossRef] [PubMed]
  169. Scheel, T.K.; Luna, J.M.; Liniger, M.; Nishiuchi, E.; Rozen-Gagnon, K.; Shlomai, A.; Auray, G.; Gerber, M.; Fak, J.; Keller, I.; et al. A Broad RNA Virus Survey Reveals Both miRNA Dependence and Functional Sequestration. Cell Host Microbe 2016, 19, 409–423. [Google Scholar] [CrossRef] [PubMed]
  170. Luna, J.M.; Scheel, T.K.; Danino, T.; Shaw, K.S.; Mele, A.; Fak, J.J.; Nishiuchi, E.; Takacs, C.N.; Catanese, M.T.; de Jong, Y.P.; et al. Hepatitis C virus RNA functionally sequesters miR-122. Cell 2015, 160, 1099–1110. [Google Scholar] [CrossRef] [PubMed]
  171. Wang, S.; Qiu, L.; Yan, X.; Jin, W.; Wang, Y.; Chen, L.; Wu, E.; Ye, X.; Gao, G.F.; Wang, F.; et al. Loss of microRNA 122 expression in patients with hepatitis B enhances hepatitis B virus replication through cyclin G(1) -modulated P53 activity. Hepatology 2012, 55, 730–741. [Google Scholar] [CrossRef] [PubMed]
  172. Li, C.; Wang, Y.; Wang, S.; Wu, B.; Hao, J.; Fan, H.; Ju, Y.; Ding, Y.; Chen, L.; Chu, X.; et al. Hepatitis B virus mRNA-mediated miR-122 inhibition upregulates PTTG1-binding protein, which promotes hepatocellular carcinoma tumor growth and cell invasion. J. Virol. 2013, 87, 2193–2205. [Google Scholar] [CrossRef] [PubMed]
  173. Ho, B.C.; Yu, S.L.; Chen, J.J.; Chang, S.Y.; Yan, B.S.; Hong, Q.S.; Singh, S.; Kao, C.L.; Chen, H.Y.; Su, K.Y.; et al. Enterovirus-induced miR-141 contributes to shutoff of host protein translation by targeting the translation initiation factor eIF4E. Cell Host Microbe 2011, 9, 58–69. [Google Scholar] [CrossRef] [PubMed]
  174. Trobaugh, D.W.; Gardner, C.L.; Sun, C.; Haddow, A.D.; Wang, E.; Chapnik, E.; Mildner, A.; Weaver, S.C.; Ryman, K.D.; Klimstra, W.B. RNA viruses can hijack vertebrate microRNAs to suppress innate immunity. Nature 2014, 506, 245–248. [Google Scholar] [CrossRef] [PubMed]
  175. Linnstaedt, S.D.; Gottwein, E.; Skalsky, R.L.; Luftig, M.A.; Cullen, B.R. Virally induced cellular microRNA miR-155 plays a key role in B-cell immortalization by Epstein-Barr virus. J. Virol. 2010, 84, 11670–11678. [Google Scholar] [CrossRef] [PubMed]
  176. Mansouri, S.; Pan, Q.; Blencowe, B.J.; Claycomb, J.M.; Frappier, L. Epstein-Barr virus EBNA1 protein regulates viral latency through effects on let-7 microRNA and dicer. J. Virol. 2014, 88, 11166–11177. [Google Scholar] [CrossRef] [PubMed]
  177. Ellis-Connell, A.L.; Iempridee, T.; Xu, I.; Mertz, J.E. Cellular microRNAs 200b and 429 regulate the Epstein-Barr virus switch between latency and lytic replication. J. Virol. 2010, 84, 10329–10343. [Google Scholar] [CrossRef] [PubMed]
  178. Chiang, K.; Liu, H.; Rice, A.P. miR-132 enhances HIV-1 replication. Virology 2013, 438, 1–4. [Google Scholar] [CrossRef] [PubMed]
  179. Huang, J.; Wang, F.; Argyris, E.; Chen, K.; Liang, Z.; Tian, H.; Huang, W.; Squires, K.; Verlinghieri, G.; Zhang, H. Cellular microRNAs contribute to HIV-1 latency in resting primary CD4+ T lymphocytes. Nat. Med. 2007, 13, 1241–1247. [Google Scholar] [CrossRef] [PubMed]
  180. Nathans, R.; Chu, C.Y.; Serquina, A.K.; Lu, C.C.; Cao, H.; Rana, T.M. Cellular microRNA and P bodies modulate host-HIV-1 interactions. Mol. Cell 2009, 34, 696–709. [Google Scholar] [CrossRef] [PubMed]
  181. Triboulet, R.; Mari, B.; Lin, Y.L.; Chable-Bessia, C.; Bennasser, Y.; Lebrigand, K.; Cardinaud, B.; Maurin, T.; Barbry, P.; Baillat, V.; et al. Suppression of microRNA-silencing pathway by HIV-1 during virus replication. Science 2007, 315, 1579–1582. [Google Scholar] [CrossRef] [PubMed]
  182. Sung, T.L.; Rice, A.P. miR-198 inhibits HIV-1 gene expression and replication in monocytes and its mechanism of action appears to involve repression of cyclin T1. PLoS Pathog. 2009, 5, e1000263. [Google Scholar] [CrossRef] [PubMed]
  183. Chiang, K.; Sung, T.L.; Rice, A.P. Regulation of cyclin T1 and HIV-1 Replication by microRNAs in resting CD4+ T lymphocytes. J. Virol. 2012, 86, 3244–3252. [Google Scholar] [CrossRef] [PubMed]
  184. Klein, M.E.; Lioy, D.T.; Ma, L.; Impey, S.; Mandel, G.; Goodman, R.H. Homeostatic regulation of MeCP2 expression by a CREB-induced microRNA. Nat. Neurosci. 2007, 10, 1513–1514. [Google Scholar] [CrossRef] [PubMed]
  185. Alvarez-Saavedra, M.; Antoun, G.; Yanagiya, A.; Oliva-Hernandez, R.; Cornejo-Palma, D.; Perez-Iratxeta, C.; Sonenberg, N.; Cheng, H.Y. miRNA-132 orchestrates chromatin remodeling and translational control of the circadian clock. Hum. Mol. Genet. 2011, 20, 731–751. [Google Scholar] [CrossRef] [PubMed]
  186. Zhang, H.S.; Chen, X.Y.; Wu, T.C.; Sang, W.W.; Ruan, Z. MiR-34a is involved in Tat-induced HIV-1 long terminal repeat (LTR) transactivation through the SIRT1/NFkappaB pathway. FEBS Lett. 2012, 586, 4203–4207. [Google Scholar] [CrossRef] [PubMed]
  187. Zhang, H.S.; Wu, T.C.; Sang, W.W.; Ruan, Z. MiR-217 is involved in Tat-induced HIV-1 long terminal repeat (LTR) transactivation by down-regulation of SIRT1. Biochim. Biophys. Acta 2012, 1823, 1017–1023. [Google Scholar] [CrossRef] [PubMed]
  188. Chen, H.; Li, C.; Huang, J.; Cung, T.; Seiss, K.; Beamon, J.; Carrington, M.F.; Porter, L.C.; Burke, P.S.; Yang, Y.; et al. CD4+ T cells from elite controllers resist HIV-1 infection by selective upregulation of p21. J. Clin. Investig. 2011, 121, 1549–1560. [Google Scholar] [CrossRef] [PubMed]
  189. Hsu, K.; Seharaseyon, J.; Dong, P.; Bour, S.; Marban, E. Mutual functional destruction of HIV-1 Vpu and host TASK-1 channel. Mol. Cell 2004, 14, 259–267. [Google Scholar] [CrossRef]
  190. Farberov, L.; Herzig, E.; Modai, S.; Isakov, O.; Hizi, A.; Shomron, N. MicroRNA-mediated regulation of p21 and TASK1 cellular restriction factors enhances HIV-1 infection. J. Cell Sci. 2015, 128, 1607–1616. [Google Scholar] [CrossRef] [PubMed]
  191. Amaral, A.J.; Andrade, J.; Foxall, R.B.; Matoso, P.; Matos, A.M.; Soares, R.S.; Rocha, C.; Ramos, C.G.; Tendeiro, R.; Serra-Caetano, A.; et al. miRNA profiling of human naive CD4 T cells links miR-34c-5p to cell activation and HIV replication. EMBO J. 2017, 36, 346–360. [Google Scholar] [CrossRef] [PubMed]
  192. Jeang, K.T. RNAi in the regulation of mammalian viral infections. BMC Biol. 2012, 10, 58. [Google Scholar] [CrossRef] [PubMed]
  193. Zhou, R.; Rana, T.M. RNA-based mechanisms regulating host-virus interactions. Immunol. Rev. 2013, 253, 97–111. [Google Scholar] [CrossRef] [PubMed]
  194. Cullen, B.R.; Cherry, S.; tenOever, B.R. Is RNA interference a physiologically relevant innate antiviral immune response in mammals? Cell Host Microbe 2013, 14, 374–378. [Google Scholar] [CrossRef] [PubMed]
  195. Cullen, B.R. Viruses and RNA interference: Issues and controversies. J. Virol. 2014, 88, 12934–12936. [Google Scholar] [CrossRef] [PubMed]
  196. Parameswaran, P.; Sklan, E.; Wilkins, C.; Burgon, T.; Samuel, M.A.; Lu, R.; Ansel, K.M.; Heissmeyer, V.; Einav, S.; Jackson, W.; et al. Six RNA viruses and forty-one hosts: Viral small RNAs and modulation of small RNA repertoires in vertebrate and invertebrate systems. PLoS Pathog. 2010, 6, e1000764. [Google Scholar] [CrossRef] [PubMed]
  197. Lin, Y.T.; Kincaid, R.P.; Arasappan, D.; Dowd, S.E.; Hunicke-Smith, S.P.; Sullivan, C.S. Small RNA profiling reveals antisense transcription throughout the KSHV genome and novel small RNAs. RNA 2010, 16, 1540–1558. [Google Scholar] [CrossRef] [PubMed]
  198. Bogerd, H.P.; Whisnant, A.W.; Kennedy, E.M.; Flores, O.; Cullen, B.R. Derivation and characterization of Dicer- and microRNA-deficient human cells. RNA 2014, 20, 923–937. [Google Scholar] [CrossRef] [PubMed]
  199. Bogerd, H.P.; Skalsky, R.L.; Kennedy, E.M.; Furuse, Y.; Whisnant, A.W.; Flores, O.; Schultz, K.L.; Putnam, N.; Barrows, N.J.; Sherry, B.; et al. Replication of many human viruses is refractory to inhibition by endogenous cellular microRNAs. J. Virol. 2014, 88, 8065–8076. [Google Scholar] [CrossRef] [PubMed]
  200. Seo, G.J.; Kincaid, R.P.; Phanaksri, T.; Burke, J.M.; Pare, J.M.; Cox, J.E.; Hsiang, T.Y.; Krug, R.M.; Sullivan, C.S. Reciprocal inhibition between intracellular antiviral signaling and the RNAi machinery in mammalian cells. Cell Host Microbe 2013, 14, 435–445. [Google Scholar] [CrossRef] [PubMed]
  201. Backes, S.; Langlois, R.A.; Schmid, S.; Varble, A.; Shim, J.V.; Sachs, D.; tenOever, B.R. The Mammalian response to virus infection is independent of small RNA silencing. Cell Rep. 2014, 8, 114–125. [Google Scholar] [CrossRef] [PubMed]
  202. Maillard, P.V.; Ciaudo, C.; Marchais, A.; Li, Y.; Jay, F.; Ding, S.W.; Voinnet, O. Antiviral RNA interference in mammalian cells. Science 2013, 342, 235–238. [Google Scholar] [CrossRef] [PubMed]
  203. Li, Y.; Lu, J.; Han, Y.; Fan, X.; Ding, S.W. RNA interference functions as an antiviral immunity mechanism in mammals. Science 2013, 342, 231–234. [Google Scholar] [CrossRef] [PubMed]
  204. Maillard, P.V.; Van der Veen, A.G.; Deddouche-Grass, S.; Rogers, N.C.; Merits, A.; Reis, E.S.C. Inactivation of the type I interferon pathway reveals long double-stranded RNA-mediated RNA interference in mammalian cells. EMBO J. 2016, 35, 2505–2518. [Google Scholar] [CrossRef] [PubMed]
  205. Li, Y.; Basavappa, M.; Lu, J.; Dong, S.; Cronkite, D.A.; Prior, J.T.; Reinecker, H.C.; Hertzog, P.; Han, Y.; Li, W.X.; et al. Induction and suppression of antiviral RNA interference by influenza A virus in mammalian cells. Nat. Microbiol. 2016, 2, 16250. [Google Scholar] [CrossRef] [PubMed]
  206. Qiu, Y.; Xu, Y.; Zhang, Y.; Zhou, H.; Deng, Y.Q.; Li, X.F.; Miao, M.; Zhang, Q.; Zhong, B.; Hu, Y.; et al. Human Virus-Derived Small RNAs Can Confer Antiviral Immunity in Mammals. Immunity 2017, 46, 992–1004. [Google Scholar] [CrossRef] [PubMed]
  207. Schuster, S.; Tholen, L.E.; Overheul, G.J.; van Kuppeveld, F.J.M.; van Rij, R.P. Deletion of Cytoplasmic Double-Stranded RNA Sensors Does Not Uncover Viral Small Interfering RNA Production in Human Cells. mSphere 2017, 2. [Google Scholar] [CrossRef] [PubMed]
  208. Lecellier, C.H.; Dunoyer, P.; Arar, K.; Lehmann-Che, J.; Eyquem, S.; Himber, C.; Saib, A.; Voinnet, O. A cellular microRNA mediates antiviral defense in human cells. Science 2005, 308, 557–560. [Google Scholar] [CrossRef] [PubMed]
  209. Pedersen, I.M.; Cheng, G.; Wieland, S.; Volinia, S.; Croce, C.M.; Chisari, F.V.; David, M. Interferon modulation of cellular microRNAs as an antiviral mechanism. Nature 2007, 449, 919–922. [Google Scholar] [CrossRef] [PubMed]
  210. Otsuka, M.; Jing, Q.; Georgel, P.; New, L.; Chen, J.; Mols, J.; Kang, Y.J.; Jiang, Z.; Du, X.; Cook, R.; et al. Hypersusceptibility to vesicular stomatitis virus infection in Dicer1-deficient mice is due to impaired miR24 and miR93 expression. Immunity 2007, 27, 123–134. [Google Scholar] [CrossRef] [PubMed]
  211. Zheng, Z.; Ke, X.; Wang, M.; He, S.; Li, Q.; Zheng, C.; Zhang, Z.; Liu, Y.; Wang, H. Human microRNA hsa-miR-296-5p suppresses enterovirus 71 replication by targeting the viral genome. J. Virol. 2013, 87, 5645–5656. [Google Scholar] [CrossRef] [PubMed]
  212. Robertson, K.A.; Hsieh, W.Y.; Forster, T.; Blanc, M.; Lu, H.; Crick, P.J.; Yutuc, E.; Watterson, S.; Martin, K.; Griffiths, S.J.; et al. An Interferon Regulated MicroRNA Provides Broad Cell-Intrinsic Antiviral Immunity through Multihit Host-Directed Targeting of the Sterol Pathway. PLoS Biol. 2016, 14, e1002364. [Google Scholar] [CrossRef] [PubMed]
  213. Backes, S.; Shapiro, J.S.; Sabin, L.R.; Pham, A.M.; Reyes, I.; Moss, B.; Cherry, S.; tenOever, B.R. Degradation of host microRNAs by poxvirus poly(A) polymerase reveals terminal RNA methylation as a protective antiviral mechanism. Cell Host Microbe 2012, 12, 200–210. [Google Scholar] [CrossRef] [PubMed]
  214. Sanghvi, V.R.; Steel, L.F. A re-examination of global suppression of RNA interference by HIV-1. PLoS ONE 2011, 6, e17246. [Google Scholar] [CrossRef] [PubMed]
  215. Langlois, R.A.; Albrecht, R.A.; Kimble, B.; Sutton, T.; Shapiro, J.S.; Finch, C.; Angel, M.; Chua, M.A.; Gonzalez-Reiche, A.S.; Xu, K.; et al. MicroRNA-based strategy to mitigate the risk of gain-of-function influenza studies. Nat. Biotechnol. 2013, 31, 844–847. [Google Scholar] [CrossRef] [PubMed]
  216. Heiss, B.L.; Maximova, O.A.; Pletnev, A.G. Insertion of microRNA targets into the flavivirus genome alters its highly neurovirulent phenotype. J. Virol. 2011, 85, 1464–1472. [Google Scholar] [CrossRef] [PubMed]
  217. Bai, X.T.; Nicot, C. miR-28-3p is a cellular restriction factor that inhibits human T cell leukemia virus, type 1 (HTLV-1) replication and virus infection. J. Biol. Chem. 2015, 290, 5381–5390. [Google Scholar] [CrossRef] [PubMed]
  218. Presloid, J.B.; Novella, I.S. RNA Viruses and RNAi: Quasispecies Implications for Viral Escape. Viruses 2015, 7, 3226–3240. [Google Scholar] [CrossRef] [PubMed]
  219. Hariharan, M.; Scaria, V.; Pillai, B.; Brahmachari, S.K. Targets for human encoded microRNAs in HIV genes. Biochem. Biophys. Res. Commun. 2005, 337, 1214–1218. [Google Scholar] [CrossRef] [PubMed]
  220. Ahluwalia, J.K.; Khan, S.Z.; Soni, K.; Rawat, P.; Gupta, A.; Hariharan, M.; Scaria, V.; Lalwani, M.; Pillai, B.; Mitra, D.; et al. Human cellular microRNA hsa-miR-29a interferes with viral nef protein expression and HIV-1 replication. Retrovirology 2008, 5, 117. [Google Scholar] [CrossRef] [PubMed]
  221. Houzet, L.; Klase, Z.; Yeung, M.L.; Wu, A.; Le, S.Y.; Quinones, M.; Jeang, K.T. The extent of sequence complementarity correlates with the potency of cellular miRNA-mediated restriction of HIV-1. Nucleic Acids Res. 2012, 40, 11684–11696. [Google Scholar] [CrossRef] [PubMed]
  222. Whisnant, A.W.; Bogerd, H.P.; Flores, O.; Ho, P.; Powers, J.G.; Sharova, N.; Stevenson, M.; Chen, C.H.; Cullen, B.R. In-depth analysis of the interaction of HIV-1 with cellular microRNA biogenesis and effector mechanisms. MBio 2013, 4, e000193. [Google Scholar] [CrossRef] [PubMed]
  223. Wang, X.; Ye, L.; Hou, W.; Zhou, Y.; Wang, Y.J.; Metzger, D.S.; Ho, W.Z. Cellular microRNA expression correlates with susceptibility of monocytes/macrophages to HIV-1 infection. Blood 2009, 113, 671–674. [Google Scholar] [CrossRef] [PubMed]
  224. Wang, P.; Qu, X.; Zhou, X.; Shen, Y.; Ji, H.; Fu, Z.; Deng, J.; Lu, P.; Yu, W.; Lu, H.; et al. Two cellular microRNAs, miR-196b and miR-1290, contribute to HIV-1 latency. Virology 2015, 486, 228–238. [Google Scholar] [CrossRef] [PubMed]
  225. Sun, G.; Li, H.; Wu, X.; Covarrubias, M.; Scherer, L.; Meinking, K.; Luk, B.; Chomchan, P.; Alluin, J.; Gombart, A.F.; et al. Interplay between HIV-1 infection and host microRNAs. Nucleic Acids Res. 2012, 40, 2181–2196. [Google Scholar] [CrossRef] [PubMed]
  226. Yeung, M.L.; Bennasser, Y.; Myers, T.G.; Jiang, G.; Benkirane, M.; Jeang, K.T. Changes in microRNA expression profiles in HIV-1-transfected human cells. Retrovirology 2005, 2, 81. [Google Scholar] [CrossRef] [PubMed]
  227. Simon, V.; Ho, D.D. HIV-1 dynamics in vivo: Implications for therapy. Nat. Rev. Microbiol. 2003, 1, 181–190. [Google Scholar] [CrossRef] [PubMed]
  228. Mantri, C.K.; Pandhare Dash, J.; Mantri, J.V.; Dash, C.C. Cocaine enhances HIV-1 replication in CD4+ T cells by down-regulating MiR-125b. PLoS ONE 2012, 7, e51387. [Google Scholar] [CrossRef] [PubMed]
  229. Shen, C.J.; Jia, Y.H.; Tian, R.R.; Ding, M.; Zhang, C.; Wang, J.H. Translation of Pur-alpha is targeted by cellular miRNAs to modulate the differentiation-dependent susceptibility of monocytes to HIV-1 infection. FASEB J. 2012, 26, 4755–4764. [Google Scholar] [CrossRef] [PubMed]
  230. Fischer, S.; Handrick, R.; Aschrafi, A.; Otte, K. Unveiling the principle of microRNA-mediated redundancy in cellular pathway regulation. RNA Biol. 2015, 12, 238–247. [Google Scholar] [CrossRef] [PubMed]
  231. Ma, L.; Shen, C.J.; Cohen, E.A.; Xiong, S.D.; Wang, J.H. miRNA-1236 inhibits HIV-1 infection of monocytes by repressing translation of cellular factor VprBP. PLoS ONE 2014, 9, e99535. [Google Scholar] [CrossRef] [PubMed]
  232. Swaminathan, G.; Rossi, F.; Sierra, L.J.; Gupta, A.; Navas-Martin, S.; Martin-Garcia, J. A role for microRNA-155 modulation in the anti-HIV-1 effects of Toll-like receptor 3 stimulation in macrophages. PLoS Pathog. 2012, 8, e1002937. [Google Scholar] [CrossRef] [PubMed]
  233. Ruelas, D.S.; Chan, J.K.; Oh, E.; Heidersbach, A.J.; Hebbeler, A.M.; Chavez, L.; Verdin, E.; Rape, M.; Greene, W.C. MicroRNA-155 Reinforces HIV Latency. J. Biol. Chem. 2015, 290, 13736–13748. [Google Scholar] [CrossRef] [PubMed]
  234. Lodge, R.; Ferreira Barbosa, J.A.; Lombard-Vadnais, F.; Gilmore, J.C.; Deshiere, A.; Gosselin, A.; Wiche Salinas, T.R.; Bego, M.G.; Power, C.; Routy, J.P.; et al. Host MicroRNAs-221 and -222 Inhibit HIV-1 Entry in Macrophages by Targeting the CD4 Viral Receptor. Cell Rep. 2017, 21, 141–153. [Google Scholar] [CrossRef] [PubMed]
  235. Pilakka-Kanthikeel, S.; Raymond, A.; Atluri, V.S.; Sagar, V.; Saxena, S.K.; Diaz, P.; Chevelon, S.; Concepcion, M.; Nair, M. Sterile alpha motif and histidine/aspartic acid domain-containing protein 1 (SAMHD1)-facilitated HIV restriction in astrocytes is regulated by miRNA-181a. J. Neuroinflamm. 2015, 12, 66. [Google Scholar] [CrossRef] [PubMed]
  236. Jin, C.; Peng, X.; Liu, F.; Cheng, L.; Lu, X.; Yao, H.; Wu, H.; Wu, N. MicroRNA-181 expression regulates specific post-transcriptional level of SAMHD1 expression in vitro. Biochem. Biophys. Res. Commun. 2014, 452, 760–767. [Google Scholar] [CrossRef] [PubMed]
  237. Witwer, K.W.; Buchanan, E.L.; Myers, S.L.; McAlexander, M.A. miRNAs and SAMHD1 regulation in vitro and in a model of HIV CNS disease. J. Neuroinflamm. 2015, 12, 159. [Google Scholar] [CrossRef] [PubMed]
  238. Chable-Bessia, C.; Meziane, O.; Latreille, D.; Triboulet, R.; Zamborlini, A.; Wagschal, A.; Jacquet, J.M.; Reynes, J.; Levy, Y.; Saib, A.; et al. Suppression of HIV-1 replication by microRNA effectors. Retrovirology 2009, 6, 26. [Google Scholar] [CrossRef] [PubMed]
  239. Adoro, S.; Cubillos-Ruiz, J.R.; Chen, X.; Deruaz, M.; Vrbanac, V.D.; Song, M.; Park, S.; Murooka, T.T.; Dudek, T.E.; Luster, A.D.; et al. IL-21 induces antiviral microRNA-29 in CD4 T cells to limit HIV-1 infection. Nat. Commun. 2015, 6, 7562. [Google Scholar] [CrossRef] [PubMed]
  240. MacKay, C.R.; Wang, J.P.; Kurt-Jones, E.A. Dicer’s role as an antiviral: Still an enigma. Curr. Opin. Immunol. 2014, 26, 49–55. [Google Scholar] [CrossRef] [PubMed]
  241. Eckenfelder, A.; Segeral, E.; Pinzon, N.; Ulveling, D.; Amadori, C.; Charpentier, M.; Nidelet, S.; Concordet, J.P.; Zagury, J.F.; Paillart, J.C.; et al. Argonaute proteins regulate HIV-1 multiply spliced RNA and viral production in a Dicer independent manner. Nucleic Acids Res. 2017, 45, 4158–4173. [Google Scholar] [CrossRef] [PubMed]
  242. Bouttier, M.; Saumet, A.; Peter, M.; Courgnaud, V.; Schmidt, U.; Cazevieille, C.; Bertrand, E.; Lecellier, C.H. Retroviral GAG proteins recruit AGO2 on viral RNAs without affecting RNA accumulation and translation. Nucleic Acids Res. 2012, 40, 775–786. [Google Scholar] [CrossRef] [PubMed]
  243. Vongrad, V.; Imig, J.; Mohammadi, P.; Kishore, S.; Jaskiewicz, L.; Hall, J.; Gunthard, H.F.; Beerenwinkel, N.; Metzner, K.J. HIV-1 RNAs are Not Part of the Argonaute 2 Associated RNA Interference Pathway in Macrophages. PLoS ONE 2015, 10, e0132127. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  244. Chen, A.K.; Sengupta, P.; Waki, K.; Van Engelenburg, S.B.; Ochiya, T.; Ablan, S.D.; Freed, E.O.; Lippincott-Schwartz, J. MicroRNA binding to the HIV-1 Gag protein inhibits Gag assembly and virus production. Proc. Natl. Acad. Sci. USA 2014, 111, E2676–E2683. [Google Scholar] [CrossRef] [PubMed]
  245. Lin, J.; Cullen, B.R. Analysis of the interaction of primate retroviruses with the human RNA interference machinery. J. Virol. 2007, 81, 12218–12226. [Google Scholar] [CrossRef] [PubMed]
  246. Umbach, J.L.; Cullen, B.R. The role of RNAi and microRNAs in animal virus replication and antiviral immunity. Genes Dev. 2009, 23, 1151–1164. [Google Scholar] [CrossRef] [PubMed]
  247. Harwig, A.; Das, A.T.; Berkhout, B. Retroviral microRNAs. Curr. Opin. Virol. 2014, 7, 47–54. [Google Scholar] [CrossRef] [PubMed]
  248. Cullen, B.R. Viruses and microRNAs. Nat. Genet. 2006, 38, S25–S30. [Google Scholar] [CrossRef] [PubMed]
  249. Gottwein, E.; Cullen, B.R. Viral and cellular microRNAs as determinants of viral pathogenesis and immunity. Cell Host Microbe 2008, 3, 375–387. [Google Scholar] [CrossRef] [PubMed]
  250. Kincaid, R.P.; Sullivan, C.S. Virus-encoded microRNAs: An overview and a look to the future. PLoS Pathog. 2012, 8, e1003018. [Google Scholar] [CrossRef] [PubMed]
  251. Rosewick, N.; Momont, M.; Durkin, K.; Takeda, H.; Caiment, F.; Cleuter, Y.; Vernin, C.; Mortreux, F.; Wattel, E.; Burny, A.; et al. Deep sequencing reveals abundant noncanonical retroviral microRNAs in B-cell leukemia/lymphoma. Proc. Natl. Acad. Sci. USA 2013, 110, 2306–2311. [Google Scholar] [CrossRef] [PubMed]
  252. Kincaid, R.P.; Chen, Y.; Cox, J.E.; Rethwilm, A.; Sullivan, C.S. Noncanonical microRNA (miRNA) biogenesis gives rise to retroviral mimics of lymphoproliferative and immunosuppressive host miRNAs. MBio 2014, 5, e00074. [Google Scholar] [CrossRef] [PubMed]
  253. Whisnant, A.W.; Kehl, T.; Bao, Q.; Materniak, M.; Kuzmak, J.; Lochelt, M.; Cullen, B.R. Identification of novel, highly expressed retroviral microRNAs in cells infected by bovine foamy virus. J. Virol. 2014, 88, 4679–4686. [Google Scholar] [CrossRef] [PubMed]
  254. Gillet, N.A.; Hamaidia, M.; de Brogniez, A.; Gutierrez, G.; Renotte, N.; Reichert, M.; Trono, K.; Willems, L. Bovine Leukemia Virus Small Noncoding RNAs Are Functional Elements That Regulate Replication and Contribute to Oncogenesis In Vivo. PLoS Pathog. 2016, 12, e1005588. [Google Scholar] [CrossRef] [PubMed]
  255. Pfeffer, S.; Sewer, A.; Lagos-Quintana, M.; Sheridan, R.; Sander, C.; Grasser, F.A.; van Dyk, L.F.; Ho, C.K.; Shuman, S.; Chien, M.; et al. Identification of microRNAs of the herpesvirus family. Nat. Methods 2005, 2, 269–276. [Google Scholar] [CrossRef] [PubMed]
  256. Radke, K.; Sigala, T.J.; Grossman, D. Transcription of bovine leukemia virus in peripheral blood cells obtained during early infection in vivo. Microb. Pathog. 1992, 12, 319–331. [Google Scholar] [CrossRef]
  257. Rovnak, J.; Casey, J.W. Assessment of bovine leukemia virus transcripts in vivo. J. Virol. 1999, 73, 8890–8897. [Google Scholar] [PubMed]
  258. Yao, Y.; Smith, L.P.; Nair, V.; Watson, M. An avian retrovirus uses canonical expression and processing mechanisms to generate viral microRNA. J. Virol. 2014, 88, 2–9. [Google Scholar] [CrossRef] [PubMed]
  259. Bennasser, Y.; Le, S.Y.; Yeung, M.L.; Jeang, K.T. HIV-1 encoded candidate micro-RNAs and their cellular targets. Retrovirology 2004, 1, 43. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  260. Couturier, J.P.; Root-Bernstein, R.S. HIV may produce inhibitory microRNAs (miRNAs) that block production of CD28, CD4 and some interleukins. J. Theor. Biol. 2005, 235, 169–184. [Google Scholar] [CrossRef] [PubMed]
  261. Omoto, S.; Ito, M.; Tsutsumi, Y.; Ichikawa, Y.; Okuyama, H.; Brisibe, E.A.; Saksena, N.K.; Fujii, Y.R. HIV-1 nef suppression by virally encoded microRNA. Retrovirology 2004, 1, 44. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  262. Omoto, S.; Fujii, Y.R. Regulation of human immunodeficiency virus 1 transcription by nef microRNA. J. Gen. Virol. 2005, 86, 751–755. [Google Scholar] [CrossRef] [PubMed]
  263. Kaul, D.; Ahlawat, A.; Gupta, S.D. HIV-1 genome-encoded hiv1-mir-H1 impairs cellular responses to infection. Mol. Cell. Biochem. 2009, 323, 143–148. [Google Scholar] [CrossRef] [PubMed]
  264. Lamers, S.L.; Fogel, G.B.; McGrath, M.S. HIV-miR-H1 evolvability during HIV pathogenesis. Biosystems 2010, 101, 88–96. [Google Scholar] [CrossRef] [PubMed]
  265. Klase, Z.; Kale, P.; Winograd, R.; Gupta, M.V.; Heydarian, M.; Berro, R.; McCaffrey, T.; Kashanchi, F. HIV-1 TAR element is processed by Dicer to yield a viral micro-RNA involved in chromatin remodeling of the viral LTR. BMC Mol. Biol. 2007, 8, 63. [Google Scholar] [CrossRef] [PubMed]
  266. Klase, Z.; Winograd, R.; Davis, J.; Carpio, L.; Hildreth, R.; Heydarian, M.; Fu, S.; McCaffrey, T.; Meiri, E.; Ayash-Rashkovsky, M.; et al. HIV-1 TAR miRNA protects against apoptosis by altering cellular gene expression. Retrovirology 2009, 6, 18. [Google Scholar] [CrossRef] [PubMed]
  267. Ouellet, D.L.; Plante, I.; Landry, P.; Barat, C.; Janelle, M.E.; Flamand, L.; Tremblay, M.J.; Provost, P. Identification of functional microRNAs released through asymmetrical processing of HIV-1 TAR element. Nucleic Acids Res. 2008, 36, 2353–2365. [Google Scholar] [CrossRef] [PubMed]
  268. Ouellet, D.L.; Vigneault-Edwards, J.; Letourneau, K.; Gobeil, L.A.; Plante, I.; Burnett, J.C.; Rossi, J.J.; Provost, P. Regulation of host gene expression by HIV-1 TAR microRNAs. Retrovirology 2013, 10, 86. [Google Scholar] [CrossRef] [PubMed]
  269. Narayanan, A.; Iordanskiy, S.; Das, R.; Van Duyne, R.; Santos, S.; Jaworski, E.; Guendel, I.; Sampey, G.; Dalby, E.; Iglesias-Ussel, M.; et al. Exosomes derived from HIV-1-infected cells contain trans-activation response element RNA. J. Biol. Chem. 2013, 288, 20014–20033. [Google Scholar] [CrossRef] [PubMed]
  270. Zhang, Y.; Fan, M.; Geng, G.; Liu, B.; Huang, Z.; Luo, H.; Zhou, J.; Guo, X.; Cai, W.; Zhang, H. A novel HIV-1-encoded microRNA enhances its viral replication by targeting the TATA box region. Retrovirology 2014, 11, 23. [Google Scholar] [CrossRef] [PubMed]
  271. Schopman, N.C.; Willemsen, M.; Liu, Y.P.; Bradley, T.; van Kampen, A.; Baas, F.; Berkhout, B.; Haasnoot, J. Deep sequencing of virus-infected cells reveals HIV-encoded small RNAs. Nucleic Acids Res. 2012, 40, 414–427. [Google Scholar] [CrossRef] [PubMed]
  272. Lee, R.C.; Feinbaum, R.L.; Ambros, V. The C. elegans heterochronic gene lin-4 encodes small RNAs with antisense complementarity to lin-14. Cell 1993, 75, 843–854. [Google Scholar] [CrossRef]
  273. Cullen, B.R. How do viruses avoid inhibition by endogenous cellular microRNAs? PLoS Pathog. 2013, 9, e1003694. [Google Scholar] [CrossRef] [PubMed]
  274. Brown, B.D.; 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. [Google Scholar] [CrossRef] [PubMed]
  275. Mullokandov, G.; Baccarini, A.; Ruzo, A.; Jayaprakash, A.D.; Tung, N.; Israelow, B.; Evans, M.J.; Sachidanandam, R.; Brown, B.D. High-throughput assessment of microRNA activity and function using microRNA sensor and decoy libraries. Nat. Methods 2012, 9, 840–846. [Google Scholar] [CrossRef] [PubMed]
  276. De Boer, R.J.; Ribeiro, R.M.; Perelson, A.S. Current estimates for HIV-1 production imply rapid viral clearance in lymphoid tissues. PLoS Comput. Biol. 2010, 6, e1000906. [Google Scholar] [CrossRef] [PubMed]
  277. Bogerd, H.P.; Kennedy, E.M.; Whisnant, A.W.; Cullen, B.R. Induced Packaging of Cellular MicroRNAs into HIV-1 Virions Can Inhibit Infectivity. MBio 2017, 8, e02125-16. [Google Scholar] [CrossRef] [PubMed]
  278. Selbach, M.; Schwanhausser, B.; Thierfelder, N.; Fang, Z.; Khanin, R.; Rajewsky, N. Widespread changes in protein synthesis induced by microRNAs. Nature 2008, 455, 58–63. [Google Scholar] [CrossRef] [PubMed]
  279. Hausser, J.; Zavolan, M. Identification and consequences of miRNA-target interactions—Beyond repression of gene expression. Nat. Rev. Genet. 2014, 15, 599–612. [Google Scholar] [CrossRef] [PubMed]
  280. Ruegger, S.; Grosshans, H. MicroRNA turnover: When, how, and why. Trends Biochem. Sci. 2012, 37, 436–446. [Google Scholar] [CrossRef] [PubMed]
  281. Aguado, L.C.; Schmid, S.; Sachs, D.; Shim, J.V.; Lim, J.K.; tenOever, B.R. microRNA Function Is Limited to Cytokine Control in the Acute Response to Virus Infection. Cell Host Microbe 2015, 18, 714–722. [Google Scholar] [CrossRef] [PubMed]
  282. Ebert, M.S.; Neilson, J.R.; Sharp, P.A. MicroRNA sponges: Competitive inhibitors of small RNAs in mammalian cells. Nat. Methods 2007, 4, 721–726. [Google Scholar] [CrossRef] [PubMed]
  283. Ebert, M.S.; Sharp, P.A. MicroRNA sponges: Progress and possibilities. RNA 2010, 16, 2043–2050. [Google Scholar] [CrossRef] [PubMed]
  284. Denzler, R.; Agarwal, V.; Stefano, J.; Bartel, D.P.; Stoffel, M. Assessing the ceRNA hypothesis with quantitative measurements of miRNA and target abundance. Mol. Cell 2014, 54, 766–776. [Google Scholar] [CrossRef] [PubMed]
  285. Broderick, J.A.; Zamore, P.D. Competitive endogenous RNAs cannot alter microRNA function in vivo. Mol. Cell 2014, 54, 711–713. [Google Scholar] [CrossRef] [PubMed]
  286. Thomson, D.W.; Dinger, M.E. Endogenous microRNA sponges: Evidence and controversy. Nat. Rev. Genet. 2016, 17, 272–283. [Google Scholar] [CrossRef] [PubMed]
  287. Cazalla, D.; Yario, T.; Steitz, J.A. Down-regulation of a host microRNA by a Herpesvirus saimiri noncoding RNA. Science 2010, 328, 1563–1566. [Google Scholar] [CrossRef] [PubMed]
  288. Liu, N.; Zhang, J.; Jiao, T.; Li, Z.; Peng, J.; Cui, Z.; Ye, X. Hepatitis B virus inhibits apoptosis of hepatoma cells by sponging the MicroRNA 15a/16 cluster. J. Virol. 2013, 87, 13370–13378. [Google Scholar] [CrossRef] [PubMed]
  289. Libri, V.; Helwak, A.; Miesen, P.; Santhakumar, D.; Borger, J.G.; Kudla, G.; Grey, F.; Tollervey, D.; Buck, A.H. Murine cytomegalovirus encodes a miR-27 inhibitor disguised as a target. Proc. Natl. Acad. Sci. USA 2012, 109, 279–284. [Google Scholar] [CrossRef] [PubMed]
  290. Marcinowski, L.; Tanguy, M.; Krmpotic, A.; Radle, B.; Lisnic, V.J.; Tuddenham, L.; Chane-Woon-Ming, B.; Ruzsics, Z.; Erhard, F.; Benkartek, C.; et al. Degradation of cellular mir-27 by a novel, highly abundant viral transcript is important for efficient virus replication in vivo. PLoS Pathog. 2012, 8, e1002510. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  291. Lee, S.; Song, J.; Kim, S.; Kim, J.; Hong, Y.; Kim, Y.; Kim, D.; Baek, D.; Ahn, K. Selective degradation of host MicroRNAs by an intergenic HCMV noncoding RNA accelerates virus production. Cell Host Microbe 2013, 13, 678–690. [Google Scholar] [CrossRef] [PubMed]
  292. Lee, N.S.; Dohjima, T.; Bauer, G.; Li, H.; Li, M.J.; Ehsani, A.; Salvaterra, P.; Rossi, J. Expression of small interfering RNAs targeted against HIV-1 rev transcripts in human cells. Nat. Biotechnol. 2002, 20, 500–505. [Google Scholar] [CrossRef] [PubMed]
  293. Novina, C.D.; Murray, M.F.; Dykxhoorn, D.M.; Beresford, P.J.; Riess, J.; Lee, S.K.; Collman, R.G.; Lieberman, J.; Shankar, P.; Sharp, P.A. siRNA-directed inhibition of HIV-1 infection. Nat. Med. 2002, 8, 681–686. [Google Scholar] [CrossRef] [PubMed]
  294. Jacque, J.M.; Triques, K.; Stevenson, M. Modulation of HIV-1 replication by RNA interference. Nature 2002, 418, 435–438. [Google Scholar] [CrossRef] [PubMed]
  295. Capodici, J.; Kariko, K.; Weissman, D. Inhibition of HIV-1 infection by small interfering RNA-mediated RNA interference. J. Immunol. 2002, 169, 5196–5201. [Google Scholar] [CrossRef] [PubMed]
  296. Boden, D.; Pusch, O.; Lee, F.; Tucker, L.; Shank, P.R.; Ramratnam, B. Promoter choice affects the potency of HIV-1 specific RNA interference. Nucleic Acids Res. 2003, 31, 5033–5038. [Google Scholar] [CrossRef] [PubMed]
  297. Boden, D.; Pusch, O.; Lee, F.; Tucker, L.; Ramratnam, B. Human immunodeficiency virus type 1 escape from RNA interference. J. Virol. 2003, 77, 11531–11535. [Google Scholar] [CrossRef] [PubMed]
  298. Pusch, O.; Boden, D.; Silbermann, R.; Lee, F.; Tucker, L.; Ramratnam, B. Nucleotide sequence homology requirements of HIV-1-specific short hairpin RNA. Nucleic Acids Res. 2003, 31, 6444–6449. [Google Scholar] [CrossRef] [PubMed]
  299. Das, A.T.; Brummelkamp, T.R.; Westerhout, E.M.; Vink, M.; Madiredjo, M.; Bernards, R.; Berkhout, B. Human immunodeficiency virus type 1 escapes from RNA interference-mediated inhibition. J. Virol. 2004, 78, 2601–2605. [Google Scholar] [CrossRef] [PubMed]
  300. Westerhout, E.M.; Ooms, M.; Vink, M.; Das, A.T.; Berkhout, B. HIV-1 can escape from RNA interference by evolving an alternative structure in its RNA genome. Nucleic Acids Res. 2005, 33, 796–804. [Google Scholar] [CrossRef] [PubMed]
  301. ter Brake, O.; Konstantinova, P.; Ceylan, M.; Berkhout, B. Silencing of HIV-1 with RNA interference: A multiple shRNA approach. Mol. Ther. 2006, 14, 883–892. [Google Scholar] [CrossRef] [PubMed]
  302. Westerhout, E.M.; Berkhout, B. A systematic analysis of the effect of target RNA structure on RNA interference. Nucleic Acids Res. 2007, 35, 4322–4330. [Google Scholar] [CrossRef] [PubMed]
  303. Nevot, M.; Martrus, G.; Clotet, B.; Martinez, M.A. RNA interference as a tool for exploring HIV-1 robustness. J. Mol. Biol. 2011, 413, 84–96. [Google Scholar] [CrossRef] [PubMed]
  304. Shah, P.S.; Pham, N.P.; Schaffer, D.V. HIV develops indirect cross-resistance to combinatorial RNAi targeting two distinct and spatially distant sites. Mol. Ther. 2012, 20, 840–848. [Google Scholar] [CrossRef] [PubMed]
  305. Liu, C.; Liang, Z.; Kong, X. Efficacy Analysis of Combinatorial siRNAs against HIV Derived from One Double Hairpin RNA Precursor. Front. Microbiol. 2017, 8, 1651. [Google Scholar] [CrossRef] [PubMed]
  306. Didiano, D.; Hobert, O. Perfect seed pairing is not a generally reliable predictor for miRNA-target interactions. Nat. Struct. Mol. Biol. 2006, 13, 849–851. [Google Scholar] [CrossRef] [PubMed]
  307. Baek, D.; Villen, J.; Shin, C.; Camargo, F.D.; Gygi, S.P.; Bartel, D.P. The impact of microRNAs on protein output. Nature 2008, 455, 64–71. [Google Scholar] [CrossRef] [PubMed]
  308. Cloonan, N. Re-thinking miRNA-mRNA interactions: Intertwining issues confound target discovery. Bioessays 2015, 37, 379–388. [Google Scholar] [CrossRef] [PubMed]
  309. Long, D.; Lee, R.; Williams, P.; Chan, C.Y.; Ambros, V.; Ding, Y. Potent effect of target structure on microRNA function. Nat. Struct. Mol. Biol. 2007, 14, 287–294. [Google Scholar] [CrossRef] [PubMed]
  310. Marin, R.M.; Vanicek, J. Efficient use of accessibility in microRNA target prediction. Nucleic Acids Res. 2011, 39, 19–29. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Schematic representation of the canonical microRNA (miRNA) biogenesis pathway in humans: The RNA Polymerase II transcribes the nuclear chromosome-encoded miRNA gene into primary miRNA (pri-miRNA) that comprises a stem, a terminal loop, and single-stranded 5′ and 3′ RNA tails. The nuclear microprocessor complex-resident Drosha, in concert with its cofactor DGCR8, recognizes the pri-miRNA and cleaves asymmetrically at defined sites in the stem. The released stem-loop structure termed precursor miRNA (pre-miRNA) is then transported to the cytoplasm by the Exportin 5-RanGTP protein complex. The cytoplasmic Dicer, in concert with potential cofactors like transactivating response RNA-binding protein (TRBP) and protein activator of the interferon-induced protein kinase (PACT), recognizes the 2-nucleotide 3′ overhang and/or the 5′ phosphorylated end of the pre-miRNA and cleaves at a defined distance from those termini. The ensuing asymmetric ~22-nucleotides-long miRNA duplex, marked by the 5′ monophosphate and a 2-nt 3′ overhang on both strands, is then loaded onto the AGO protein complex in an ATP-dependent process aided by chaperones and co-chaperones. The miRNA duplexes are then unwound and one of the strands (passenger strand), selected on the basis of 5′ nucleotide composition and thermodynamic stability, concomitantly undergoes dissociation or degradation. The resultant mature miRNA (guide strand), in the context of the effector nucleoprotein complex called miRNA-induced silencing complex (miRISC), then binds the target mRNA leading to translational suppression or mRNA deadenylation or mRNA degradation.
Figure 1. Schematic representation of the canonical microRNA (miRNA) biogenesis pathway in humans: The RNA Polymerase II transcribes the nuclear chromosome-encoded miRNA gene into primary miRNA (pri-miRNA) that comprises a stem, a terminal loop, and single-stranded 5′ and 3′ RNA tails. The nuclear microprocessor complex-resident Drosha, in concert with its cofactor DGCR8, recognizes the pri-miRNA and cleaves asymmetrically at defined sites in the stem. The released stem-loop structure termed precursor miRNA (pre-miRNA) is then transported to the cytoplasm by the Exportin 5-RanGTP protein complex. The cytoplasmic Dicer, in concert with potential cofactors like transactivating response RNA-binding protein (TRBP) and protein activator of the interferon-induced protein kinase (PACT), recognizes the 2-nucleotide 3′ overhang and/or the 5′ phosphorylated end of the pre-miRNA and cleaves at a defined distance from those termini. The ensuing asymmetric ~22-nucleotides-long miRNA duplex, marked by the 5′ monophosphate and a 2-nt 3′ overhang on both strands, is then loaded onto the AGO protein complex in an ATP-dependent process aided by chaperones and co-chaperones. The miRNA duplexes are then unwound and one of the strands (passenger strand), selected on the basis of 5′ nucleotide composition and thermodynamic stability, concomitantly undergoes dissociation or degradation. The resultant mature miRNA (guide strand), in the context of the effector nucleoprotein complex called miRNA-induced silencing complex (miRISC), then binds the target mRNA leading to translational suppression or mRNA deadenylation or mRNA degradation.
Viruses 10 00110 g001
Figure 2. Schematic highlighting the binding sites of cellular miRNAs reported to directly target and bind HIV-1 RNA, and discussed in this review. The cellular miRNAs that are also incorporated into the HIV-1 particles are underlined.
Figure 2. Schematic highlighting the binding sites of cellular miRNAs reported to directly target and bind HIV-1 RNA, and discussed in this review. The cellular miRNAs that are also incorporated into the HIV-1 particles are underlined.
Viruses 10 00110 g002
Figure 3. Schematic showing different stages of HIV-1 replication in a target cell and the potential site(s) of miRISC action. While the incoming HIV-1 RNA genome is protected by the viral capsid shell, thus precluding any access to the miRISC, the viral transcripts exported from the nucleus into the cytoplasm are potentially accessible and hence are viable targets of miRISC. However, it remains unclear whether the endogenous levels of cellular miRNAs are sufficient to effectively counter the high number of viral transcripts typically produced during HIV-1 infection. Yellow arrow = Indicates/depicts progression of the different stages of HIV-1 replication. Black arrow = Indicates/depicts the potential interaction of miRNAs with the HIV-1 RNAs. Black T bar = Indicates/depicts the protection provided by the viral capsid shell to the viral RNA genome against the cellular miRNAs.
Figure 3. Schematic showing different stages of HIV-1 replication in a target cell and the potential site(s) of miRISC action. While the incoming HIV-1 RNA genome is protected by the viral capsid shell, thus precluding any access to the miRISC, the viral transcripts exported from the nucleus into the cytoplasm are potentially accessible and hence are viable targets of miRISC. However, it remains unclear whether the endogenous levels of cellular miRNAs are sufficient to effectively counter the high number of viral transcripts typically produced during HIV-1 infection. Yellow arrow = Indicates/depicts progression of the different stages of HIV-1 replication. Black arrow = Indicates/depicts the potential interaction of miRNAs with the HIV-1 RNAs. Black T bar = Indicates/depicts the protection provided by the viral capsid shell to the viral RNA genome against the cellular miRNAs.
Viruses 10 00110 g003
Table 1. Details of cellular miRNAs reported to target HIV-1 RNA, and discussed in this review.
Table 1. Details of cellular miRNAs reported to target HIV-1 RNA, and discussed in this review.
miRNASequence (5′-3′)Target Region on HIV-1 RNATargeted Sequence on HIV-1 RNA (5′-3′)Effect on HIV-1 ReplicationReference
1miR-149-5pUCUGGCUCCGUGUCUUCACUCCC GagGGGAGTGGGGGGACCCGGCCATANegativeHouzet et al. 2012 [221]
2miR-423-3pAGCUCGGUCUGAGGCCCCUCAGUGagUAUAAAACUCUAAGAGCCGAGCANegativeWhisnant et al. 2013 [222]
3miR-92a-3pUAUUGCACUUGUCCCGGCCUGU PolGCATTGGGAATCATTCAAGCACAACCANegativeHouzet et al. 2012 [221]
4miR-155-5pUUAAUGCUAAUCGUGAUAGGGGU VifGUACUUGGCACUAGCAGCAUUAA NegativeWhisnant et al. 2013 [222]
5miR-133bUUUGGUCCCCUUCAACCAGCUA EnvGACCTGGATGGAGTGGGACAGANegativeHouzet et al. 2012 [221]
6miR-138-5pAGCUGGUGUUGUGAAUCAGGCCG EnvATTGAAGAATCGCAAAACCAGCANegativeHouzet et al. 2012 [221]
7miR-125b-5pUCCCUGAGACCCUAACUUGUGA EnvACGAGGAUUGUGGAACUUCUGGGACGCAGGGGNegativeHuang et al. 2007 [179]
8miR-150-5pUCUCCCAACCCUUGUACCAGUG EnvUCUGGGACGCAGGGGGUGGGAANegativeHuang et al. 2007 [179]; Wang et al. 2009 [223]
9miR-1290UGGAUUUUUGGAUCAGGGANef/3′UTRUCCUUGAUCTGUGGAUCUANegativeWang et al. 2015 [224]
10miR-196bUAGGUAGUUUCCUGUUGUUGGGNef/3′UTRCUACCACACACAAGGCUACUUCNegativeWang et al. 2015 [224]
11miR-223-3pUGUCAGUUUGUCAAAUACCCCA Nef/3′UTRAGGGGUCAGAUAUCCACUGACCNegativeHuang et al. 2007 [179]; Sun et al. 2012 [225]; Wang et al. 2009 [223]
12miR-29b-3pUAGCACCAUUUGAAAUCAGUGUU Nef/3′UTRAUACCCACUGACCUUUGGAUGGUGCUUNegativeSun et al. 2012 [225]
13miR-29a-3pUAGCACCAUCUGAAAUCGGUUA Nef/3′UTRCACUGACCUUUGGAUGGUGCUUNegativeAhluwalia et al. 2008 [220]; Nathans et al. 2009 [180]
14miR-326CCUCUGGGCCCUUCCUCCAG Nef/3′UTRTGGGATGGAGGACCCGGAGGNegativeHouzet et al. 2012 [221]
15miR-382-5pGAAGUUGUUCGUGGUGGAUUCG U3/3′UTRCGAGCUUGCUACAAGGGACUUUNegativeHuang et al. 2007 [179]; Wang et al. 2009 [223]
16miR-28-5pAAGGAGCUCACAGUCUAUUGAG U3/3′UTRAUCUGAGCCUGGGAGCUCUCNegativeHuang et al. 2007 [179]; Wang et al. 2009 [223]
Table 2. Cellular miRNAs reported to target the host-dependency factors (HDFs) and modulate HIV-1 replication negatively or positively, and discussed in this review.
Table 2. Cellular miRNAs reported to target the host-dependency factors (HDFs) and modulate HIV-1 replication negatively or positively, and discussed in this review.
miRNATarget HDF/Cellular mRNAEffect on HIV-1 ReplicationReference
1miR-17-5pPCAF 3′UTRReduction in HIV-1 infection in PBMCs and Jurkat cellsTriboulet et al. 2007 [181]
2miR-20aPCAF 3′UTRReduction in HIV-1 infection in PBMCs and Jurkat cellsTriboulet et al. 2007 [181]
3miR-198Cyclin T1 3′UTRImpaired HIV-1 replication in monocytesSung and Rice, 2009 [182]
4miR-15aPur-alpha 3′UTRImpaired HIV-1 replication in monocytesShen et al. 2012 [229]
5miR-15bPur-alpha 3′UTRImpaired HIV-1 replication in monocytesShen et al. 2012 [229]
6miR-16Pur-alpha 3′UTRImpaired HIV-1 replication in monocytesShen et al. 2012 [229]
7miR-20aPur-alpha 3′UTRImpaired HIV-1 replication in monocytesShen et al. 2012 [229]
8miR-93Pur-alpha 3′UTRImpaired HIV-1 replication in monocytesShen et al. 2012 [229]
9miR-106bPur-alpha 3′UTRImpaired HIV-1 replication in monocytesShen et al. 2012 [229]
10miR-155ADAM10 3′UTR, NUP153 3′UTR, LEDGF/p75 3′UTRTRIM32 3′UTR
SAMHD1
Reduction in HIV-1 late RT products and viral DNA integration in MDMs
Promotes reactivation of latent HIV-1 via NF-kB signaling in J-Lat 5A8 cells
Overexpression of miR-155 enhanced HIV-1 replication in astrocytes
Swaminathan et al. 2012 [232]
Ruelas et al. 2015 [233]
Pilakka-Kanthikeel et al. 2015 [235]
11miR-27bCyclin T1 3′UTRImpaired HIV-1 replication in resting CD4+ T-cellsChiang et al. 2012 [183]
12miR-29bCyclin T1 3′UTRImpaired HIV-1 replication in resting CD4+ T-cellsChiang et al. 2012 [183]
13miR-150Cyclin T1 3′UTRImpaired HIV-1 replication in resting CD4+ T-cellsChiang et al. 2012 [183]
14miR-223Cyclin T1 3′UTRImpaired HIV-1 replication in resting CD4+ T-cellsChiang et al. 2012 [183]
15miR-1236VprBP 3′UTRImpaired HIV-1 replication in monocytesMa et al. 2014 [231]
16miR-181aSAMHD1Overexpression of miR-181a enhanced HIV-1 replication in astrocytes Pilakka-Kanthikeel et al. 2015 [235]
17miR-132MeCP2Increased HIV-1 infection in Jurkat T-cellsChiang et al. 2013 [178]
18Let-7cP21 3′UTRIncreased HIV-1 replication in JLTRG-R5 and HeLa-CCR5 cellsFarberov et al. 2015 [190]
19miR-124aTASK1 3′UTRIncreased HIV-1 replication in JLTRG-R5 and HeLa-CCR5 cellsFarberov et al. 2015 [190]
20miR-34a-5pTASK1 3′UTRIncreased HIV-1 replication in JLTRG-R5 and HeLa-CCR5 cellsFarberov et al. 2015 [190]
21miR-34c-5pSeveral genes involved in TCR signaling and activation of naïve CD4+ T-cellsIncreased HIV-1 replication in Jurkat T-cellsAmaral et al. 2017 [191]
22miR-221CD4 3′UTRInhibition of HIV-1 entry in macrophagesLodge et al. 2017 [234]
23miR-222CD4 3′UTRInhibition of HIV-1 entry in macrophagesLodge et al. 2017 [234]

Share and Cite

MDPI and ACS Style

Balasubramaniam, M.; Pandhare, J.; Dash, C. Are microRNAs Important Players in HIV-1 Infection? An Update. Viruses 2018, 10, 110. https://0-doi-org.brum.beds.ac.uk/10.3390/v10030110

AMA Style

Balasubramaniam M, Pandhare J, Dash C. Are microRNAs Important Players in HIV-1 Infection? An Update. Viruses. 2018; 10(3):110. https://0-doi-org.brum.beds.ac.uk/10.3390/v10030110

Chicago/Turabian Style

Balasubramaniam, Muthukumar, Jui Pandhare, and Chandravanu Dash. 2018. "Are microRNAs Important Players in HIV-1 Infection? An Update" Viruses 10, no. 3: 110. https://0-doi-org.brum.beds.ac.uk/10.3390/v10030110

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop