Next Article in Journal
Application of Human Stem Cells to Model Genetic Sensorineural Hearing Loss and Meniere Disease
Next Article in Special Issue
Single-Cell Transcriptomics of Mtb/HIV Co-Infection
Previous Article in Journal
Substances of Natural Origin in Medicine: Plants vs. Cancer
Previous Article in Special Issue
Transcriptomic Classification of Pituitary Neuroendocrine Tumors Causing Acromegaly
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Aid or Antagonize: Nuclear Long Noncoding RNAs Regulate Host Responses and Outcomes of Viral Infections

1
Disease Intervention and Prevention Program, Texas Biomedical Research Institute, San Antonio, TX 78227, USA
2
Host-Pathogen Interaction Program, Texas Biomedical Research Institute, San Antonio, TX 78227, USA
*
Author to whom correspondence should be addressed.
Cells 2023, 12(7), 987; https://doi.org/10.3390/cells12070987
Submission received: 6 October 2022 / Revised: 12 March 2023 / Accepted: 15 March 2023 / Published: 23 March 2023
(This article belongs to the Special Issue Transcriptome in Human Disease)

Abstract

:
Long noncoding RNAs (lncRNAs) are transcripts measuring >200 bp in length and devoid of protein-coding potential. LncRNAs exceed the number of protein-coding mRNAs and regulate cellular, developmental, and immune pathways through diverse molecular mechanisms. In recent years, lncRNAs have emerged as epigenetic regulators with prominent roles in health and disease. Many lncRNAs, either host or virus-encoded, have been implicated in critical cellular defense processes, such as cytokine and antiviral gene expression, the regulation of cell signaling pathways, and the activation of transcription factors. In addition, cellular and viral lncRNAs regulate virus gene expression. Viral infections and associated immune responses alter the expression of host lncRNAs regulating immune responses, host metabolism, and viral replication. The influence of lncRNAs on the pathogenesis and outcomes of viral infections is being widely explored because virus-induced lncRNAs can serve as diagnostic and therapeutic targets. Future studies should focus on thoroughly characterizing lncRNA expressions in virus-infected primary cells, investigating their role in disease prognosis, and developing biologically relevant animal or organoid models to determine their suitability for specific therapeutic targeting. Many cellular and viral lncRNAs localize in the nucleus and epigenetically modulate viral transcription, latency, and host responses to infection. In this review, we provide an overview of the role of nuclear lncRNAs in the pathogenesis and outcomes of viral infections, such as the Influenza A virus, Sendai Virus, Respiratory Syncytial Virus, Hepatitis C virus, Human Immunodeficiency Virus, and Herpes Simplex Virus. We also address significant advances and barriers in characterizing lncRNA function and explore the potential of lncRNAs as therapeutic targets.

1. Introduction

During the past decade, several genome-wide RNAi [1,2,3,4,5,6,7,8,9,10,11,12,13,14] and CRISPR [15,16,17,18,19,20,21,22] studies have identified host proteins critical for viral replication. Protein-coding open reading frames constitute less than 2% of the human genome [23], and the bulk of the transcriptome is noncoding RNA (ncRNA). Among the different ncRNA subclasses, long noncoding RNAs (lncRNAs) were reported to constitute nearly 68% of the transcriptome [24]. Many lncRNAs regulate various cellular processes [25,26] and are emerging as versatile regulators of gene expression with prominent roles in health and disease [27,28].
A variety of viral infections alter host lncRNA expressions. Dramatic changes in lncRNA expressions have been observed in the cells infected with Influenza A Virus (IAV) [29], Sendai Virus (Sev), Respiratory Syncytial Virus (RSV) [30], Hepatitis C Virus (HCV) [31], Adenovirus [32], Human Papilloma Virus (HPV) [33], pathogenic human enterovirus [34], Severe Acute Respiratory Syndrome Coronavirus (SARS-CoV and SARS-CoV-2) [35,36,37,38,39,40,41,42,43,44,45], Human Immunodeficiency Virus (HIV) [46,47,48,49,50,51], Muscovy Duck ReoVirus (MDRV), and Herpes Simplex Virus (HSV) [52]. These studies are only beginning to reveal the myriad changes in lncRNA expression upon viral infection and indicate the role of lncRNAs in viral pathogenesis.
The pathogen-associated molecular patterns (PAMPs) and damage-associated molecular patterns (DAMPs) released during infections, stress, and non-programmed cell death are detected by the pattern recognition receptors (PRRs), such as Toll-like receptors (TLR), retinoic acid-inducible gene (RIG)-like receptors (RLRs), nucleotide oligomerization domain (NOD)-like Receptors (NLRs), and C-type lectin receptors (CLRs). The activation of PRRs leads to the transcription of inflammatory genes induced by ATF2 and NF-κB, the transcription of specific antiviral genes induced by IRF3 and IRF7, and the synthesis of Type I interferons (IFN-I). PRR activation by virus infection [53,54,55,56,57,58] and various ligands, such as lipopolysaccharides (TLR4) or Poly I: C induces the expression of lncRNAs [59,60,61,62,63,64]. In addition, stimulating cells by cytokines, such as IFN-I [65] and TNF-α [66], induces differential expressions of lncRNAs. Viral infections, including specific viral proteins can upregulate the expression of stress-induced and other lncRNAs [67,68]. Thus, the transcriptome of virus-infected cells presents an opportunity to discover and characterize novel lncRNAs that may play a significant role in cellular defense, immune response, and viral propagation. For instance, IFN–alpha (IFNα) stimulation or infection with RNA viruses upregulates lncRNA ISIR [69]. ISIR activates the Interferon Regulatory Factor-3 (IRF3) and strengthens the interferon response to viral infections. Indeed, several transcriptomic studies have highlighted the marked dysregulation of lncRNAs in virally infected cells [70]. We have listed prominent nuclear RNAs with a significant impact on immune response, viral replication, and latency in Table 1. We also discuss the molecular mechanisms of their action wih more details in this review.
Overlapping patterns in lncRNA expression, in response to virus infections, suggest the functional role of lncRNAs in the clinical manifestations of these infections. Similarly, intersecting changes in global lncRNA expression patterns in SARS-CoV and influenza virus infection indicate a lncRNA-based signature of respiratory virus infection and a functional role for the virus-induced lncRNAs in clinical outcomes [95]. After performing a whole-transcriptome analysis of the host response to severe acute respiratory syndrome coronavirus (SARS-CoV) infection across four founder mouse strains, Peng et al. found several noncoding RNAs to be similarly regulated in SARS-CoV and influenza virus-infected mice [95]. However, the functional mechanisms and impact of these overlapping patterns of lncRNAs in respiratory viral infections are yet to be determined. In addition, the virus-induced lncRNAs may have diagnostic potential, such as the antiviral lncRNA EDAL induced by multiple neurotropic viruses in mice [96]. Thus, careful analyses of lncRNA expression and function in infected cells are critical to improving our understanding of viral pathogenesis.
Nevertheless, diversity in the form and function of lncRNAs makes them both intriguing and challenging to study. The human DNA is predicted to encode over 100,000 lncRNAs [97,98], but only a tiny fraction of these have been characterized. Several types of lncRNAs are described in the literature, including but not limited to lncRNAs transcribed from intergenic, enhancer, and promoter regions, as well as sense and antisense lncRNAs that overlap other protein-coding genes [23,99,100]. Most lncRNAs are expressed at lower levels than protein-coding mRNAs. Several factors, including repressive histone modifications at lncRNA gene promoters [101,102], transcription through phosphorylation-deficient Polymerase II (pol II), weak or aberrant splicing, and termination contribute to lower transcription levels of lncRNAs than the protein-coding mRNAs. Additionally, degradation by nuclear exosomes leads to overall diminished expression levels of most lncRNAs [103]. Nevertheless, the expression levels of several lncRNAs are either similar or even exceed those of protein-coding mRNAs. For example, MALAT1 and NEAT1 are expressed ubiquitously at high levels in most cells [104].
A significant fraction of lncRNA is preferentially localized in the nucleus. This nuclear retention is due to specific sequence motifs encoded within some lncRNAs, such as Alu repeats [105], AGCCC motif of BORG [106], E and M fragments of MALAT1 [107], repeating RNA domain (RDD) of human FIRRE [108], and retained introns in the nuclear TUG1 [109]. The nuclear retention of Kaposi Sarcoma-associated Herpes Virus (KSHV)-encoded lncRNA and PAN RNA (polyadenylated nuclear RNA) is dependent on the presence of an expression and nuclear retention element (ENE) [110]. The ENE contains a uracil (U)-rich internal loop that interacts with the 3′-poly(A) tail to form a triple helical loop to protect PAN RNA from a rapid nuclear deadenylation-dependent decay by exonucleases [111,112,113]. In addition, viral ORF57 stabilizes the expression of PAN RNA and increases its nuclear accumulation [88,114,115]. RNA-stabilizing ENE-like structures are also found in human lncRNAs, such as MALAT1 and MENβ [116]. Protein-coding mRNAs and many lncRNAs share the mechanisms for posttranscriptional processing, nuclear export, and trafficking within the cells [117]. Like the protein-coding mRNAs, many lncRNAs are exported from the nucleus by the nuclear export complexes, TREX (transcription export complex), and NFX (nuclear RNA export factor) [117,118,119]. However, some lncRNAs are retained in the nucleus due to poor binding to the nuclear export complexes [119,120]. In addition, functional nuclear lncRNAs escape exosome degradation through specific, high-affinity interactions with DNA and proteins that tether them to the chromatin.
Functional nuclear lncRNA transcripts employ diverse mechanisms to regulate gene expression, including but not limited to the recruitment, depletion, or relocalization of chromatin-modifying proteins, transcription factors, and RNA; the direct interaction with DNA; and the regulation of chromatin organization, as well as intra/inter-chromosomal interactions, transcription, and splicing [108,121,122,123,124,125,126,127]. Nuclear lncRNAs have been shown to regulate the localization of splicing factors and impact the expression of antiviral genes [83,128]. LncRNAs are typically expressed at significantly lower levels than protein-coding mRNAs. Per-cell copies of lncRNAs (0.3–1000) are remarkably low compared to their protein binding partners. Recent data show that lncRNAs adopt distinct mechanisms to affect dramatic changes in target gene expression, even with a few copies. A few lncRNA molecules seed and organize functional territory wherein the lncRNAs recruit diffusible RNAs or proteins, thus enriching the functional effectors at a specific genomic site [129,130,131]. Quinodoz et al. showed that most lncRNAs included in their study remain at their target loci close to the lncRNA transcription site and do not diffuse elsewhere in the nucleus or cytoplasm [130]. Markaki et al. further elucidated how only a few lncRNA molecules can initiate the “crowding” of transcriptional regulators at several target loci [129]. Xist lncRNA silences over 1000 genes on the X-chromosome. Markaki et al. showed that just two Xist RNA molecules could recruit a multiprotein structure and increase the local concentration of regulators to silence the transcription of target genes. X-chromosome compaction and densification of a silencer protein, SPEN, induces silencing across the entire X-chromosome. Thus, X-chromosome genes that do not directly interact with Xist RNA are also silenced [129]. Markaki et al. also showed that Xist remains restricted on the X-chromosome [129]. Other lncRNAs, such as MALAT1, interact with several genes far from their transcription site and across chromosomes. These observations raise further questions about how some lncRNAs stay localized to their transcription site, and if any specific features dictate lncRNA movement and localization. For some lncRNA loci, the transcript does not exhibit any regulatory function. However, the process of lncRNA transcription itself may contribute to the regulation of adjacent gene expression by remodeling chromatin or recruiting transcriptional regulatory factors [132]. Thus, nuclear lncRNAs are versatile tools for rapidly activating or suppressing specific genes or gene networks, and some viruses have been shown to hijack this machinery to establish infections successfully.
Here, we review the impact of nuclear lncRNAs on the regulation of gene expression and viral disease outcomes. Most nuclear lncRNAs described here utilize epigenetic mechanisms to regulate gene transcription. Nevertheless, we also have a few nuclear lncRNAs that play a significant role in viral infections by regulating mRNA splicing of the host response genes.

2. Chromatin–lncRNA Interactions Regulate Viral Infections

Chromatin structure plays a critical role in activating and repressing transcription. Many lncRNAs modulate gene expression within the spatial proximity of their transcription site and distant gene networks through either sequence-specific, direct DNA-binding, or indirectly through their chromatin-binding protein partners. Emerging data indicate that lncRNA-chromatin interactions regulate the antiviral interferon (IFN) response, viral transcription, and latency. We have listed several lncRNAs with known antiviral or proviral activity in Table 1. The precise mechanism of action of some lncRNAs remains to be further explored. This section discusses known nuclear lncRNAs regulating immune response and viral transcription through direct interaction with chromatin and histone modification.
Direct interaction: Many nuclear lncRNAs interact with the double-stranded DNA directly in a sequence-specific manner. These interactions form triple helices [133,134,135,136,137,138,139,140] and R-loops [141]. The triple helices recruit coactivator or corepressor proteins to activate [134,136,139] or repress [133,135] gene transcription, respectively. The lncRNA–DNA triple helices can form near the transcription start site (proximal) or at distant regulatory regions (distal). Lnc-MxA is an IFN-induced lncRNA upregulated during IAV infection. Lnc-MxA binds to the IFNB1 promoter forming a triplex, which then interferes with the binding of IRF3 and p65 transcription factors to the IFNB1 promoter, resulting in the abrogation of the IFNβ transcription [61] (Figure 1). Type I IFNs, such as IFNβ, promote the transcriptional activation of hundreds of interferon-stimulated genes (ISGs), many of which inhibit virus replication [142]. Thus, Lnc-MXA enhances viral replication by dampening the interferon response [61].
Elevated expression levels of lncRNA MIR4435-2HG have been reported in primary myeloid-derived dendritic cells (mDCs) isolated from patients who spontaneously controlled HIV replication (elite controllers, ECs) [134]. An elevated expression of MIR4435-2HG increased triple-helix formations at an intronic gene enhancer and enhanced RPTOR1 (Regulatory Associated Protein Of MTOR Complex I) gene expression. Activating chromatin marker H3K27ac is enriched at the site of the triple helix, likely through the specific recruitment of histone acetyltransferases [134]. RPTOR1 increased glycolysis and metabolic activity of mDCs in response to TLR3 stimulation. Hartana et al. identified a role for MIR4435-2HG in enhancing the metabolic activity of mDCs, which is likely to increase the functional responsiveness of mDCs, thereby facilitating more effective immune activity in ECs [134].
Virally encoded lncRNAs also use this mechanism to modulate viral gene transcription. For example, the Epstein–Barr virus (EBV) forms virus-induced nodular structures (VINORCS). VINORCS are composed of viral and cellular proteins required for viral replication. EBV-encoded lncRNA BHLF1 localizes in the viral replication compartment [93], forming an RNA–DNA hybrid at the virus transcription start site [94]. This hybrid structure then recruits RNA-binding proteins to form VINORCs and facilitate selective processing and the export of viral mRNAs, thus enhancing viral replication.
Histone modification: Post-translational histone modifications, such as phosphorylation, acetylation, methylation, ubiquitination, SUMOylation, and GlcNAcylation, are key regulators of the chromatin state and transcriptional activity. Nuclear lncRNAs interact with the proteins that add (writers) [143,144,145], remove (erasers) [144,146,147], or recognize (readers) [148,149] these histone modifications and modulate their functions. Several lncRNAs use this mechanism to regulate the expression of ISGs. The expression of some ISG-regulating lncRNAs is significantly modulated in virus-infected cells. For example, viruses such as IAV, SeV, MDRV, and HSV significantly downregulate lncRNA NRAV, which inhibits the expression of critical ISGs through histone modifications at the promoters of these genes. In this context, an NRAV overexpression leads to reduced H3K4 trimethylation (H3K4me3), activating transcription and enriched repressive H3K27 trimethylation (H3K27me3) at the transcription start sites of ISGs [57]. Although the exact mechanism of how NRAV regulates histone trimethylation is unknown, NRAV is shown to bind a regulatory protein ZONAB, which may be involved in histone modifications. The NRAV-mediated regulation of ISG transcription was attributed, at least partially, to its interaction with ZONAB [57]. Likewise, a host cell-encoded lncRNA induced in response to IAV infection, termed, “Inhibiting IAV Replication by Promoting IFN and ISG Expression” (IVRPIE), enhances the expression of IFN-β and several critical ISGs, including IRF1, IFIT1, IFIT3, MxA, ISG15, and IFI44L, through histone modifications at these loci [63]. MxA and ISG15 can suppress the replication of highly pathogenic Influenza A viruses [150]. Likewise, IFITM1 and IFITM3 can inhibit an early step of Influenza A virus replication [2]. Thus, IVRPI inhibits IAV replication by enhancing the expression of IFN-β and antiviral ISG proteins. HCV-induced lncRNA RP11-288L9.4 inhibits the expression of IFNα-inducible Protein 6 (IFI6) through histone modifications [72].
The host-encoded lncRNA HEAL is expressed at high levels in human macrophages upon HIV infection and binds directly to the HIV promoter, along with an RNA-binding protein fused in liposarcoma (FUS) [85]. The HEAL–FUS complex recruits the histone acetyltransferase p300 to enhance H3K27 acetylation and enriches Transcription Elongation Factor P-TEFb to the HIV promoter, thus increasing HIV transcription [85] (schematic presentation of HEAL–FUS-driven HIV transcription in Figure 1). LncRNA MALAT1 interacts with Polycomb Repressive Complex 2 (PRC2) proteins, namely EZH2, Suz12, and EED, which then catalyze the methylation of Histone H3 at Lysine 27 to repress gene transcription. In tumor cells, MALAT1 facilitates EZH2-binding to its target loci, driving H3K27 trimethylation-mediated repression of multiple tumor-suppressor genes, such as E-cadherin [151,152], NDRG1 [153], p21, and p27 [154]. MALAT1 is expressed at high levels in HIV-infected cells, where it enhances HIV transcription from latent provirus [78]; it localizes EZH2 to its target sites in tumors; and in HIV infection, it sequesters EZH2 away from the HIV long terminal repeat (LTR), thus preventing PRC2-mediated H3K27 trimethylation and promoting HIV viral reactivation [78].
The KSHV-lytic protein (K-Rta) induces the expression of a cellular lncRNA KIKAT/LINC01061 (KSHV-Induced KDM4A-Associated Transcript) [68]. Yang et al. showed that KIKAT interacts with a histone lysine demethylase (KDM4A) and re-localizes KDM4A from the transcription start site (TSS) of the Angiomotin (AMOT) gene. The KIKAT-mediated relocation of KDM4A increased the AMOT transcription and angiomotin-dependent cell migration, thus implicating its role in angiogenesis in Kaposi’s sarcoma [68]. KIKAT was also shown to enhance KSHV reactivation. While KDM4A was shown to regulate KSHV replication [155,156], its binding on the KSHV genome was not impacted by KIKAT [68].
Virus-encoded lncRNAs also utilize histone modification strategies to establish and regulate their latent infection. This mechanism has been demonstrated in HIV infection, where a virus-encoded antisense lncRNA (antisense transcript; Ast) recruits chromatin remodeling proteins such as DNMT3a, EZH2, and HDAC-1 to HIV 5’ long terminal repeat (LTR). These proteins mediate H3K9 dimethylation, H3K27 trimethylation, and histone deacetylation, resulting in the epigenetic silencing of viral transcription [86,157,158].
KSHV remains persistent for a lifetime in patients. KSHV remains latent for a long duration, followed by a short lytic cycle. During latency, the KSHV episome is tethered to the host genome through KSHV latency-associated nuclear antigen (LANA) protein (KSHV life cycle is reviewed in [159]). LANA protein binds the viral episome and the host nucleosomal proteins to tether the viral episome to the cellular genome. The KSHV-encoded lncRNA, PAN RNA ([160], is the most abundant viral transcript [114]. The absence of PAN RNA results in reduced virus production [88,161,162]. The KSHV-lytic protein (K-Rta) induces PAN RNA expression at very high levels [163]. PAN RNA interacts with KSHV latency-associated nuclear antigen (LANA) and sequesters LANA from the viral DNA episomes, thus facilitating lytic reactivation [114,164]. A recent study showed that PAN RNA sequesters LANA-interacting nucleosomal protein CHD4 (chromodoain helicase binding protein 4) [165]. The CHD4 and LANA complex co-localize to the episome, where CHD4 prevents the aggregation of RNA polymerase II on the KSHV episome and inhibits KSHV reactivation [165]. Thus, PAN RNA facilitates KSHV reactivation by the sequestration of the LANA/CHD4 complex from the KSHV episome [165]. PAN RNA encodes two main cis-acting elements, the Mta response element, (MRE) and the expression and nuclear retention element (ENE). Gutierrez et al. found that ENE is not required for viral replication but is essential for the nuclear retention of PAN RNA [110]. Viral protein ORF59 binds to PAN RNA during reactivation, recruiting chromatin-modifying factors to the viral genome [166]. PAN RNA physically interacts with the viral promoter, lysine demethylases UTX and JMJD3, and lysine methyltransferase MLL2 [87]. It recruits histone demethylases to the viral genome [88]. In addition, PAN RNA interaction sites have also been detected on the host genome, suggesting that it potentially interacts with transcriptional regulators and chromatin modifiers to modulate cellular gene expression, immune response, and cell cycle control [88,162].

3. Virus-Induced lncRNAs Regulate the Transcription and Splicing of Host and Viral Genes

Virus-induced lncRNAs modulate antiviral interferon responses by regulating the activation, availability, and localization of transcription factors. Lnc-000641, a pseudorabies virus (PRV)-induced lncRNA, inhibits the phosphorylation of upstream-activating kinases and transcription factors (Jak and STAT1), thereby reducing downstream IFNα transcription, thus facilitating increased PRV replication [73]. Along similar lines, HCV-induced lncRNA, Lethe, interacts with NF-κB subunit RelA and inhibits RelA-mediated DNA binding [66]. This prevents RelA-mediated transcriptions of the activating antiviral factors, 2′,5′-oligoadenylate synthetase (OAS), interferon regulatory factor 1 (IRF1), and protein kinase R (PKR), thus, enhancing HCV replication [71]. LPS-stimulated or virus-infected human dendritic cells (DCs) upregulate the expression of lncRNA LUCAT1, which functions as a potent regulator of the IFN-α/β response [167]. LUCAT1 sequesters STAT1 in the nucleus preventing STAT1 from binding to the promoters of ISGs and blocking their expression [167] (Figure 1). The virus infection or activation through PAMPs can also downregulate the expression of lncRNAs that regulate the immune response. For example, Aznaourova et al. recently showed that SARS-CoV-2 infection or PAMP-mediated stimulation inhibits the expression of a nuclear lncRNA, PIRAT. PIRAT recruits transcription factor PU.1 to pseudogenes and suppresses PU.1 binding to promoters of alarmin genes (S100A8 and S100A), thus inhibiting alarmin gene expression. Alarmins promote the production of inflammatory cytokines. The SARS-CoV-2 infection also enhances the expression of another lncRNA, LUCAT1, which augments alarmin gene transcription. Thus, the SARS-CoV-2 infection upregulates LUCAT1 and downregulates PIRAT, increasing alarmin production and aggravating inflammatory mediators contributing to the severity of COVID-19 [168]. LncRNA LUARIS (a.k.a. lncRNA#32) upregulated the ISG expression through interactions with host proteins HNRNPU and ATF2, resulting in the inhibition of encephalomyocarditis virus (EMCV), Hepatitis B, and Hepatitis C virus replication [76]. While HNRNPU stabilized LUARIS transcript, the LUARIS–ATF2 interaction was found to be critical for activating ISG transcription, indicating that LUARIS enhanced recruitment of the transcription factor ATF2 to the promoters of ISGs, thus enhancing ATF2-mediated transcription (Figure 1). In addition, 7SK snRNA sequesters P-TEFb, a general transcription elongation factor and human co-factor for HIV-1 transactivator (Tat) protein, into the catalytically inactive 7SK snRNP and inhibits HIV transcription [79,80,81,82]. The human T-cell leukemia virus (HTLV)-encoded antisense lncRNA is recruited to the CC chemokine receptor (CCR4) and enhances its transcription to support the proliferation of HTLV-infected cells [169].
Many lncRNAs regulate neighboring genetic loci in a transcript-dependent manner by interfering with the recruitment of transcription factors or Poll II at the promoter, altering chromatin modification or reducing accessibility. LncRNAs can form RNA–DNA triplexes that enrich gene regulatory proteins at the neighboring loci. For example, PTENpg1 localizes on the promoter of its adjacent locus, PTEN, and recruits histone methyl transferases (EZH2 and DNMT3a) to the PTEN promoter, dampening PTEN transcription [170]. LncRNAs also act as a scaffold for the locus-specific recruitment of chromatin-modifying enzymes. For example, lncRNA APOAS1 provides a scaffold for the chromatin-modifying histone demethylase protein, LSD1, to localize the APOA1 gene and repress APOA1 gene expression [171]. HCV infection induces one such lncRNA that regulates transcription of its neighboring protein-coding gene and alters the antiviral immune response. HCV-infected cells express a nuclear lncRNA GCSIR (GPR55 cis-regulatory suppressor of immune response RNA, a.k.a Lnc-ITM2C-1) that enhances the transcription of its neighboring gene, GPR55, through yet unknown molecular mechanisms. The GPR55 protein, in turn, inhibits the expression of several ISGs, thus, dampening antiviral responses [60].
In addition to regulating transcription, neighboring or intragenic lncRNAs can also modulate the splicing of their protein-coding neighbors. RUNX1 transcription and protein expression are tightly controlled by several lncRNAs transcribed from the neighboring loci. RUNX1 represses HIV-1 replication in T cells by binding to the HIV-1 LTR [172]. RUNX1 gene encodes three transcript variants that produce three different protein isoforms, RUNX1a, b, and c. Among the three, RUNX1b and c have been shown to bind the HIV–LTR and suppress HIV transcription [83]. LINC01426 (a.k.a. uc002yug.2) is transcribed from a locus upstream of RUNX1. It enhances the recruitment of splicing factors (MBNL1 and SFRS1) to the regional RNA duplexes resulting in increased RUNXa isoform and the relative reduction of RUNX1b and c isoform expressions [83]. In addition, LINC01426 increases the production of HIV protein Tat through unknown mechanisms. Thus, LINC01426 promotes viral reactivation by inhibiting transcription repressive forms of RUNX1 and enhancing Tat expression [83]. Another lncRNA RUNXOR is transcribed from a promoter upstream of the RUNX1 gene that overlaps with RUNX1 mRNA. RUNXOR increases H3K4me3 marks at the promoter of RUNX1 and activates RUNX1 transcription. Interestingly, myeloid-derived suppressor cells (MDSCs) in people living with HIV (PLWH) showed increased RUNXOR expression (Figure 1). An increased expression of RUNXOR in MDSCs results in the expression of critical immunosuppressive molecules that cause T cell suppression in PLWH [173].
Although there is an increasing number of lncRNAs identified as novel regulators of host–virus interaction, the precise functional mechanisms of many remain unknown. HCV, IAV, and the Semliki Forest virus (SFV) induce a nuclear lncRNA transcript eosinophil granule ontogeny transcript (EGOT). The molecular mechanism of EGOT-mediated suppression of the IFN-signaling pathway and enhancement of viral production is yet to be determined [31]. BISPR is an IFN-stimulated lncRNA that significantly regulates the antiviral BST2 gene expression through yet unknown mechanisms [174,175]. IAV-induced lncRNA TSPOAP1-AS1 inhibits the expression of IFNβ and other ISGs through unknown mechanisms [62]. Interestingly, TSPOAP1-AS1 is localized in the nucleus and cytoplasm, but IAV-infected cells showed increased nuclear levels of TSPOAP1-AS1, implicating a nuclear mechanism of ISG inhibition. ZIKA virus (ZKIV) infection induces lncRNA OASL-IT1. OASL-IT1 enhances IFNβ and ISG (Mx1 and IFITM1) expression and inhibits ZKIV replication through unclear mechanisms [75]. IAV H1N1, H3N2, H7N7 strains, and VSV-infected cells upregulate a nuclear lncRNA VIN. The molecular mechanisms of VIN-mediated upregulation of viral gene expression remain to be determined [29].

4. Heterogeneity in lncRNA Form, Function, and Phenotype

Some lncRNA transcripts utilize diverse mechanisms that occasionally produce contrasting phenotypes highlighting the complexities of their gene regulatory functions. In addition, many lncRNAs are transcribed from genomic regions that encode sequences regulating chromatin structure, which makes investigating molecular mechanisms and interpreting lncRNA functions very challenging in these instances.
This is best exemplified by lncRNA Ifng-as1. In mice and humans, Ifng-as1 [Nettoie Salmonella pas Theiler’s (NEST), also named TMEVPG1] is transcribed from the opposite strand of the IFNG protein-coding region. Ifng-as1 was initially identified as a susceptibility locus for Theiler’s virus persistence in mice [176]. In follow-up studies, the RNAi-mediated knockdown of human IFNG–AS1 expression in human T-helper (Th1) cells and significantly reduced IFNG transcriptions [177,178]. Similarly, the transgenic expression of mouse Ifng-as1 enhanced the IFN-γ expression and established resistance to Salmonella enterica and increased persistence of Theilers’ virus in mice [179]. These contrasting phenotypes indicated the essential role of lncRNAs in modulating immune responses to distinct pathogens and disease outcomes. The ectopically expressed Ifng-as1 transcript binds to a histone methyltransferase complex component (WDR5) and enhances H3K4 trimethylation at the Ifng locus in an in vitro cell line model [179]. These approaches indicated that Ifng-as1 is a trans-acting lncRNA that recruits chromatin modifiers in a sequence-specific manner. A recent study employed CRISPR tools to compare Ifng-as1 knockout (KO; DNA+RNA product deletion) and Ifng-as1-polyA knock-in (KI; truncated non-functional RNA) modification in mice [180]. Deleting DNA (KO) and truncating RNA (KI) inhibited Ifng gene expression, but KO mice showed more severe impairment in defense against infection. This study further determined that deleting the Ifng-as1 locus (KO) eliminates one of the CTCF-binding sites, thus disrupting the chromatin looping required for optimal Ifng gene expression. The truncated Ifng-as1 (KI) does not affect chromatin architecture but diminishes Ifng expression. The Ifng-as1 transcript will likely enhance Ifng expression by recruiting and enriching transcription factors or chromatin modifiers. These critical experiments dissecting the effects of the lncRNA transcript and chromatin structure showed that Ifng-as1 regulates Ifng expression in cis, and Ifng-as1 locus impacts the chromatin organization independent of the Ifng-as1 transcription or lncRNA sequence (Figure 1).
Human IFNG-AS1 expression was significantly higher in CD4+ Th1 cells, the antigen-specific memory precursor, and the central memory CD8+ T than in the effector memory T cells in LCMV-infected mice. Similarly, the IFNG–AS1 expression is maintained long term (up to a decade) at high levels in human memory T cells [180] and abundantly expressed in activated Natural Killer (NK) cells [181]. These findings highlight the differences in IFNG–AS1 expressions in various immune cell phenotypes and indicate its functional relevance in acute and memory responses to viral infections. From a clinical perspective, polymorphisms in the human IFNG–AS1 gene have been associated with autoimmune and inflammatory disorders [182,183,184], further underscoring their role in immunity and chronic inflammatory diseases in humans.
An IFN-stimulated nuclear lncRNA, NRIR (a negative regulator of interferon response; a.k.a. lncRNA-CMPK2), produces stimulation-specific contrasting phenotypes. Specifically, NRIR inhibited the transcription of several ISGs and enhanced HCV replication in hepatocytes [65]. Moreover, NRIR inhibited the expression of IFITM3, a well-characterized ISG, in endothelial and epithelial cells during Hantaan virus infection [185]. However, in monocytes, NRIR silencing significantly reduced the LPS-induced expression of ISGs, including MX1, IFITM3, ISG15, and chemokines such as CXCL10. Although these contrasting findings strengthen the role of NRIR as a regulator of IFN responses, they highlight the cell-type and stimulus-specific functions of lncRNAs [186].
The lncRNA NEAT1 sequesters both protein and RNA in the nuclear bodies. An infection with IAV and HSV induces the expression of lncRNA NEAT1, which sequesters splicing factor proline glutamine-rich (SFPQ/PSF) to the paraspeckles. SFPQ/PSF acts as a repressor of IL-8 and HSV viral genes [187], but at the same time, it is essential for IAV mRNA polyadenylation [128]. NEAT1 activates the antiviral gene IL-8 transcription by sequestering SFPQ/PSF [187]. However, NEAT1 also recruits STAT3 to viral gene promoters and upregulates the viral replication in HSV-1 infections [74]. Thus, NEAT1 functions as an antiviral factor by inducing cytokine response but is simultaneously hijacked by HSV to facilitate viral gene expression. Similar to HSV-1 infection, NEAT1 upregulation by Hantaan virus (HTNV) infection promotes RIG-I and DDX60 transcription by relocating SFPQ from the promoters of both genes to paraspeckles [56]. Since RIG-I and DDX-60 expression are essential for interferon γ (IFN- γ) production [188,189], it appears that the induction of NEAT1 enhances antiviral responses against HTNV. During the HIV-1 replication cycle, NEAT1 sequesters unspliced HIV transcripts in nuclear paraspeckle bodies, thus, preventing the nuclear export of HIV mRNA and promoting the long-term persistence of HIV [51].
Viral infections can also hijack lncRNA functions by manipulating their RBP partners, which are critical for lncRNA activity. For example, HIV integration induces double-stranded breaks (DSB) that initiate the apoptosis pathway in the infected CD4+ T cells. Unlike CD4+ T cells, HIV-infected macrophages have been reported to evade DSB-induced apoptosis by accelerating the decay of lincRNA-p21, a lncRNA that inhibits the transcription of pro-survival genes induced during the canonical DNA damage pathway. Two protein binding partners, namely HuR and hnRNP-K, are critical for the stability and function of lincRNA-p21. HIV infection of macrophages results in the sequestration of HuR and hnRNP-K in the cytoplasm, where it increases lincRNA-p21 decay and reduces lincRNA-p21 levels in the cells. Decreased availability of hnRNP-K in the nucleus reduces the functional nuclear lncRNA-p21/hnRNP-K complex required to suppress pro-survival genes. Thus, by sequestering the proteins essential for maintaining lincRNA-p21 stability and function, HIV inhibits DSB-induced cell death and promotes its persistence in infected macrophages [84]. The expression of lncRNA SAF is significantly upregulated in HIV-1-infected human monocyte-derived macrophages (MDM) and HIV-1-infected airway macrophages obtained by the bronchoalveolar lavage of HIV-1-infected individuals. The downregulation of SAF increases caspase-3/7 activity levels in virus-infected MDMs, thus inducing apoptosis. Although the mechanisms of SAF-mediated regulation of caspase3/7 activity are not completely understood, it is proposed to be a potential target to cause cell death in HIV-infected macrophages and reduce overall HIV burden [190].
Viral lncRNAs have also been shown to manipulate large cellular gene networks through diverse mechanisms. The Epstein–Barr virus (EBV), a tumor-causing virus, is associated with various human cancers [191,192] and encodes noncoding RNAs termed BamHI A rightward transcripts (BARTs) expressed at high levels in EBV-associated epithelial tumors [93]. BARTs include microRNAs and lncRNAs [93,192,193]. An alternative splicing of BARTs results in multiple spliced forms of BART lncRNA, with putative open reading frames in BARF0, RK-BARF0, RPMS1, and A73, none of which encode proteins [194,195]. BART lncRNAs regulate EBV lytic replication [196] and an extensive cellular gene network that influences adhesion, oxidoreductase activity, inflammation, and metastasis [91]. BART lncRNAs regulate the expression of tumor suppressor gene RASA unfolded protein response (UPR) genes and may contribute to host DNA methylation [91]. High levels of CpG island methylation leading to host gene silencing is associated with EBV-positive gastric carcinomas [197]. BART lncRNA significantly inhibits mitochondrial antiviral signaling (MAVS)-induced IFNB1 promoter activity [92]. Verhoeven et al. [92] observed that BART lncRNA RMS1 is associated with RNA Polymerase II (Pol II) and the CREB-binding protein (CBP/p300) complex in the nucleus. CBP activates transcription by recruiting the transcriptional machinery and also functions as a histone acetyltransferase (HAT) to alter chromatin structure. Thus, BART lncRNAs may mediate epigenetic regulation of gene expression through an interaction with the chromatin remodeling complex. They further showed that BART lncRNA RMS1 stalled Pol II at the promoter region of IFNB1 and inhibited its transcription [92]. Verhoeven et al. showed that BART lncRNA RMS1 expression could inhibit MAVS-induced HAT activity, further indicating that BART lncRNA may regulate chromatin remodeling during the gene transcription process in viral infection [92]. An overexpression of BART lncRNA RMS1 also upregulates the transcription of IKZF3 mRNA, which encodes Aiolos protein normally expressed only in lymphoid cells. Aiolos is expressed in solid and liquid tumors at high levels, promoting tumor cell survival and metastasis [92]. Thus, BART lncRNA contributes to viral oncogenesis through multiple mechanisms. In addition, EBV-encoded miRNAs regulated host protein-coding as well as lncRNAs in the EBV-infected cells (reviewed in [198]).
Similarly, human cytomegalovirus (HCMV) encodes at least four known lncRNAs (RNA1.2, RNA2.7, RNA4.9, and RNA 5.0). RNA4.9 localizes in the nuclear viral replication complex (VRC) [199], whereas RNA1.2, RNA2.7, and RNA5.0 predominantly localize to the cytoplasm [199,200]. Repressive histone modifications around the major immediate early promoter (MIEP) region inhibit HCMV lytic cycle and latency in myeloid cells [201,202,203,204]. RNA4.9 is transcribed in latently infected CD14 (+) monocytes and CD34 (+) cells, tethers the components of the PRC complex to the MIEP, enriches the repressive H3K27me3 mark at MIEP, and inhibits viral transcription [90]. RNA4.9 also mediates the formation of the RNA–DNA hybrid and the initiation of viral DNA replication in the lytic phase [199]. Though RNA2.7 localizes primarily in the cytoplasm, it has been shown to function in the nucleus and mitochondria. RNA2.7 binds Polymerase II (PolII) and blocks its interaction with phosphorylated cyclin-dependent kinase (pCDK), thus inhibiting the phosphorylation of Pol II. The inhibition of Pol II phosphorylation leads to host cell cycle arrests and facilitates viral replication [205]. RNA2.7 interacts directly with the mitochondrial complex protein GRIM-19, prevents its relocalization, and maintains high ATP production levels during lytic infection [206]. In addition, RNA2.7 is shown to have additional functions with yet unknown molecular mechanisms, such as the inhibition of apoptosis and maintenance of latency by the suppression of lytic gene expression in latent cells [207]. RNA2.7 stabilizes cellular transcripts that promote cellular motility and viral spread in lytic infection [208]. RNA1.2 is expressed at high levels during lytic infection, inhibits cellular NF-kB activation, and mediates the extracellular release of IL-6 [209].
Herpes simplex virus (HSV)-encoded Latency-associated transcript (LAT) is the only viral transcript expressed during the latent infection of neurons and plays an important role in HSV latency. LAT long non-coding transcript accumulates in the nucleus and contributes to the silencing of viral lytic genes by the heterochromatization of their promoters. LAT is suspected to be involved in the recruitment of chromatin remodeling complexes during heterochomatization, although the mechanisms are unclear [210,211,212,213,214]. In addition, the LAT transcript encodes microRNA (miRNA), small RNA (sRNA), short non-coding RNA (sncRNA), and open reading frame (ORF) encoding proteins that mediate numerous functions to maintain latency. The functions of LAT-encoded miRNAs, sRNAs, sncRNAs, and ORFs have been studied extensively and reviewed previously [215,216].
Numerous studies have revealed that while some lncRNAs affect multiple viral infections, several virus-induced lncRNAs act independently or in concert to regulate a single pathway. Although our understanding of lncRNA function in viral pathogenesis has significantly improved, future studies should focus on identifying and characterizing lncRNAs that have pronounced effects on virus replication, infection outcomes, and, more importantly, their molecular mechanisms. Investigating the mechanisms of lncRNA functions has immense therapeutic potential, as lncRNAs could serve as molecular targets for future antiviral therapy.

5. Therapeutic Potential of lncRNAs

Given their proven role in cellular defense and viral pathogenesis, careful functional studies are needed to define the diagnostic or therapeutic potential of lncRNAs. While most research focuses on understanding lncRNA functional biology, recent studies also explored their potential diagnostic and therapeutic targets [217]. Most studies described in this review have used lncRNA knockdown or overexpression to decipher the functional impact of lncRNAs on the expression of their protein-coding targets, viral replication, and immune response. A diverse array of modalities, such as RNA interference using si/shRNA [218], antisense oligonucleotides (ASO) [219], CRISPR-mediated knockout (CRISPR-KO), knock-in (CRISPR-KI) [180], transcriptional activation or inhibition (CRISPRi/a), and CRISPR-mediated RNA silencing [220,221,222,223] are available to manipulate lncRNA expression for functional studies (Table 2). In addition, various strategies have been employed to augment lncRNA expression in the local genomic context. Talen-mediated knock-in of strong promoters upstream of lncRNA transcription start site [224] and transcriptional activation using CRISPR/dCas9 (CRISPRa) [225] have been used to activate lncRNA expression from their endogenous loci. CRISPR-display is a compelling strategy that allows for the site-specific delivery of lncRNA transcripts using CRISPR/dCas9 and guide the RNA sequence fused with the lncRNA sequence [226]. Small-scale lncRNA knockdown screens have yielded significant insights into the role of virus-induced lncRNAs on viral pathogenesis [77,227]. Many experimental methods are used to probe lncRNA interactions with their protein-binding partner or chromatin and investigate lncRNA functions. However, most methods are technically challenging, need skilled researchers, specialized equipment, and reagents, and, thus, are prohibitively expensive for most laboratories. Several computational methods are being developed to predict lncRNA interactions and functions systematically. For example, lncRNA–DNA interactions that may regulate transcription could be predicted using computational methods such as Triplexator [228,229], Triplex Domain Finder [230], LongTarget [231], and Triple [232,233], which examine whether lncRNAs form triplexes with target promoters and enhancers. “Super-lncRNAs” [234] predict lncRNAs that bind super-enhancers through triplex formation. However, the functional impact of lncRNAs on virus infection must be rigorously investigated using genome-wide functional studies in cellular models and primary cells.
The thorough investigation of lncRNA in tumor biology [235] using preclinical models led to the development of lncRNA-based diagnostic [236] and therapeutic modalities for cancer, with some showing promising results in human clinical trials [237]. LncRNAs that show the potential to enforce viral latency or reactivation could be good candidates for direct therapeutic targeting. The epigenetic modulation of viral latency using latency reversal agents such as HDAC inhibitors can affect a wide gene network and show off-target effects. Recruiting or deterring an lncRNA to or from a virus promoter could provide specificity in altering viral latency. An interaction with a specific RNA-binding protein partner is essential to lncRNA function; accordingly, small molecule therapeutics may be used directly to disrupt the critical interactions [84,238]. Computational tools are being researched and developed to use nucleotide sequences and structural motifs of lncRNAs to predict their subcellular localization [239,240,241,242], interactions with RNA-binding proteins, and functions. However, successful therapeutic targeting of lncRNAs will depend on our ability to precisely identify relevant RNA motif/s and better understand the structural and functional features of lncRNAs.

6. Conclusions

In recent years, we have seen an avalanche of new information on lncRNA expression in virus-infected cells and the identification of several lncRNAs affecting viral replication and host immune responses, all of which have improved our understanding of the diverse functional potential of lncRNAs. This is a burgeoning area of research beyond mammalian species [243]. Some aspects of lncRNAs have not yet been studied in infectious disease research. For example, cellular stress, such as DNA damage [244,245], rapamycin treatment [246], and cellular differentiation [247], regulate the subcellular localization of lncRNAs. A recent study showed that an influenza virus-induced murine lncRNA, Lnc45, resides mainly in the nucleus in uninfected cells but translocates to the cytoplasm in H5N1-infected cells and dramatically impedes viral replication [248]. This observation highlights the need for a systematic study of the cellular redistribution of lncRNAs in viral infections. Some lncRNAs are highly conserved across species, but most show lower sequence conservation than protein-coding genes [249,250,251]. Therefore, it is necessary to examine the sequence and functions of relevant lncRNAs in higher animal models that are closer to humans like non-human primate (NHP) species and evaluate the use of NHP models in preclinical studies for the therapeutic targeting of lncRNAs. The systemic or specific delivery of lncRNAs or lncRNA-inhibiting RNA-based therapeutics is currently under investigation for the clinical management of several human diseases [252]. Optimal delivery modalities to target specific cells and tissues are being intensely studied. An in-depth analysis of lncRNAs in viral infections, particularly, those establishing latent reservoirs, such as HIV infection, has enormous potential for discovering novel regulatory mechanisms associated with immune response/inflammation, viral replication, and long-term viral persistence. These studies can potentially lead to identifying novel and highly selective therapeutic targets.

Author Contributions

V.K., M.M. and S.K. conceptualized the framework, and all authors (V.K., M.M., S.J., S.K.) contributed to the writing and editing of the review. All authors have read and agreed to the published version of the manuscript.

Funding

V.K., S.J., M.M. and S.K. are supported by the Texas Biomedical Research Institute, Texas Biomed Forum award (2019) to V.K., and San Antonio precision partnerships award to S.K. Research reported in this publication was supported in parts by the National Institute Of Allergy and Infectious Diseases of the National Institutes of Health under Award Number R56AI150371 (S.K.), R01AI157850 (S.K.), R21 AI140956 (S.K.) and R01DA042524 and R01DA052845 to MM. The content is solely the responsibility of the authors and does not necessarily represent the official views of the National Institutes of Health.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Brass, A.L.; Dykxhoorn, D.M.; Benita, Y.; Yan, N.; Engelman, A.; Xavier, R.J.; Lieberman, J.; Elledge, S.J. Identification of host proteins required for HIV infection through a functional genomic screen. Science 2008, 319, 921–926. [Google Scholar] [CrossRef]
  2. Brass, A.L.; Huang, I.C.; Benita, Y.; John, S.P.; Krishnan, M.N.; Feeley, E.M.; Ryan, B.J.; Weyer, J.L.; van der Weyden, L.; Fikrig, E.; et al. The IFITM proteins mediate cellular resistance to influenza A H1N1 virus, West Nile virus, and dengue virus. Cell 2009, 139, 1243–1254. [Google Scholar] [CrossRef] [Green Version]
  3. Cherry, S.; Doukas, T.; Armknecht, S.; Whelan, S.; Wang, H.; Sarnow, P.; Perrimon, N. Genome-wide RNAi screen reveals a specific sensitivity of IRES-containing RNA viruses to host translation inhibition. Genes. Dev. 2005, 19, 445–452. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Deffrasnes, C.; Marsh, G.A.; Foo, C.H.; Rootes, C.L.; Gould, C.M.; Grusovin, J.; Monaghan, P.; Lo, M.K.; Tompkins, S.M.; Adams, T.E.; et al. Genome-wide siRNA Screening at Biosafety Level 4 Reveals a Crucial Role for Fibrillarin in Henipavirus Infection. PLoS Pathog. 2016, 12, e1005478. [Google Scholar] [CrossRef] [Green Version]
  5. Karlas, A.; Machuy, N.; Shin, Y.; Pleissner, K.P.; Artarini, A.; Heuer, D.; Becker, D.; Khalil, H.; Ogilvie, L.A.; Hess, S.; et al. Genome-wide RNAi screen identifies human host factors crucial for influenza virus replication. Nature 2010, 463, 818–822. [Google Scholar] [CrossRef]
  6. Konig, R.; Zhou, Y.; Elleder, D.; Diamond, T.L.; Bonamy, G.M.; Irelan, J.T.; Chiang, C.Y.; Tu, B.P.; De Jesus, P.D.; Lilley, C.E.; et al. Global analysis of host-pathogen interactions that regulate early-stage HIV-1 replication. Cell 2008, 135, 49–60. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Lipovsky, A.; Popa, A.; Pimienta, G.; Wyler, M.; Bhan, A.; Kuruvilla, L.; Guie, M.A.; Poffenberger, A.C.; Nelson, C.D.; Atwood, W.J.; et al. Genome-wide siRNA screen identifies the retromer as a cellular entry factor for human papillomavirus. Proc. Natl. Acad. Sci. USA 2013, 110, 7452–7457. [Google Scholar] [CrossRef] [Green Version]
  8. Martin, S.; Chiramel, A.I.; Schmidt, M.L.; Chen, Y.C.; Whitt, N.; Watt, A.; Dunham, E.C.; Shifflett, K.; Traeger, S.; Leske, A.; et al. A genome-wide siRNA screen identifies a druggable host pathway essential for the Ebola virus life cycle. Genome Med. 2018, 10, 58. [Google Scholar] [CrossRef]
  9. Ooi, Y.S.; Stiles, K.M.; Liu, C.Y.; Taylor, G.M.; Kielian, M. Genome-wide RNAi screen identifies novel host proteins required for alphavirus entry. PLoS Pathog. 2013, 9, e1003835. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  10. Sivan, G.; Martin, S.E.; Myers, T.G.; Buehler, E.; Szymczyk, K.H.; Ormanoglu, P.; Moss, B. Human genome-wide RNAi screen reveals a role for nuclear pore proteins in poxvirus morphogenesis. Proc. Natl. Acad. Sci. USA 2013, 110, 3519–3524. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  11. Wu, K.X.; Phuektes, P.; Kumar, P.; Goh, G.Y.; Moreau, D.; Chow, V.T.; Bard, F.; Chu, J.J. Human genome-wide RNAi screen reveals host factors required for enterovirus 71 replication. Nat. Commun. 2016, 7, 13150. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Yasunaga, A.; Hanna, S.L.; Li, J.; Cho, H.; Rose, P.P.; Spiridigliozzi, A.; Gold, B.; Diamond, M.S.; Cherry, S. Genome-wide RNAi screen identifies broadly-acting host factors that inhibit arbovirus infection. PLoS Pathog. 2014, 10, e1003914. [Google Scholar] [CrossRef] [Green Version]
  13. Yeung, M.L.; Houzet, L.; Yedavalli, V.S.; Jeang, K.T. A genome-wide short hairpin RNA screening of jurkat T-cells for human proteins contributing to productive HIV-1 replication. J. Biol. Chem. 2009, 284, 19463–19473. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Zhou, H.; Xu, M.; Huang, Q.; Gates, A.T.; Zhang, X.D.; Castle, J.C.; Stec, E.; Ferrer, M.; Strulovici, B.; Hazuda, D.J.; et al. Genome-scale RNAi screen for host factors required for HIV replication. Cell Host Microbe 2008, 4, 495–504. [Google Scholar] [CrossRef] [Green Version]
  15. Baggen, J.; Persoons, L.; Vanstreels, E.; Jansen, S.; Van Looveren, D.; Boeckx, B.; Geudens, V.; De Man, J.; Jochmans, D.; Wauters, J.; et al. Genome-wide CRISPR screening identifies TMEM106B as a proviral host factor for SARS-CoV-2. Nat. Genet. 2021, 53, 435–444. [Google Scholar] [CrossRef]
  16. Krasnopolsky, S.; Kuzmina, A.; Taube, R. Genome-wide CRISPR knockout screen identifies ZNF304 as a silencer of HIV transcription that promotes viral latency. PLoS Pathog. 2020, 16, e1008834. [Google Scholar] [CrossRef]
  17. Kulsuptrakul, J.; Wang, R.; Meyers, N.L.; Ott, M.; Puschnik, A.S. A genome-wide CRISPR screen identifies UFMylation and TRAMP-like complexes as host factors required for hepatitis A virus infection. Cell Rep. 2021, 34, 108859. [Google Scholar] [CrossRef]
  18. Li, B.; Clohisey, S.M.; Chia, B.S.; Wang, B.; Cui, A.; Eisenhaure, T.; Schweitzer, L.D.; Hoover, P.; Parkinson, N.J.; Nachshon, A.; et al. Genome-wide CRISPR screen identifies host dependency factors for influenza A virus infection. Nat. Commun. 2020, 11, 164. [Google Scholar] [CrossRef] [Green Version]
  19. Li, Y.; Muffat, J.; Omer Javed, A.; Keys, H.R.; Lungjangwa, T.; Bosch, I.; Khan, M.; Virgilio, M.C.; Gehrke, L.; Sabatini, D.M.; et al. Genome-wide CRISPR screen for Zika virus resistance in human neural cells. Proc. Natl. Acad. Sci. USA 2019, 116, 9527–9532. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  20. Park, R.J.; Wang, T.; Koundakjian, D.; Hultquist, J.F.; Lamothe-Molina, P.; Monel, B.; Schumann, K.; Yu, H.; Krupzcak, K.M.; Garcia-Beltran, W.; et al. A genome-wide CRISPR screen identifies a restricted set of HIV host dependency factors. Nat. Genet. 2017, 49, 193–203. [Google Scholar] [CrossRef]
  21. Thamamongood, T.; Aebischer, A.; Wagner, V.; Chang, M.W.; Elling, R.; Benner, C.; Garcia-Sastre, A.; Kochs, G.; Beer, M.; Schwemmle, M. A Genome-Wide CRISPR-Cas9 Screen Reveals the Requirement of Host Cell Sulfation for Schmallenberg Virus Infection. J. Virol. 2020, 94, e00752-20. [Google Scholar] [CrossRef] [PubMed]
  22. Zhu, Y.; Feng, F.; Hu, G.; Wang, Y.; Yu, Y.; Zhu, Y.; Xu, W.; Cai, X.; Sun, Z.; Han, W.; et al. A genome-wide CRISPR screen identifies host factors that regulate SARS-CoV-2 entry. Nat. Commun. 2021, 12, 961. [Google Scholar] [CrossRef] [PubMed]
  23. Consortium, E.P.; Bernstein, B.E.; Birney, E.; Dunham, I.; Green, E.D.; Gunter, C.; Snyder, M. An integrated encyclopedia of DNA elements in the human genome. Nature 2012, 489, 57–74. [Google Scholar] [CrossRef] [Green Version]
  24. Iyer, M.K.; Niknafs, Y.S.; Malik, R.; Singhal, U.; Sahu, A.; Hosono, Y.; Barrette, T.R.; Prensner, J.R.; Evans, J.R.; Zhao, S.; et al. The landscape of long noncoding RNAs in the human transcriptome. Nat. Genet. 2015, 47, 199–208. [Google Scholar] [CrossRef] [PubMed]
  25. Nagano, T.; Fraser, P. No-nonsense functions for long noncoding RNAs. Cell 2011, 145, 178–181. [Google Scholar] [CrossRef] [Green Version]
  26. Wilusz, J.E.; Sunwoo, H.; Spector, D.L. Long noncoding RNAs: Functional surprises from the RNA world. Genes. Dev. 2009, 23, 1494–1504. [Google Scholar] [CrossRef] [Green Version]
  27. Chen, Y.G.; Satpathy, A.T.; Chang, H.Y. Gene regulation in the immune system by long noncoding RNAs. Nat. Immunol. 2017, 18, 962–972. [Google Scholar] [CrossRef]
  28. Wapinski, O.; Chang, H.Y. Long noncoding RNAs and human disease. Trends Cell Biol. 2011, 21, 354–361. [Google Scholar] [CrossRef]
  29. Winterling, C.; Koch, M.; Koeppel, M.; Garcia-Alcalde, F.; Karlas, A.; Meyer, T.F. Evidence for a crucial role of a host non-coding RNA in influenza A virus replication. RNA Biol. 2014, 11, 66–75. [Google Scholar] [CrossRef] [PubMed]
  30. Sun, S.; Yao, M.; Yuan, L.; Qiao, J. Long-chain non-coding RNA n337374 relieves symptoms of respiratory syncytial virus-induced asthma by inhibiting dendritic cell maturation via the CD86 and the ERK pathway. Allergol. Immunopathol. (Madr) 2021, 49, 100–107. [Google Scholar] [CrossRef] [PubMed]
  31. Carnero, E.; Barriocanal, M.; Prior, C.; Pablo Unfried, J.; Segura, V.; Guruceaga, E.; Enguita, M.; Smerdou, C.; Gastaminza, P.; Fortes, P. Long noncoding RNA EGOT negatively affects the antiviral response and favors HCV replication. EMBO Rep. 2016, 17, 1013–1028. [Google Scholar] [CrossRef] [Green Version]
  32. Zhao, H.; Chen, M.; Lind, S.B.; Pettersson, U. Distinct temporal changes in host cell lncRNA expression during the course of an adenovirus infection. Virology 2016, 492, 242–250. [Google Scholar] [CrossRef]
  33. Liu, H.; Xu, J.; Yang, Y.; Wang, X.; Wu, E.; Majerciak, V.; Zhang, T.; Steenbergen, R.D.M.; Wang, H.K.; Banerjee, N.S.; et al. Oncogenic HPV promotes the expression of the long noncoding RNA lnc-FANCI-2 through E7 and YY1. Proc. Natl. Acad. Sci. USA 2021, 118, e2014195118. [Google Scholar] [CrossRef]
  34. Kuo, R.L.; Chen, Y.T.; Li, H.A.; Wu, C.C.; Chiang, H.C.; Lin, J.Y.; Huang, H.I.; Shih, S.R.; Chin-Ming Tan, B. Molecular determinants and heterogeneity underlying host response to EV-A71 infection at single-cell resolution. RNA Biol. 2021, 18, 796–808. [Google Scholar] [CrossRef]
  35. Devadoss, D.; Acharya, A.; Manevski, M.; Pandey, K.; Borchert, G.M.; Nair, M.; Mirsaeidi, M.; Byrareddy, S.N.; Chand, H.S. Distinct Mucoinflammatory Phenotype and the Immunomodulatory Long Noncoding Transcripts Associated with SARS-CoV-2 Airway Infection. medRxiv 2021. [Google Scholar] [CrossRef]
  36. Laha, S.; Saha, C.; Dutta, S.; Basu, M.; Chatterjee, R.; Ghosh, S.; Bhattacharyya, N.P. In silico analysis of altered expression of long non-coding RNA in SARS-CoV-2 infected cells and their possible regulation by STAT1, STAT3 and interferon regulatory factors. Heliyon 2021, 7, e06395. [Google Scholar] [CrossRef] [PubMed]
  37. Morenikeji, O.B.; Bernard, K.; Strutton, E.; Wallace, M.; Thomas, B.N. Evolutionarily Conserved Long Non-coding RNA Regulates Gene Expression in Cytokine Storm During COVID-19. Front. Bioeng. Biotechnol. 2020, 8, 582953. [Google Scholar] [CrossRef]
  38. Mukherjee, S.; Banerjee, B.; Karasik, D.; Frenkel-Morgenstern, M. mRNA-lncRNA Co-Expression Network Analysis Reveals the Role of lncRNAs in Immune Dysfunction during Severe SARS-CoV-2 Infection. Viruses 2021, 13, 402. [Google Scholar] [CrossRef]
  39. Natarelli, L.; Parca, L.; Mazza, T.; Weber, C.; Virgili, F.; Fratantonio, D. MicroRNAs and Long Non-Coding RNAs as Potential Candidates to Target Specific Motifs of SARS-CoV-2. Noncoding RNA 2021, 7, 14. [Google Scholar] [CrossRef] [PubMed]
  40. Shaath, H.; Alajez, N.M. Identification of PBMC-based molecular signature associational with COVID-19 disease severity. Heliyon 2021, 7, e06866. [Google Scholar] [CrossRef]
  41. Tang, H.; Gao, Y.; Li, Z.; Miao, Y.; Huang, Z.; Liu, X.; Xie, L.; Li, H.; Wen, W.; Zheng, Y.; et al. The noncoding and coding transcriptional landscape of the peripheral immune response in patients with COVID-19. Clin. Transl. Med. 2020, 10, e200. [Google Scholar] [CrossRef]
  42. Turjya, R.R.; Khan, M.A.; Mir Md Khademul Islam, A.B. Perversely expressed long noncoding RNAs can alter host response and viral proliferation in SARS-CoV-2 infection. Future Virol. 2020, 15, 577–593. [Google Scholar] [CrossRef]
  43. Vishnubalaji, R.; Shaath, H.; Alajez, N.M. Protein Coding and Long Noncoding RNA (lncRNA) Transcriptional Landscape in SARS-CoV-2 Infected Bronchial Epithelial Cells Highlight a Role for Interferon and Inflammatory Response. Genes. 2020, 11, 760. [Google Scholar] [CrossRef]
  44. Wu, Y.; Zhao, T.; Deng, R.; Xia, X.; Li, B.; Wang, X. A study of differential circRNA and lncRNA expressions in COVID-19-infected peripheral blood. Sci. Rep. 2021, 11, 7991. [Google Scholar] [CrossRef]
  45. Xiong, Y.; Liu, Y.; Cao, L.; Wang, D.; Guo, M.; Jiang, A.; Guo, D.; Hu, W.; Yang, J.; Tang, Z.; et al. Transcriptomic characteristics of bronchoalveolar lavage fluid and peripheral blood mononuclear cells in COVID-19 patients. Emerg. Microbes Infect. 2020, 9, 761–770. [Google Scholar] [CrossRef] [PubMed]
  46. Chang, S.T.; Sova, P.; Peng, X.; Weiss, J.; Law, G.L.; Palermo, R.E.; Katze, M.G. Next-generation sequencing reveals HIV-1-mediated suppression of T cell activation and RNA processing and regulation of noncoding RNA expression in a CD4+ T cell line. MBio 2011, 2, e00134-11. [Google Scholar] [CrossRef] [Green Version]
  47. Peng, X.; Sova, P.; Green, R.R.; Thomas, M.J.; Korth, M.J.; Proll, S.; Xu, J.; Cheng, Y.; Yi, K.; Chen, L.; et al. Deep sequencing of HIV-infected cells: Insights into nascent transcription and host-directed therapy. J. Virol. 2014, 88, 8768–8782. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  48. Schynkel, T.; Szaniawski, M.A.; Spivak, A.M.; Bosque, A.; Planelles, V.; Vandekerckhove, L.; Trypsteen, W. Interferon-Mediated Long Non-Coding RNA Response in Macrophages in the Context of HIV. Int. J. Mol. Sci. 2020, 21, 7741. [Google Scholar] [CrossRef] [PubMed]
  49. Trypsteen, W.; Mohammadi, P.; Van Hecke, C.; Mestdagh, P.; Lefever, S.; Saeys, Y.; De Bleser, P.; Vandesompele, J.; Ciuffi, A.; Vandekerckhove, L.; et al. Differential expression of lncRNAs during the HIV replication cycle: An underestimated layer in the HIV-host interplay. Sci. Rep. 2016, 6, 36111. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  50. Trypsteen, W.; White, C.H.; Mukim, A.; Spina, C.A.; De Spiegelaere, W.; Lefever, S.; Planelles, V.; Bosque, A.; Woelk, C.H.; Vandekerckhove, L.; et al. Long non-coding RNAs and latent HIV—A search for novel targets for latency reversal. PLoS ONE 2019, 14, e0224879. [Google Scholar] [CrossRef] [PubMed]
  51. Zhang, Q.; Chen, C.Y.; Yedavalli, V.S.; Jeang, K.T. NEAT1 long noncoding RNA and paraspeckle bodies modulate HIV-1 posttranscriptional expression. MBio 2013, 4, e00596-12. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Ouyang, J.; Hu, J.; Chen, J.L. lncRNAs regulate the innate immune response to viral infection. Wiley Interdiscip. Rev. RNA 2016, 7, 129–143. [Google Scholar] [CrossRef]
  53. Fan, J.; Cheng, M.; Chi, X.; Liu, X.; Yang, W. A Human Long Non-coding RNA LncATV Promotes Virus Replication Through Restricting RIG-I-Mediated Innate Immunity. Front. Immunol. 2019, 10, 1711. [Google Scholar] [CrossRef] [Green Version]
  54. Jiang, M.; Zhang, S.; Yang, Z.; Lin, H.; Zhu, J.; Liu, L.; Wang, W.; Liu, S.; Liu, W.; Ma, Y.; et al. Self-Recognition of an Inducible Host lncRNA by RIG-I Feedback Restricts Innate Immune Response. Cell 2018, 173, 906–919 e913. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Lin, H.; Jiang, M.; Liu, L.; Yang, Z.; Ma, Z.; Liu, S.; Ma, Y.; Zhang, L.; Cao, X. The long noncoding RNA Lnczc3h7a promotes a TRIM25-mediated RIG-I antiviral innate immune response. Nat. Immunol. 2019, 20, 812–823. [Google Scholar] [CrossRef] [PubMed]
  56. Ma, H.; Han, P.; Ye, W.; Chen, H.; Zheng, X.; Cheng, L.; Zhang, L.; Yu, L.; Wu, X.; Xu, Z.; et al. The Long Noncoding RNA NEAT1 Exerts Antihantaviral Effects by Acting as Positive Feedback for RIG-I Signaling. J. Virol. 2017, 91, e02250-16. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Ouyang, J.; Zhu, X.; Chen, Y.; Wei, H.; Chen, Q.; Chi, X.; Qi, B.; Zhang, L.; Zhao, Y.; Gao, G.F.; et al. NRAV, a long noncoding RNA, modulates antiviral responses through suppression of interferon-stimulated gene transcription. Cell Host Microbe 2014, 16, 616–626. [Google Scholar] [CrossRef] [Green Version]
  58. Xie, Q.; Chen, S.; Tian, R.; Huang, X.; Deng, R.; Xue, B.; Qin, Y.; Xu, Y.; Wang, J.; Guo, M.; et al. Long Noncoding RNA ITPRIP-1 Positively Regulates the Innate Immune Response through Promotion of Oligomerization and Activation of MDA5. J. Virol. 2018, 92, e00507-18. [Google Scholar] [CrossRef] [Green Version]
  59. Gonzalez-Moro, I.; Olazagoitia-Garmendia, A.; Colli, M.L.; Cobo-Vuilleumier, N.; Postler, T.S.; Marselli, L.; Marchetti, P.; Ghosh, S.; Gauthier, B.R.; Eizirik, D.L.; et al. The T1D-associated lncRNA Lnc13 modulates human pancreatic beta cell inflammation by allele-specific stabilization of STAT1 mRNA. Proc. Natl. Acad. Sci. USA 2020, 117, 9022–9031. [Google Scholar] [CrossRef]
  60. Hu, P.; Wilhelm, J.; Gerresheim, G.K.; Shalamova, L.A.; Niepmann, M. Lnc-ITM2C-1 and GPR55 Are Proviral Host Factors for Hepatitis C Virus. Viruses 2019, 11, 549. [Google Scholar] [CrossRef] [Green Version]
  61. Li, X.; Guo, G.; Lu, M.; Chai, W.; Li, Y.; Tong, X.; Li, J.; Jia, X.; Liu, W.; Qi, D.; et al. Long Noncoding RNA Lnc-MxA Inhibits Beta Interferon Transcription by Forming RNA-DNA Triplexes at Its Promoter. J. Virol. 2019, 93. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Wang, Q.; Zhang, D.; Feng, W.; Guo, Y.; Sun, X.; Zhang, M.; Guan, Z.; Duan, M. Long noncoding RNA TSPOAP1 antisense RNA 1 negatively modulates type I IFN signaling to facilitate influenza A virus replication. J. Med. Virol. 2019, 94, 557–566. [Google Scholar] [CrossRef] [PubMed]
  63. Zhao, L.; Xia, M.; Wang, K.; Lai, C.; Fan, H.; Gu, H.; Yang, P.; Wang, X. A Long Non-coding RNA IVRPIE Promotes Host Antiviral Immune Responses Through Regulating Interferon beta1 and ISG Expression. Front. Microbiol. 2020, 11, 260. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Zhang, Q.; Chao, T.C.; Patil, V.S.; Qin, Y.; Tiwari, S.K.; Chiou, J.; Dobin, A.; Tsai, C.M.; Li, Z.; Dang, J.; et al. The long noncoding RNA ROCKI regulates inflammatory gene expression. EMBO J. 2019, 38, e100041. [Google Scholar] [CrossRef]
  65. Kambara, H.; Niazi, F.; Kostadinova, L.; Moonka, D.K.; Siegel, C.T.; Post, A.B.; Carnero, E.; Barriocanal, M.; Fortes, P.; Anthony, D.D.; et al. Negative regulation of the interferon response by an interferon-induced long non-coding RNA. Nucleic Acids Res. 2014, 42, 10668–10680. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Rapicavoli, N.A.; Qu, K.; Zhang, J.; Mikhail, M.; Laberge, R.M.; Chang, H.Y. A mammalian pseudogene lncRNA at the interface of inflammation and anti-inflammatory therapeutics. Elife 2013, 2, e00762. [Google Scholar] [CrossRef]
  67. Sonkoly, E.; Bata-Csorgo, Z.; Pivarcsi, A.; Polyanka, H.; Kenderessy-Szabo, A.; Molnar, G.; Szentpali, K.; Bari, L.; Megyeri, K.; Mandi, Y.; et al. Identification and characterization of a novel, psoriasis susceptibility-related noncoding RNA gene, PRINS. J. Biol. Chem. 2005, 280, 24159–24167. [Google Scholar] [CrossRef] [Green Version]
  68. Yang, W.S.; Yeh, W.W.; Campbell, M.; Chang, L.; Chang, P.C. Long non-coding RNA KIKAT/LINC01061 as a novel epigenetic regulator that relocates KDM4A on chromatin and modulates viral reactivation. PLoS Pathog. 2021, 17, e1009670. [Google Scholar] [CrossRef] [PubMed]
  69. Xu, J.; Wang, P.; Li, Z.; Li, Z.; Han, D.; Wen, M.; Zhao, Q.; Zhang, L.; Ma, Y.; Liu, W.; et al. IRF3-binding lncRNA-ISIR strengthens interferon production in viral infection and autoinflammation. Cell Rep. 2021, 37, 109926. [Google Scholar] [CrossRef]
  70. Qiu, L.; Wang, T.; Tang, Q.; Li, G.; Wu, P.; Chen, K. Long Non-coding RNAs: Regulators of Viral Infection and the Interferon Antiviral Response. Front. Microbiol. 2018, 9, 1621. [Google Scholar] [CrossRef]
  71. Xiong, Y.; Yuan, J.; Zhang, C.; Zhu, Y.; Kuang, X.; Lan, L.; Wang, X. The STAT3-regulated long non-coding RNA Lethe promote the HCV replication. Biomed. Pharmacother. 2015, 72, 165–171. [Google Scholar] [CrossRef] [PubMed]
  72. Liu, X.; Duan, X.; Holmes, J.A.; Li, W.; Lee, S.H.; Tu, Z.; Zhu, C.; Salloum, S.; Lidofsky, A.; Schaefer, E.A.; et al. A Long Noncoding RNA Regulates Hepatitis C Virus Infection Through Interferon Alpha-Inducible Protein 6. Hepatology 2019, 69, 1004–1019. [Google Scholar] [CrossRef] [PubMed]
  73. Fang, L.; Gao, Y.; Liu, X.; Bai, J.; Jiang, P.; Wang, X. Long non-coding RNA LNC_000641 regulates pseudorabies virus replication. Vet. Res. 2021, 52, 52. [Google Scholar] [CrossRef] [PubMed]
  74. Wang, Z.; Fan, P.; Zhao, Y.; Zhang, S.; Lu, J.; Xie, W.; Jiang, Y.; Lei, F.; Xu, N.; Zhang, Y. NEAT1 modulates herpes simplex virus-1 replication by regulating viral gene transcription. Cell Mol. Life Sci. 2017, 74, 1117–1131. [Google Scholar] [CrossRef] [Green Version]
  75. Wang, Y.; Huo, Z.; Lin, Q.; Lin, Y.; Chen, C.; Huang, Y.; Huang, C.; Zhang, J.; He, J.; Liu, C.; et al. Positive Feedback Loop of Long Noncoding RNA OASL-IT1 and Innate Immune Response Restricts the Replication of Zika Virus in Epithelial A549 Cells. J. Innate Immun. 2021, 13, 179–193. [Google Scholar] [CrossRef]
  76. Nishitsuji, H.; Ujino, S.; Yoshio, S.; Sugiyama, M.; Mizokami, M.; Kanto, T.; Shimotohno, K. Long noncoding RNA #32 contributes to antiviral responses by controlling interferon-stimulated gene expression. Proc. Natl. Acad. Sci. USA 2016, 113, 10388–10393. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Li, J.; Chen, C.; Ma, X.; Geng, G.; Liu, B.; Zhang, Y.; Zhang, S.; Zhong, F.; Liu, C.; Yin, Y.; et al. Long noncoding RNA NRON contributes to HIV-1 latency by specifically inducing tat protein degradation. Nat. Commun. 2016, 7, 11730. [Google Scholar] [CrossRef] [Green Version]
  78. Qu, D.; Sun, W.W.; Li, L.; Ma, L.; Sun, L.; Jin, X.; Li, T.; Hou, W.; Wang, J.H. Long noncoding RNA MALAT1 releases epigenetic silencing of HIV-1 replication by displacing the polycomb repressive complex 2 from binding to the LTR promoter. Nucleic Acids Res. 2019, 47, 3013–3027. [Google Scholar] [CrossRef] [Green Version]
  79. Nguyen, V.T.; Kiss, T.; Michels, A.A.; Bensaude, O. 7SK small nuclear RNA binds to and inhibits the activity of CDK9/cyclin T complexes. Nature 2001, 414, 322–325. [Google Scholar] [CrossRef] [PubMed]
  80. Contreras, X.; Barboric, M.; Lenasi, T.; Peterlin, B.M. HMBA releases P-TEFb from HEXIM1 and 7SK snRNA via PI3K/Akt and activates HIV transcription. PLoS Pathog. 2007, 3, 1459–1469. [Google Scholar] [CrossRef] [Green Version]
  81. Budhiraja, S.; Famiglietti, M.; Bosque, A.; Planelles, V.; Rice, A.P. Cyclin T1 and CDK9 T-loop phosphorylation are downregulated during establishment of HIV-1 latency in primary resting memory CD4+ T cells. J. Virol. 2013, 87, 1211–1220. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Eilebrecht, S.; Benecke, B.J.; Benecke, A.G. Latent HIV-1 TAR Regulates 7SK-responsive P-TEFb Target Genes and Targets Cellular Immune Responses in the Absence of Tat. Genomics Proteomics Bioinformatics 2017, 15, 313–323. [Google Scholar] [CrossRef] [PubMed]
  83. Huan, C.; Li, Z.; Ning, S.; Wang, H.; Yu, X.F.; Zhang, W. Long Noncoding RNA uc002yug.2 Activates HIV-1 Latency through Regulation of mRNA Levels of Various RUNX1 Isoforms and Increased Tat Expression. J. Virol. 2018, 92, e01844-17. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Barichievy, S.; Naidoo, J.; Boulle, M.; Scholefield, J.; Parihar, S.P.; Coussens, A.K.; Brombacher, F.; Sigal, A.; Mhlanga, M.M. Viral Apoptosis Evasion via the MAPK Pathway by Use of a Host Long Noncoding RNA. Front. Cell Infect. Microbiol. 2018, 8, 263. [Google Scholar] [CrossRef]
  85. Chao, T.C.; Zhang, Q.; Li, Z.; Tiwari, S.K.; Qin, Y.; Yau, E.; Sanchez, A.; Singh, G.; Chang, K.; Kaul, M.; et al. The Long Noncoding RNA HEAL Regulates HIV-1 Replication through Epigenetic Regulation of the HIV-1 Promoter. mBio 2019, 10, e02016-19. [Google Scholar] [CrossRef] [Green Version]
  86. Saayman, S.; Ackley, A.; Turner, A.W.; Famiglietti, M.; Bosque, A.; Clemson, M.; Planelles, V.; Morris, K.V. An HIV-encoded antisense long noncoding RNA epigenetically regulates viral transcription. Mol. Ther. 2014, 22, 1164–1175. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Rossetto, C.C.; Pari, G. KSHV PAN RNA associates with demethylases UTX and JMJD3 to activate lytic replication through a physical interaction with the virus genome. PLoS Pathog. 2012, 8, e1002680. [Google Scholar] [CrossRef]
  88. Rossetto, C.C.; Tarrant-Elorza, M.; Verma, S.; Purushothaman, P.; Pari, G.S. Regulation of viral and cellular gene expression by Kaposi’s sarcoma-associated herpesvirus polyadenylated nuclear RNA. J. Virol. 2013, 87, 5540–5553. [Google Scholar] [CrossRef] [Green Version]
  89. Rossetto, C.C.; Tarrant-Elorza, M.; Verma, S.; Purushothaman, P.; Pari, G.S. Correction for Rossetto et al., Regulation of Viral and Cellular Gene Expression by Kaposi’s Sarcoma-Associated Herpesvirus Polyadenylated Nuclear RNA. J. Virol. 2016, 90, 4255. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  90. Rossetto, C.C.; Tarrant-Elorza, M.; Pari, G.S. Cis and trans acting factors involved in human cytomegalovirus experimental and natural latent infection of CD14 (+) monocytes and CD34 (+) cells. PLoS Pathog. 2013, 9, e1003366. [Google Scholar] [CrossRef] [Green Version]
  91. Marquitz, A.R.; Mathur, A.; Edwards, R.H.; Raab-Traub, N. Host Gene Expression Is Regulated by Two Types of Noncoding RNAs Transcribed from the Epstein-Barr Virus BamHI A Rightward Transcript Region. J. Virol. 2015, 89, 11256–11268. [Google Scholar] [CrossRef] [Green Version]
  92. Verhoeven, R.J.A.; Tong, S.; Mok, B.W.; Liu, J.; He, S.; Zong, J.; Chen, Y.; Tsao, S.W.; Lung, M.L.; Chen, H. Epstein-Barr Virus BART Long Non-coding RNAs Function as Epigenetic Modulators in Nasopharyngeal Carcinoma. Front. Oncol. 2019, 9, 1120. [Google Scholar] [CrossRef] [Green Version]
  93. Park, R.; Miller, G. Epstein-Barr Virus-Induced Nodules on Viral Replication Compartments Contain RNA Processing Proteins and a Viral Long Noncoding RNA. J. Virol. 2018, 92, e01254-18. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Rennekamp, A.J.; Lieberman, P.M. Initiation of Epstein-Barr virus lytic replication requires transcription and the formation of a stable RNA-DNA hybrid molecule at OriLyt. J. Virol. 2011, 85, 2837–2850. [Google Scholar] [CrossRef] [Green Version]
  95. Peng, X.; Gralinski, L.; Armour, C.D.; Ferris, M.T.; Thomas, M.J.; Proll, S.; Bradel-Tretheway, B.G.; Korth, M.J.; Castle, J.C.; Biery, M.C.; et al. Unique signatures of long noncoding RNA expression in response to virus infection and altered innate immune signaling. mBio 2010, 1, e00206-10. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Sui, B.; Chen, D.; Liu, W.; Wu, Q.; Tian, B.; Li, Y.; Hou, J.; Liu, S.; Xie, J.; Jiang, H.; et al. A novel antiviral lncRNA, EDAL, shields a T309 O-GlcNAcylation site to promote EZH2 lysosomal degradation. Genome Biol. 2020, 21, 228. [Google Scholar] [CrossRef]
  97. Fang, S.; Zhang, L.; Guo, J.; Niu, Y.; Wu, Y.; Li, H.; Zhao, L.; Li, X.; Teng, X.; Sun, X.; et al. NONCODEV5: A comprehensive annotation database for long non-coding RNAs. Nucleic Acids Res. 2018, 46, D308–D314. [Google Scholar] [CrossRef] [PubMed]
  98. Uszczynska-Ratajczak, B.; Lagarde, J.; Frankish, A.; Guigo, R.; Johnson, R. Towards a complete map of the human long non-coding RNA transcriptome. Nat. Rev. Genet. 2018, 19, 535–548. [Google Scholar] [CrossRef]
  99. Derrien, T.; Johnson, R.; Bussotti, G.; Tanzer, A.; Djebali, S.; Tilgner, H.; Guernec, G.; Martin, D.; Merkel, A.; Knowles, D.G.; et al. The GENCODE v7 catalog of human long noncoding RNAs: Analysis of their gene structure, evolution, and expression. Genome Res. 2012, 22, 1775–1789. [Google Scholar] [CrossRef] [Green Version]
  100. Griffiths-Jones, S. Annotating noncoding RNA genes. Annu. Rev. Genomics Hum. Genet. 2007, 8, 279–298. [Google Scholar] [CrossRef] [Green Version]
  101. Lagarde, J.; Uszczynska-Ratajczak, B.; Carbonell, S.; Perez-Lluch, S.; Abad, A.; Davis, C.; Gingeras, T.R.; Frankish, A.; Harrow, J.; Guigo, R.; et al. High-throughput annotation of full-length long noncoding RNAs with capture long-read sequencing. Nat. Genet. 2017, 49, 1731–1740. [Google Scholar] [CrossRef] [Green Version]
  102. Mele, M.; Mattioli, K.; Mallard, W.; Shechner, D.M.; Gerhardinger, C.; Rinn, J.L. Chromatin environment, transcriptional regulation, and splicing distinguish lincRNAs and mRNAs. Genome Res. 2017, 27, 27–37. [Google Scholar] [CrossRef] [Green Version]
  103. Schlackow, M.; Nojima, T.; Gomes, T.; Dhir, A.; Carmo-Fonseca, M.; Proudfoot, N.J. Distinctive Patterns of Transcription and RNA Processing for Human lincRNAs. Mol. Cell 2017, 65, 25–38. [Google Scholar] [CrossRef] [Green Version]
  104. West, J.A.; Davis, C.P.; Sunwoo, H.; Simon, M.D.; Sadreyev, R.I.; Wang, P.I.; Tolstorukov, M.Y.; Kingston, R.E. The long noncoding RNAs NEAT1 and MALAT1 bind active chromatin sites. Mol. Cell 2014, 55, 791–802. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Lubelsky, Y.; Ulitsky, I. Sequences enriched in Alu repeats drive nuclear localization of long RNAs in human cells. Nature 2018, 555, 107–111. [Google Scholar] [CrossRef]
  106. Zhang, B.; Gunawardane, L.; Niazi, F.; Jahanbani, F.; Chen, X.; Valadkhan, S. A novel RNA motif mediates the strict nuclear localization of a long noncoding RNA. Mol. Cell Biol. 2014, 34, 2318–2329. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  107. Miyagawa, R.; Tano, K.; Mizuno, R.; Nakamura, Y.; Ijiri, K.; Rakwal, R.; Shibato, J.; Masuo, Y.; Mayeda, A.; Hirose, T.; et al. Identification of cis- and trans-acting factors involved in the localization of MALAT-1 noncoding RNA to nuclear speckles. RNA 2012, 18, 738–751. [Google Scholar] [CrossRef] [Green Version]
  108. Hacisuleyman, E.; Goff, L.A.; Trapnell, C.; Williams, A.; Henao-Mejia, J.; Sun, L.; McClanahan, P.; Hendrickson, D.G.; Sauvageau, M.; Kelley, D.R.; et al. Topological organization of multichromosomal regions by the long intergenic noncoding RNA Firre. Nat. Struct. Mol. Biol. 2014, 21, 198–206. [Google Scholar] [CrossRef]
  109. Dumbovic, G.; Braunschweig, U.; Langner, H.K.; Smallegan, M.; Biayna, J.; Hass, E.P.; Jastrzebska, K.; Blencowe, B.; Cech, T.R.; Caruthers, M.H.; et al. Nuclear compartmentalization of TERT mRNA and TUG1 lncRNA is driven by intron retention. Nat. Commun. 2021, 12, 3308. [Google Scholar] [CrossRef] [PubMed]
  110. Gutierrez, I.V.; Dayton, J.; Harger, S.; Rossetto, C.C. The Expression and Nuclear Retention Element of Polyadenylated Nuclear RNA Is Not Required for Productive Lytic Replication of Kaposi’s Sarcoma-Associated Herpesvirus. J. Virol. 2021, 95, e0009621. [Google Scholar] [CrossRef] [PubMed]
  111. Conrad, N.K.; Mili, S.; Marshall, E.L.; Shu, M.D.; Steitz, J.A. Identification of a rapid mammalian deadenylation-dependent decay pathway and its inhibition by a viral RNA element. Mol. Cell 2006, 24, 943–953. [Google Scholar] [CrossRef] [PubMed]
  112. Conrad, N.K.; Shu, M.D.; Uyhazi, K.E.; Steitz, J.A. Mutational analysis of a viral RNA element that counteracts rapid RNA decay by interaction with the polyadenylate tail. Proc. Natl. Acad. Sci. USA 2007, 104, 10412–10417. [Google Scholar] [CrossRef] [Green Version]
  113. Conrad, N.K.; Steitz, J.A. A Kaposi’s sarcoma virus RNA element that increases the nuclear abundance of intronless transcripts. EMBO J. 2005, 24, 1831–1841. [Google Scholar] [CrossRef] [Green Version]
  114. Rossetto, C.C.; Pari, G.S. PAN’s Labyrinth: Molecular biology of Kaposi’s sarcoma-associated herpesvirus (KSHV) PAN RNA, a multifunctional long noncoding RNA. Viruses 2014, 6, 4212–4226. [Google Scholar] [CrossRef] [Green Version]
  115. Sahin, B.B.; Patel, D.; Conrad, N.K. Kaposi’s sarcoma-associated herpesvirus ORF57 protein binds and protects a nuclear noncoding RNA from cellular RNA decay pathways. PLoS Pathog. 2010, 6, e1000799. [Google Scholar] [CrossRef] [PubMed]
  116. Brown, J.A.; Valenstein, M.L.; Yario, T.A.; Tycowski, K.T.; Steitz, J.A. Formation of triple-helical structures by the 3’-end sequences of MALAT1 and MENbeta noncoding RNAs. Proc. Natl. Acad. Sci. USA 2012, 109, 19202–19207. [Google Scholar] [CrossRef] [Green Version]
  117. Hautbergue, G.M. RNA Nuclear Export: From Neurological Disorders to Cancer. Adv. Exp. Med. Biol. 2017, 1007, 89–109. [Google Scholar] [CrossRef]
  118. Guttman, M.; Amit, I.; Garber, M.; French, C.; Lin, M.F.; Feldser, D.; Huarte, M.; Zuk, O.; Carey, B.W.; Cassady, J.P.; et al. Chromatin signature reveals over a thousand highly conserved large non-coding RNAs in mammals. Nature 2009, 458, 223–227. [Google Scholar] [CrossRef]
  119. Viphakone, N.; Sudbery, I.; Griffith, L.; Heath, C.G.; Sims, D.; Wilson, S.A. Co-transcriptional Loading of RNA Export Factors Shapes the Human Transcriptome. Mol. Cell 2019, 75, 310–323 e318. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  120. Cohen, H.R.; Panning, B. XIST RNA exhibits nuclear retention and exhibits reduced association with the export factor TAP/NXF1. Chromosoma 2007, 116, 373–383. [Google Scholar] [CrossRef]
  121. Bergmann, J.H.; Spector, D.L. Long non-coding RNAs: Modulators of nuclear structure and function. Curr. Opin. Cell Biol. 2014, 26, 10–18. [Google Scholar] [CrossRef] [Green Version]
  122. Clemson, C.M.; Hutchinson, J.N.; Sara, S.A.; Ensminger, A.W.; Fox, A.H.; Chess, A.; Lawrence, J.B. An architectural role for a nuclear noncoding RNA: NEAT1 RNA is essential for the structure of paraspeckles. Mol. Cell 2009, 33, 717–726. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Engreitz, J.M.; Pandya-Jones, A.; McDonel, P.; Shishkin, A.; Sirokman, K.; Surka, C.; Kadri, S.; Xing, J.; Goren, A.; Lander, E.S.; et al. The Xist lncRNA exploits three-dimensional genome architecture to spread across the X chromosome. Science 2013, 341, 1237973. [Google Scholar] [CrossRef] [Green Version]
  124. Guttman, M.; Rinn, J.L. Modular regulatory principles of large non-coding RNAs. Nature 2012, 482, 339–346. [Google Scholar] [CrossRef] [Green Version]
  125. Mao, Y.S.; Sunwoo, H.; Zhang, B.; Spector, D.L. Direct visualization of the co-transcriptional assembly of a nuclear body by noncoding RNAs. Nat. Cell Biol. 2011, 13, 95–101. [Google Scholar] [CrossRef] [PubMed]
  126. Rinn, J.L.; Chang, H.Y. Genome regulation by long noncoding RNAs. Annu. Rev. Biochem. 2012, 81, 145–166. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  127. Wang, K.C.; Chang, H.Y. Molecular mechanisms of long noncoding RNAs. Mol. Cell 2011, 43, 904–914. [Google Scholar] [CrossRef] [Green Version]
  128. Landeras-Bueno, S.; Jorba, N.; Perez-Cidoncha, M.; Ortin, J. The splicing factor proline-glutamine rich (SFPQ/PSF) is involved in influenza virus transcription. PLoS Pathog. 2011, 7, e1002397. [Google Scholar] [CrossRef]
  129. Markaki, Y.; Gan Chong, J.; Wang, Y.; Jacobson, E.C.; Luong, C.; Tan, S.Y.X.; Jachowicz, J.W.; Strehle, M.; Maestrini, D.; Banerjee, A.K.; et al. Xist nucleates local protein gradients to propagate silencing across the X chromosome. Cell 2021, 184, 6174–6192 e6132. [Google Scholar] [CrossRef]
  130. Quinodoz, S.A.; Jachowicz, J.W.; Bhat, P.; Ollikainen, N.; Banerjee, A.K.; Goronzy, I.N.; Blanco, M.R.; Chovanec, P.; Chow, A.; Markaki, Y.; et al. RNA promotes the formation of spatial compartments in the nucleus. Cell 2021, 184, 5775–5790 e5730. [Google Scholar] [CrossRef]
  131. Quinodoz, S.A.; Ollikainen, N.; Tabak, B.; Palla, A.; Schmidt, J.M.; Detmar, E.; Lai, M.M.; Shishkin, A.A.; Bhat, P.; Takei, Y.; et al. Higher-Order Inter-chromosomal Hubs Shape 3D Genome Organization in the Nucleus. Cell 2018, 174, 744–757 e724. [Google Scholar] [CrossRef] [Green Version]
  132. Engreitz, J.M.; Haines, J.E.; Perez, E.M.; Munson, G.; Chen, J.; Kane, M.; McDonel, P.E.; Guttman, M.; Lander, E.S. Local regulation of gene expression by lncRNA promoters, transcription and splicing. Nature 2016, 539, 452–455. [Google Scholar] [CrossRef] [Green Version]
  133. Grote, P.; Wittler, L.; Hendrix, D.; Koch, F.; Wahrisch, S.; Beisaw, A.; Macura, K.; Blass, G.; Kellis, M.; Werber, M.; et al. The tissue-specific lncRNA Fendrr is an essential regulator of heart and body wall development in the mouse. Dev. Cell 2013, 24, 206–214. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Hartana, C.A.; Rassadkina, Y.; Gao, C.; Martin-Gayo, E.; Walker, B.D.; Lichterfeld, M.; Yu, X.G. Long noncoding RNA MIR4435-2HG enhances metabolic function of myeloid dendritic cells from HIV-1 elite controllers. J. Clin. Invest. 2021, 131, e146136. [Google Scholar] [CrossRef] [PubMed]
  135. Kalwa, M.; Hanzelmann, S.; Otto, S.; Kuo, C.C.; Franzen, J.; Joussen, S.; Fernandez-Rebollo, E.; Rath, B.; Koch, C.; Hofmann, A.; et al. The lncRNA HOTAIR impacts on mesenchymal stem cells via triple helix formation. Nucleic Acids Res. 2016, 44, 10631–10643. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  136. Mondal, T.; Subhash, S.; Vaid, R.; Enroth, S.; Uday, S.; Reinius, B.; Mitra, S.; Mohammed, A.; James, A.R.; Hoberg, E.; et al. MEG3 long noncoding RNA regulates the TGF-beta pathway genes through formation of RNA-DNA triplex structures. Nat. Commun. 2015, 6, 7743. [Google Scholar] [CrossRef] [Green Version]
  137. O’Leary, V.B.; Ovsepian, S.V.; Carrascosa, L.G.; Buske, F.A.; Radulovic, V.; Niyazi, M.; Moertl, S.; Trau, M.; Atkinson, M.J.; Anastasov, N. PARTICLE, a Triplex-Forming Long ncRNA, Regulates Locus-Specific Methylation in Response to Low-Dose Irradiation. Cell Rep. 2015, 11, 474–485. [Google Scholar] [CrossRef] [Green Version]
  138. O’Leary, V.B.; Smida, J.; Buske, F.A.; Carrascosa, L.G.; Azimzadeh, O.; Maugg, D.; Hain, S.; Tapio, S.; Heidenreich, W.; Kerr, J.; et al. PARTICLE triplexes cluster in the tumor suppressor WWOX and may extend throughout the human genome. Sci. Rep. 2017, 7, 7163. [Google Scholar] [CrossRef] [Green Version]
  139. Postepska-Igielska, A.; Giwojna, A.; Gasri-Plotnitsky, L.; Schmitt, N.; Dold, A.; Ginsberg, D.; Grummt, I. LncRNA Khps1 Regulates Expression of the Proto-oncogene SPHK1 via Triplex-Mediated Changes in Chromatin Structure. Mol. Cell 2015, 60, 626–636. [Google Scholar] [CrossRef]
  140. Zhao, Z.; Senturk, N.; Song, C.; Grummt, I. lncRNA PAPAS tethered to the rDNA enhancer recruits hypophosphorylated CHD4/NuRD to repress rRNA synthesis at elevated temperatures. Genes. Dev. 2018, 32, 836–848. [Google Scholar] [CrossRef] [PubMed]
  141. Arab, K.; Karaulanov, E.; Musheev, M.; Trnka, P.; Schafer, A.; Grummt, I.; Niehrs, C. GADD45A binds R-loops and recruits TET1 to CpG island promoters. Nat. Genet. 2019, 51, 217–223. [Google Scholar] [CrossRef]
  142. Schoggins, J.W.; Wilson, S.J.; Panis, M.; Murphy, M.Y.; Jones, C.T.; Bieniasz, P.; Rice, C.M. A diverse range of gene products are effectors of the type I interferon antiviral response. Nature 2011, 472, 481–485. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  143. Kaneko, S.; Bonasio, R.; Saldana-Meyer, R.; Yoshida, T.; Son, J.; Nishino, K.; Umezawa, A.; Reinberg, D. Interactions between JARID2 and noncoding RNAs regulate PRC2 recruitment to chromatin. Mol. Cell 2014, 53, 290–300. [Google Scholar] [CrossRef] [Green Version]
  144. Yang, L.; Lin, C.; Jin, C.; Yang, J.C.; Tanasa, B.; Li, W.; Merkurjev, D.; Ohgi, K.A.; Meng, D.; Zhang, J.; et al. lncRNA-dependent mechanisms of androgen-receptor-regulated gene activation programs. Nature 2013, 500, 598–602. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Yang, Y.W.; Flynn, R.A.; Chen, Y.; Qu, K.; Wan, B.; Wang, K.C.; Lei, M.; Chang, H.Y. Essential role of lncRNA binding for WDR5 maintenance of active chromatin and embryonic stem cell pluripotency. Elife 2014, 3, e02046. [Google Scholar] [CrossRef] [PubMed]
  146. Nagano, T.; Mitchell, J.A.; Sanz, L.A.; Pauler, F.M.; Ferguson-Smith, A.C.; Feil, R.; Fraser, P. The Air noncoding RNA epigenetically silences transcription by targeting G9a to chromatin. Science 2008, 322, 1717–1720. [Google Scholar] [CrossRef] [Green Version]
  147. Yang, L.; Lin, C.; Liu, W.; Zhang, J.; Ohgi, K.A.; Grinstein, J.D.; Dorrestein, P.C.; Rosenfeld, M.G. ncRNA- and Pc2 methylation-dependent gene relocation between nuclear structures mediates gene activation programs. Cell 2011, 147, 773–788. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  148. Guttman, M.; Donaghey, J.; Carey, B.W.; Garber, M.; Grenier, J.K.; Munson, G.; Young, G.; Lucas, A.B.; Ach, R.; Bruhn, L.; et al. lincRNAs act in the circuitry controlling pluripotency and differentiation. Nature 2011, 477, 295–300. [Google Scholar] [CrossRef] [Green Version]
  149. Tsai, M.C.; Manor, O.; Wan, Y.; Mosammaparast, N.; Wang, J.K.; Lan, F.; Shi, Y.; Segal, E.; Chang, H.Y. Long noncoding RNA as modular scaffold of histone modification complexes. Science 2010, 329, 689–693. [Google Scholar] [CrossRef] [Green Version]
  150. Haller, O.; Staeheli, P.; Kochs, G. Protective role of interferon-induced Mx GTPases against influenza viruses. Rev. Sci. Tech. 2009, 28, 219–231. [Google Scholar] [CrossRef] [Green Version]
  151. Fan, Y.; Shen, B.; Tan, M.; Mu, X.; Qin, Y.; Zhang, F.; Liu, Y. TGF-beta-induced upregulation of malat1 promotes bladder cancer metastasis by associating with suz12. Clin. Cancer Res. 2014, 20, 1531–1541. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  152. Hirata, H.; Hinoda, Y.; Shahryari, V.; Deng, G.; Nakajima, K.; Tabatabai, Z.L.; Ishii, N.; Dahiya, R. Long Noncoding RNA MALAT1 Promotes Aggressive Renal Cell Carcinoma through Ezh2 and Interacts with miR-205. Cancer Res. 2015, 75, 1322–1331. [Google Scholar] [CrossRef] [Green Version]
  153. Cheng, Y.; Imanirad, P.; Jutooru, I.; Hedrick, E.; Jin, U.H.; Rodrigues Hoffman, A.; Leal de Araujo, J.; Morpurgo, B.; Golovko, A.; Safe, S. Role of metastasis-associated lung adenocarcinoma transcript-1 (MALAT-1) in pancreatic cancer. PLoS ONE 2018, 13, e0192264. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  154. Wang, X.; Sehgal, L.; Jain, N.; Khashab, T.; Mathur, R.; Samaniego, F. LncRNA MALAT1 promotes development of mantle cell lymphoma by associating with EZH2. J. Transl. Med. 2016, 14, 346. [Google Scholar] [CrossRef] [Green Version]
  155. Chang, P.C.; Fitzgerald, L.D.; Hsia, D.A.; Izumiya, Y.; Wu, C.Y.; Hsieh, W.P.; Lin, S.F.; Campbell, M.; Lam, K.S.; Luciw, P.A.; et al. Histone demethylase JMJD2A regulates Kaposi’s sarcoma-associated herpesvirus replication and is targeted by a viral transcriptional factor. J. Virol. 2011, 85, 3283–3293. [Google Scholar] [CrossRef] [Green Version]
  156. Yang, W.S.; Campbell, M.; Chang, P.C. SUMO modification of a heterochromatin histone demethylase JMJD2A enables viral gene transactivation and viral replication. PLoS Pathog. 2017, 13, e1006216. [Google Scholar] [CrossRef] [Green Version]
  157. Kobayashi-Ishihara, M.; Yamagishi, M.; Hara, T.; Matsuda, Y.; Takahashi, R.; Miyake, A.; Nakano, K.; Yamochi, T.; Ishida, T.; Watanabe, T. HIV-1-encoded antisense RNA suppresses viral replication for a prolonged period. Retrovirology 2012, 9, 38. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  158. Zapata, J.C.; Campilongo, F.; Barclay, R.A.; DeMarino, C.; Iglesias-Ussel, M.D.; Kashanchi, F.; Romerio, F. The Human Immunodeficiency Virus 1 ASP RNA promotes viral latency by recruiting the Polycomb Repressor Complex 2 and promoting nucleosome assembly. Virology 2017, 506, 34–44. [Google Scholar] [CrossRef] [PubMed]
  159. Purushothaman, P.; Dabral, P.; Gupta, N.; Sarkar, R.; Verma, S.C. KSHV Genome Replication and Maintenance. Front. Microbiol. 2016, 7, 54. [Google Scholar] [CrossRef] [Green Version]
  160. Sun, R.; Lin, S.F.; Gradoville, L.; Miller, G. Polyadenylylated nuclear RNA encoded by Kaposi sarcoma-associated herpesvirus. Proc. Natl. Acad. Sci. USA 1996, 93, 11883–11888. [Google Scholar] [CrossRef] [Green Version]
  161. Borah, S.; Darricarrere, N.; Darnell, A.; Myoung, J.; Steitz, J.A. A viral nuclear noncoding RNA binds re-localized poly(A) binding protein and is required for late KSHV gene expression. PLoS Pathog. 2011, 7, e1002300. [Google Scholar] [CrossRef]
  162. Rossetto, C.C.; Pari, G.S. Kaposi’s sarcoma-associated herpesvirus noncoding polyadenylated nuclear RNA interacts with virus- and host cell-encoded proteins and suppresses expression of genes involved in immune modulation. J. Virol. 2011, 85, 13290–13297. [Google Scholar] [CrossRef] [Green Version]
  163. Song, M.J.; Brown, H.J.; Wu, T.T.; Sun, R. Transcription activation of polyadenylated nuclear rna by rta in human herpesvirus 8/Kaposi’s sarcoma-associated herpesvirus. J. Virol. 2001, 75, 3129–3140. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  164. Campbell, M.; Kim, K.Y.; Chang, P.C.; Huerta, S.; Shevchenko, B.; Wang, D.H.; Izumiya, C.; Kung, H.J.; Izumiya, Y. A lytic viral long noncoding RNA modulates the function of a latent protein. J. Virol. 2014, 88, 1843–1848. [Google Scholar] [CrossRef] [Green Version]
  165. Kumar, A.; Lyu, Y.; Yanagihashi, Y.; Chantarasrivong, C.; Majerciak, V.; Salemi, M.; Wang, K.H.; Inagaki, T.; Chuang, F.; Davis, R.R.; et al. KSHV episome tethering sites on host chromosomes and regulation of latency-lytic switch by CHD4. Cell Rep. 2022, 39, 110788. [Google Scholar] [CrossRef]
  166. Hiura, K.; Strahan, R.; Uppal, T.; Prince, B.; Rossetto, C.C.; Verma, S.C. KSHV ORF59 and PAN RNA Recruit Histone Demethylases to the Viral Chromatin during Lytic Reactivation. Viruses 2020, 12, 420. [Google Scholar] [CrossRef] [Green Version]
  167. Agarwal, S.; Vierbuchen, T.; Ghosh, S.; Chan, J.; Jiang, Z.; Kandasamy, R.K.; Ricci, E.; Fitzgerald, K.A. The long non-coding RNA LUCAT1 is a negative feedback regulator of interferon responses in humans. Nat. Commun. 2020, 11, 6348. [Google Scholar] [CrossRef]
  168. Aznaourova, M.; Schmerer, N.; Janga, H.; Zhang, Z.; Pauck, K.; Bushe, J.; Volkers, S.M.; Wendisch, D.; Georg, P.; Ntini, E.; et al. Single-cell RNA sequencing uncovers the nuclear decoy lincRNA PIRAT as a regulator of systemic monocyte immunity during COVID-19. Proc. Natl. Acad. Sci. USA 2022, 119, e2120680119. [Google Scholar] [CrossRef]
  169. Ma, G.; Yasunaga, J.I.; Shimura, K.; Takemoto, K.; Watanabe, M.; Amano, M.; Nakata, H.; Liu, B.; Zuo, X.; Matsuoka, M. Human retroviral antisense mRNAs are retained in the nuclei of infected cells for viral persistence. Proc. Natl. Acad. Sci. USA 2021, 118, e2014783118. [Google Scholar] [CrossRef] [PubMed]
  170. Johnsson, P.; Ackley, A.; Vidarsdottir, L.; Lui, W.O.; Corcoran, M.; Grander, D.; Morris, K.V. A pseudogene long-noncoding-RNA network regulates PTEN transcription and translation in human cells. Nat. Struct. Mol. Biol. 2013, 20, 440–446. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  171. Halley, P.; Kadakkuzha, B.M.; Faghihi, M.A.; Magistri, M.; Zeier, Z.; Khorkova, O.; Coito, C.; Hsiao, J.; Lawrence, M.; Wahlestedt, C. Regulation of the apolipoprotein gene cluster by a long noncoding RNA. Cell Rep. 2014, 6, 222–230. [Google Scholar] [CrossRef] [Green Version]
  172. Klase, Z.; Yedavalli, V.S.; Houzet, L.; Perkins, M.; Maldarelli, F.; Brenchley, J.; Strebel, K.; Liu, P.; Jeang, K.T. Activation of HIV-1 from latent infection via synergy of RUNX1 inhibitor Ro5-3335 and SAHA. PLoS Pathog. 2014, 10, e1003997. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  173. Zhang, J.; Thakuri, B.K.C.; Zhao, J.; Nguyen, L.N.; Nguyen, L.N.T.; Khanal, S.; Cao, D.; Dang, X.; Schank, M.; Lu, Z.; et al. Long Noncoding RNA RUNXOR Promotes Myeloid-Derived Suppressor Cell Expansion and Functions via Enhancing Immunosuppressive Molecule Expressions during Latent HIV Infection. J. Immunol. 2021, 206, 2052–2060. [Google Scholar] [CrossRef] [PubMed]
  174. Barriocanal, M.; Carnero, E.; Segura, V.; Fortes, P. Long Non-Coding RNA BST2/BISPR is Induced by IFN and Regulates the Expression of the Antiviral Factor Tetherin. Front. Immunol. 2014, 5, 655. [Google Scholar] [CrossRef]
  175. Kambara, H.; Gunawardane, L.; Zebrowski, E.; Kostadinova, L.; Jobava, R.; Krokowski, D.; Hatzoglou, M.; Anthony, D.D.; Valadkhan, S. Regulation of Interferon-Stimulated Gene BST2 by a lncRNA Transcribed from a Shared Bidirectional Promoter. Front. Immunol. 2014, 5, 676. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  176. Vigneau, S.; Rohrlich, P.S.; Brahic, M.; Bureau, J.F. Tmevpg1, a candidate gene for the control of Theiler’s virus persistence, could be implicated in the regulation of gamma interferon. J. Virol. 2003, 77, 5632–5638. [Google Scholar] [CrossRef] [Green Version]
  177. Collier, S.P.; Collins, P.L.; Williams, C.L.; Boothby, M.R.; Aune, T.M. Cutting edge: Influence of Tmevpg1, a long intergenic noncoding RNA, on the expression of Ifng by Th1 cells. J. Immunol. 2012, 189, 2084–2088. [Google Scholar] [CrossRef] [Green Version]
  178. Collier, S.P.; Henderson, M.A.; Tossberg, J.T.; Aune, T.M. Regulation of the Th1 genomic locus from Ifng through Tmevpg1 by T-bet. J. Immunol. 2014, 193, 3959–3965. [Google Scholar] [CrossRef] [Green Version]
  179. Gomez, J.A.; Wapinski, O.L.; Yang, Y.W.; Bureau, J.F.; Gopinath, S.; Monack, D.M.; Chang, H.Y.; Brahic, M.; Kirkegaard, K. The NeST long ncRNA controls microbial susceptibility and epigenetic activation of the interferon-gamma locus. Cell 2013, 152, 743–754. [Google Scholar] [CrossRef] [Green Version]
  180. Petermann, F.; Pekowska, A.; Johnson, C.A.; Jankovic, D.; Shih, H.Y.; Jiang, K.; Hudson, W.H.; Brooks, S.R.; Sun, H.W.; Villarino, A.V.; et al. The Magnitude of IFN-gamma Responses Is Fine-Tuned by DNA Architecture and the Non-coding Transcript of Ifng-as1. Mol. Cell 2019, 75, 1229–1242 e1225. [Google Scholar] [CrossRef]
  181. Stein, N.; Berhani, O.; Schmiedel, D.; Duev-Cohen, A.; Seidel, E.; Kol, I.; Tsukerman, P.; Hecht, M.; Reches, A.; Gamliel, M.; et al. IFNG-AS1 Enhances Interferon Gamma Production in Human Natural Killer Cells. iScience 2019, 11, 466–473. [Google Scholar] [CrossRef] [Green Version]
  182. Goris, A.; Heggarty, S.; Marrosu, M.G.; Graham, C.; Billiau, A.; Vandenbroeck, K. Linkage disequilibrium analysis of chromosome 12q14-15 in multiple sclerosis: Delineation of a 118-kb interval around interferon-gamma (IFNG) that is involved in male versus female differential susceptibility. Genes. Immun. 2002, 3, 470–476. [Google Scholar] [CrossRef]
  183. Latiano, A.; Palmieri, O.; Latiano, T.; Corritore, G.; Bossa, F.; Martino, G.; Biscaglia, G.; Scimeca, D.; Valvano, M.R.; Pastore, M.; et al. Investigation of multiple susceptibility loci for inflammatory bowel disease in an Italian cohort of patients. PLoS ONE 2011, 6, e22688. [Google Scholar] [CrossRef] [Green Version]
  184. Silverberg, M.S.; Cho, J.H.; Rioux, J.D.; McGovern, D.P.; Wu, J.; Annese, V.; Achkar, J.P.; Goyette, P.; Scott, R.; Xu, W.; et al. Ulcerative colitis-risk loci on chromosomes 1p36 and 12q15 found by genome-wide association study. Nat. Genet. 2009, 41, 216–220. [Google Scholar] [CrossRef]
  185. Xu-Yang, Z.; Pei-Yu, B.; Chuan-Tao, Y.; Wei, Y.; Hong-Wei, M.; Kang, T.; Chun-Mei, Z.; Ying-Feng, L.; Xin, W.; Ping-Zhong, W.; et al. Interferon-Induced Transmembrane Protein 3 Inhibits Hantaan Virus Infection, and Its Single Nucleotide Polymorphism rs12252 Influences the Severity of Hemorrhagic Fever with Renal Syndrome. Front. Immunol. 2016, 7, 535. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  186. Mariotti, B.; Servaas, N.H.; Rossato, M.; Tamassia, N.; Cassatella, M.A.; Cossu, M.; Beretta, L.; van der Kroef, M.; Radstake, T.; Bazzoni, F. The Long Non-coding RNA NRIR Drives IFN-Response in Monocytes: Implication for Systemic Sclerosis. Front. Immunol. 2019, 10, 100. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  187. Imamura, K.; Imamachi, N.; Akizuki, G.; Kumakura, M.; Kawaguchi, A.; Nagata, K.; Kato, A.; Kawaguchi, Y.; Sato, H.; Yoneda, M.; et al. Long noncoding RNA NEAT1-dependent SFPQ relocation from promoter region to paraspeckle mediates IL8 expression upon immune stimuli. Mol. Cell 2014, 53, 393–406. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  188. Matthys, V.S.; Cimica, V.; Dalrymple, N.A.; Glennon, N.B.; Bianco, C.; Mackow, E.R. Hantavirus GnT elements mediate TRAF3 binding and inhibit RIG-I/TBK1-directed beta interferon transcription by blocking IRF3 phosphorylation. J. Virol. 2014, 88, 2246–2259. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  189. Oshiumi, H.; Miyashita, M.; Okamoto, M.; Morioka, Y.; Okabe, M.; Matsumoto, M.; Seya, T. DDX60 Is Involved in RIG-I-Dependent and Independent Antiviral Responses, and Its Function Is Attenuated by Virus-Induced EGFR Activation. Cell Rep. 2015, 11, 1193–1207. [Google Scholar] [CrossRef] [PubMed]
  190. Boliar, S.; Gludish, D.W.; Jambo, K.C.; Kamng’ona, R.; Mvaya, L.; Mwandumba, H.C.; Russell, D.G. Inhibition of the lncRNA SAF drives activation of apoptotic effector caspases in HIV-1-infected human macrophages. Proc. Natl. Acad. Sci. USA 2019, 116, 7431–7438. [Google Scholar] [CrossRef] [Green Version]
  191. Young, L.S.; Rickinson, A.B. Epstein-Barr virus: 40 years on. Nat. Rev. Cancer 2004, 4, 757–768. [Google Scholar] [CrossRef] [PubMed]
  192. Zuo, L.; Yue, W.; Du, S.; Xin, S.; Zhang, J.; Liu, L.; Li, G.; Lu, J. An update: Epstein-Barr virus and immune evasion via microRNA regulation. Virol. Sin. 2017, 32, 175–187. [Google Scholar] [CrossRef]
  193. Qiu, J.; Smith, P.; Leahy, L.; Thorley-Lawson, D.A. The Epstein-Barr virus encoded BART miRNAs potentiate tumor growth in vivo. PLoS Pathog. 2015, 11, e1004561. [Google Scholar] [CrossRef] [Green Version]
  194. Chen, H.L.; Lung, M.M.; Sham, J.S.; Choy, D.T.; Griffin, B.E.; Ng, M.H. Transcription of BamHI-A region of the EBV genome in NPC tissues and B cells. Virology 1992, 191, 193–201. [Google Scholar] [CrossRef] [PubMed]
  195. Smith, P.R.; de Jesus, O.; Turner, D.; Hollyoake, M.; Karstegl, C.E.; Griffin, B.E.; Karran, L.; Wang, Y.; Hayward, S.D.; Farrell, P.J. Structure and coding content of CST (BART) family RNAs of Epstein-Barr virus. J. Virol. 2000, 74, 3082–3092. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  196. Verhoeven, R.J.; Tong, S.; Zhang, G.; Zong, J.; Chen, Y.; Jin, D.Y.; Chen, M.R.; Pan, J.; Chen, H. NF-kappaB Signaling Regulates Expression of Epstein-Barr Virus BART MicroRNAs and Long Noncoding RNAs in Nasopharyngeal Carcinoma. J. Virol. 2016, 90, 6475–6488. [Google Scholar] [CrossRef] [Green Version]
  197. Kang, G.H.; Lee, S.; Kim, W.H.; Lee, H.W.; Kim, J.C.; Rhyu, M.G.; Ro, J.Y. Epstein-barr virus-positive gastric carcinoma demonstrates frequent aberrant methylation of multiple genes and constitutes CpG island methylator phenotype-positive gastric carcinoma. Am. J. Pathol. 2002, 160, 787–794. [Google Scholar] [CrossRef] [Green Version]
  198. Zhang, J.; Li, X.; Hu, J.; Cao, P.; Yan, Q.; Zhang, S.; Dang, W.; Lu, J. Long noncoding RNAs involvement in Epstein-Barr virus infection and tumorigenesis. Virol. J. 2020, 17, 51. [Google Scholar] [CrossRef] [Green Version]
  199. Tai-Schmiedel, J.; Karniely, S.; Lau, B.; Ezra, A.; Eliyahu, E.; Nachshon, A.; Kerr, K.; Suarez, N.; Schwartz, M.; Davison, A.J.; et al. Human cytomegalovirus long noncoding RNA4.9 regulates viral DNA replication. PLoS Pathog. 2020, 16, e1008390. [Google Scholar] [CrossRef] [Green Version]
  200. Lee, S.; Kim, H.; Hong, A.; Song, J.; Lee, S.; Kim, M.; Hwang, S.Y.; Jeong, D.; Kim, J.; Son, A.; et al. Functional and molecular dissection of HCMV long non-coding RNAs. Sci. Rep. 2022, 12, 19303. [Google Scholar] [CrossRef]
  201. Murphy, J.C.; Fischle, W.; Verdin, E.; Sinclair, J.H. Control of cytomegalovirus lytic gene expression by histone acetylation. EMBO J. 2002, 21, 1112–1120. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  202. Reeves, M.B.; Lehner, P.J.; Sissons, J.G.P.; Sinclair, J.H. An in vitro model for the regulation of human cytomegalovirus latency and reactivation in dendritic cells by chromatin remodelling. J. Gen. Virol. 2005, 86, 2949–2954. [Google Scholar] [CrossRef] [PubMed]
  203. Reeves, M.B.; MacAry, P.A.; Lehner, P.J.; Sissons, J.G.; Sinclair, J.H. Latency, chromatin remodeling, and reactivation of human cytomegalovirus in the dendritic cells of healthy carriers. Proc. Natl. Acad. Sci. USA 2005, 102, 4140–4145. [Google Scholar] [CrossRef] [Green Version]
  204. Sinclair, J.; Sissons, P. Latency and reactivation of human cytomegalovirus. J. Gen. Virol. 2006, 87, 1763–1779. [Google Scholar] [CrossRef] [PubMed]
  205. Huang, Y.; Guo, X.; Zhang, J.; Li, J.; Xu, M.; Wang, Q.; Liu, Z.; Ma, Y.; Qi, Y.; Ruan, Q. Human cytomegalovirus RNA2.7 inhibits RNA polymerase II (Pol II) Serine-2 phosphorylation by reducing the interaction between Pol II and phosphorylated cyclin-dependent kinase 9 (pCDK9). Virol. Sin. 2022, 37, 358–369. [Google Scholar] [CrossRef] [PubMed]
  206. Reeves, M.B.; Davies, A.A.; McSharry, B.P.; Wilkinson, G.W.; Sinclair, J.H. Complex I binding by a virally encoded RNA regulates mitochondria-induced cell death. Science 2007, 316, 1345–1348. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  207. Perera, M.R.; Sinclair, J.H. The Human Cytomegalovirus beta2.7 Long Non-Coding RNA Prevents Induction of Reactive Oxygen Species to Maintain Viral Gene Silencing during Latency. Int. J. Mol. Sci. 2022, 23, 11017. [Google Scholar] [CrossRef]
  208. Lau, B.; Kerr, K.; Camiolo, S.; Nightingale, K.; Gu, Q.; Antrobus, R.; Suarez, N.M.; Loney, C.; Stanton, R.J.; Weekes, M.P.; et al. Human Cytomegalovirus RNA2.7 Is Required for Upregulating Multiple Cellular Genes To Promote Cell Motility and Viral Spread Late in Lytic Infection. J. Virol. 2021, 95, e0069821. [Google Scholar] [CrossRef]
  209. Lau, B.; Kerr, K.; Gu, Q.; Nightingale, K.; Antrobus, R.; Suarez, N.M.; Stanton, R.J.; Wang, E.C.Y.; Weekes, M.P.; Davison, A.J. Human Cytomegalovirus Long Non-coding RNA1.2 Suppresses Extracellular Release of the Pro-inflammatory Cytokine IL-6 by Blocking NF-kappaB Activation. Front. Cell Infect. Microbiol. 2020, 10, 361. [Google Scholar] [CrossRef]
  210. Amelio, A.L.; Giordani, N.V.; Kubat, N.J.; O’Neil J, E.; Bloom, D.C. Deacetylation of the herpes simplex virus type 1 latency-associated transcript (LAT) enhancer and a decrease in LAT abundance precede an increase in ICP0 transcriptional permissiveness at early times postexplant. J. Virol. 2006, 80, 2063–2068. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  211. Cliffe, A.R.; Coen, D.M.; Knipe, D.M. Kinetics of facultative heterochromatin and polycomb group protein association with the herpes simplex viral genome during establishment of latent infection. mBio 2013, 4, e00590-12. [Google Scholar] [CrossRef] [Green Version]
  212. Cliffe, A.R.; Garber, D.A.; Knipe, D.M. Transcription of the herpes simplex virus latency-associated transcript promotes the formation of facultative heterochromatin on lytic promoters. J. Virol. 2009, 83, 8182–8190. [Google Scholar] [CrossRef] [Green Version]
  213. Kwiatkowski, D.L.; Thompson, H.W.; Bloom, D.C. The polycomb group protein Bmi1 binds to the herpes simplex virus 1 latent genome and maintains repressive histone marks during latency. J. Virol. 2009, 83, 8173–8181. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  214. Wang, Q.Y.; Zhou, C.; Johnson, K.E.; Colgrove, R.C.; Coen, D.M.; Knipe, D.M. Herpesviral latency-associated transcript gene promotes assembly of heterochromatin on viral lytic-gene promoters in latent infection. Proc. Natl. Acad. Sci. USA 2005, 102, 16055–16059. [Google Scholar] [CrossRef] [Green Version]
  215. Hancock, M.H.; Skalsky, R.L. Roles of Non-coding RNAs During Herpesvirus Infection. Curr. Top. Microbiol. Immunol. 2018, 419, 243–280. [Google Scholar] [CrossRef] [PubMed]
  216. Zhang, Y.; Zeng, L.S.; Wang, J.; Cai, W.Q.; Cui, W.; Song, T.J.; Peng, X.C.; Ma, Z.; Xiang, Y.; Cui, S.Z.; et al. Multifunctional Non-Coding RNAs Mediate Latent Infection and Recurrence of Herpes Simplex Viruses. Infect. Drug. Resist. 2021, 14, 5335–5349. [Google Scholar] [CrossRef] [PubMed]
  217. Arun, G.; Diermeier, S.D.; Spector, D.L. Therapeutic Targeting of Long Non-Coding RNAs in Cancer. Trends Mol. Med. 2018, 24, 257–277. [Google Scholar] [CrossRef]
  218. Lennox, K.A.; Behlke, M.A. Cellular localization of long non-coding RNAs affects silencing by RNAi more than by antisense oligonucleotides. Nucleic Acids Res. 2016, 44, 863–877. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  219. Gagnon, K.T.; Corey, D.R. Guidelines for Experiments Using Antisense Oligonucleotides and Double-Stranded RNAs. Nucleic Acid. Ther. 2019, 29, 116–122. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  220. Abudayyeh, O.O.; Gootenberg, J.S.; Essletzbichler, P.; Han, S.; Joung, J.; Belanto, J.J.; Verdine, V.; Cox, D.B.T.; Kellner, M.J.; Regev, A.; et al. RNA targeting with CRISPR-Cas13. Nature 2017, 550, 280–284. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  221. Cox, D.B.T.; Gootenberg, J.S.; Abudayyeh, O.O.; Franklin, B.; Kellner, M.J.; Joung, J.; Zhang, F. RNA editing with CRISPR-Cas13. Science 2017, 358, 1019–1027. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  222. Konermann, S.; Lotfy, P.; Brideau, N.J.; Oki, J.; Shokhirev, M.N.; Hsu, P.D. Transcriptome Engineering with RNA-Targeting Type VI-D CRISPR Effectors. Cell 2018, 173, 665–676 e614. [Google Scholar] [CrossRef] [Green Version]
  223. Smargon, A.A.; Cox, D.B.T.; Pyzocha, N.K.; Zheng, K.; Slaymaker, I.M.; Gootenberg, J.S.; Abudayyeh, O.A.; Essletzbichler, P.; Shmakov, S.; Makarova, K.S.; et al. Cas13b Is a Type VI-B CRISPR-Associated RNA-Guided RNase Differentially Regulated by Accessory Proteins Csx27 and Csx28. Mol. Cell 2017, 65, 618–630 e617. [Google Scholar] [CrossRef] [Green Version]
  224. Xiang, J.F.; Yin, Q.F.; Chen, T.; Zhang, Y.; Zhang, X.O.; Wu, Z.; Zhang, S.; Wang, H.B.; Ge, J.; Lu, X.; et al. Human colorectal cancer-specific CCAT1-L lncRNA regulates long-range chromatin interactions at the MYC locus. Cell Res. 2014, 24, 513–531. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  225. Bester, A.C.; Lee, J.D.; Chavez, A.; Lee, Y.R.; Nachmani, D.; Vora, S.; Victor, J.; Sauvageau, M.; Monteleone, E.; Rinn, J.L.; et al. An Integrated Genome-wide CRISPRa Approach to Functionalize lncRNAs in Drug Resistance. Cell 2018, 173, 649–664 e620. [Google Scholar] [CrossRef]
  226. Shechner, D.M.; Hacisuleyman, E.; Younger, S.T.; Rinn, J.L. Multiplexable, locus-specific targeting of long RNAs with CRISPR-Display. Nat. Methods 2015, 12, 664–670. [Google Scholar] [CrossRef] [PubMed]
  227. Wang, J.; Zhang, Y.; Li, Q.; Zhao, J.; Yi, D.; Ding, J.; Zhao, F.; Hu, S.; Zhou, J.; Deng, T.; et al. Influenza Virus Exploits an Interferon-Independent lncRNA to Preserve Viral RNA Synthesis through Stabilizing Viral RNA Polymerase PB1. Cell Rep. 2019, 27, 3295–3304 e3294. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  228. Buske, F.A.; Bauer, D.C.; Mattick, J.S.; Bailey, T.L. Triplexator: Detecting nucleic acid triple helices in genomic and transcriptomic data. Genome Res. 2012, 22, 1372–1381. [Google Scholar] [CrossRef] [Green Version]
  229. Buske, F.A.; Bauer, D.C.; Mattick, J.S.; Bailey, T.L. Triplex-Inspector: An analysis tool for triplex-mediated targeting of genomic loci. Bioinformatics 2013, 29, 1895–1897. [Google Scholar] [CrossRef] [Green Version]
  230. Kuo, C.C.; Hanzelmann, S.; Senturk Cetin, N.; Frank, S.; Zajzon, B.; Derks, J.P.; Akhade, V.S.; Ahuja, G.; Kanduri, C.; Grummt, I.; et al. Detection of RNA-DNA binding sites in long noncoding RNAs. Nucleic Acids Res. 2019, 47, e32. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  231. He, S.; Zhang, H.; Liu, H.; Zhu, H. LongTarget: A tool to predict lncRNA DNA-binding motifs and binding sites via Hoogsteen base-pairing analysis. Bioinformatics 2015, 31, 178–186. [Google Scholar] [CrossRef] [Green Version]
  232. Hon, J.; Martinek, T.; Rajdl, K.; Lexa, M. Triplex: An R/Bioconductor package for identification and visualization of potential intramolecular triplex patterns in DNA sequences. Bioinformatics 2013, 29, 1900–1901. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  233. Lexa, M.; Martinek, T.; Burgetova, I.; Kopecek, D.; Brazdova, M. A dynamic programming algorithm for identification of triplex-forming sequences. Bioinformatics 2011, 27, 2510–2517. [Google Scholar] [CrossRef] [Green Version]
  234. Soibam, B. Super-lncRNAs: Identification of lncRNAs that target super-enhancers via RNA:DNA:DNA triplex formation. RNA 2017, 23, 1729–1742. [Google Scholar] [CrossRef] [Green Version]
  235. Schmitt, A.M.; Chang, H.Y. Long Noncoding RNAs in Cancer Pathways. Cancer Cell 2016, 29, 452–463. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  236. Lee, G.L.; Dobi, A.; Srivastava, S. Prostate cancer: Diagnostic performance of the PCA3 urine test. Nature reviews. Urology 2011, 8, 123–124. [Google Scholar] [CrossRef]
  237. Hanna, N.; Ohana, P.; Konikoff, F.M.; Leichtmann, G.; Hubert, A.; Appelbaum, L.; Kopelman, Y.; Czerniak, A.; Hochberg, A. Phase 1/2a, dose-escalation, safety, pharmacokinetic and preliminary efficacy study of intratumoral administration of BC-819 in patients with unresectable pancreatic cancer. Cancer gene therapy 2012, 19, 374–381. [Google Scholar] [CrossRef]
  238. Connelly, C.M.; Moon, M.H.; Schneekloth, J.S., Jr. The Emerging Role of RNA as a Therapeutic Target for Small Molecules. Cell Chem. Biol. 2016, 23, 1077–1090. [Google Scholar] [CrossRef] [PubMed]
  239. Ahmad, A.; Lin, H.; Shatabda, S. Locate-R: Subcellular localization of long non-coding RNAs using nucleotide compositions. Genomics 2020, 112, 2583–2589. [Google Scholar] [CrossRef]
  240. Gudenas, B.L.; Wang, L. Prediction of LncRNA Subcellular Localization with Deep Learning from Sequence Features. Sci. Rep. 2018, 8, 16385. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  241. Li, M.; Zhao, B.; Yin, R.; Lu, C.; Guo, F.; Zeng, M. GraphLncLoc: Long non-coding RNA subcellular localization prediction using graph convolutional networks based on sequence to graph transformation. Brief. Bioinform. 2022, 24, bbac565. [Google Scholar] [CrossRef]
  242. Zeng, M.; Wu, Y.; Lu, C.; Zhang, F.; Wu, F.X.; Li, M. DeepLncLoc: A deep learning framework for long non-coding RNA subcellular localization prediction based on subsequence embedding. Brief. Bioinform. 2022, 23, bbab360. [Google Scholar] [CrossRef] [PubMed]
  243. He, Y.; Han, B.; Ding, Y.; Zhang, H.; Chang, S.; Zhang, L.; Zhao, C.; Yang, N.; Song, J. Linc-GALMD1 Regulates Viral Gene Expression in the Chicken. Front. Genet. 2019, 10, 1122. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  244. Munschauer, M.; Nguyen, C.T.; Sirokman, K.; Hartigan, C.R.; Hogstrom, L.; Engreitz, J.M.; Ulirsch, J.C.; Fulco, C.P.; Subramanian, V.; Chen, J.; et al. The NORAD lncRNA assembles a topoisomerase complex critical for genome stability. Nature 2018, 561, 132–136. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  245. Shen, Y.; Liu, S.; Fan, J.; Jin, Y.; Tian, B.; Zheng, X.; Fu, H. Nuclear retention of the lncRNA SNHG1 by doxorubicin attenuates hnRNPC-p53 protein interactions. EMBO Rep. 2017, 18, 536–548. [Google Scholar] [CrossRef] [Green Version]
  246. Carrieri, C.; Cimatti, L.; Biagioli, M.; Beugnet, A.; Zucchelli, S.; Fedele, S.; Pesce, E.; Ferrer, I.; Collavin, L.; Santoro, C.; et al. Long non-coding antisense RNA controls Uchl1 translation through an embedded SINEB2 repeat. Nature 2012, 491, 454–457. [Google Scholar] [CrossRef] [Green Version]
  247. Cai, R.; Sun, Y.; Qimuge, N.; Wang, G.; Wang, Y.; Chu, G.; Yu, T.; Yang, G.; Pang, W. Adiponectin AS lncRNA inhibits adipogenesis by transferring from nucleus to cytoplasm and attenuating Adiponectin mRNA translation. Biochim. Biophys. Acta Mol. Cell Biol. Lipids 2018, 1863, 420–432. [Google Scholar] [CrossRef]
  248. Zhang, L.; Zheng, X.; Li, J.; Wang, G.; Hu, Z.; Chen, Y.; Wang, X.; Gu, M.; Gao, R.; Hu, S.; et al. Long noncoding RNA#45 exerts broad inhibitory effect on influenza a virus replication via its stem ring arms. Virulence 2021, 12, 2443–2460. [Google Scholar] [CrossRef]
  249. Guo, C.J.; Ma, X.K.; Xing, Y.H.; Zheng, C.C.; Xu, Y.F.; Shan, L.; Zhang, J.; Wang, S.; Wang, Y.; Carmichael, G.G.; et al. Distinct Processing of lncRNAs Contributes to Non-conserved Functions in Stem Cells. Cell 2020, 181, 621–636 e622. [Google Scholar] [CrossRef]
  250. Hezroni, H.; Koppstein, D.; Schwartz, M.G.; Avrutin, A.; Bartel, D.P.; Ulitsky, I. Principles of long noncoding RNA evolution derived from direct comparison of transcriptomes in 17 species. Cell Rep. 2015, 11, 1110–1122. [Google Scholar] [CrossRef] [Green Version]
  251. Quinn, J.J.; Zhang, Q.C.; Georgiev, P.; Ilik, I.A.; Akhtar, A.; Chang, H.Y. Rapid evolutionary turnover underlies conserved lncRNA-genome interactions. Genes. Dev. 2016, 30, 191–207. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  252. Burnett, J.C.; Rossi, J.J. RNA-based therapeutics: Current progress and future prospects. Chemistry & biology 2012, 19, 60–71. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Virus-induced nuclear lncRNAs regulate gene expression using various mechanisms, such as (A) Direct interaction of Lnc-MxA with the promoter region of IFNB1, which inhibits recruitment of transcription factors and IFNB1 transcription; (B) HEAL lncRNA recruits histone-modifying proteins and transcription elongation factors at HIV promoter and enhances HIV transcription; (C) LUCAT1 sequesters STAT1 transcription factor and inhibits IFNB1 transcription, whereas LUARIS localizes ATF2 to the promoter region of interferon-stimulated genes (ISGs) and enhances their expression; (D) RUNX mRNA transcription and isoform expression are regulated by neighboring lncRNAs RUNXOR and LINC01426; (E) Ifng-as1 transcript is likely to enhance Ifng expression by recruiting and enriching transcriptional enhancers (transcription factor or chromatin modifiers). In addition, the Ifng-as1 locus impacts the chromatin organization independent of the Ifng-as1 transcription or lncRNA sequence through a CTCF-binding site encoded within the Ifng-as1 gene region; (F) BART lncRNA stalls Pol II at the promoter region of IFNB1 and inhibits its transcription. BART lncRNA also associates with CREB-binding protein (CBP/300) and inhibits its histone acetylation activity, thus inhibiting gene expression.
Figure 1. Virus-induced nuclear lncRNAs regulate gene expression using various mechanisms, such as (A) Direct interaction of Lnc-MxA with the promoter region of IFNB1, which inhibits recruitment of transcription factors and IFNB1 transcription; (B) HEAL lncRNA recruits histone-modifying proteins and transcription elongation factors at HIV promoter and enhances HIV transcription; (C) LUCAT1 sequesters STAT1 transcription factor and inhibits IFNB1 transcription, whereas LUARIS localizes ATF2 to the promoter region of interferon-stimulated genes (ISGs) and enhances their expression; (D) RUNX mRNA transcription and isoform expression are regulated by neighboring lncRNAs RUNXOR and LINC01426; (E) Ifng-as1 transcript is likely to enhance Ifng expression by recruiting and enriching transcriptional enhancers (transcription factor or chromatin modifiers). In addition, the Ifng-as1 locus impacts the chromatin organization independent of the Ifng-as1 transcription or lncRNA sequence through a CTCF-binding site encoded within the Ifng-as1 gene region; (F) BART lncRNA stalls Pol II at the promoter region of IFNB1 and inhibits its transcription. BART lncRNA also associates with CREB-binding protein (CBP/300) and inhibits its histone acetylation activity, thus inhibiting gene expression.
Cells 12 00987 g001
Table 1. Cellular and viral nuclear lncRNAs regulate viral replication and persistence.
Table 1. Cellular and viral nuclear lncRNAs regulate viral replication and persistence.
LncRNAVirus MechanismReference
Proviral lncRNAs
NRAVIAV, SeV, MDRV, HSVHistone modification and reduction in active transcription marks at ISG.Ouyang et al., 2014 [57]
TSPOAP1-AS1IAVTSPOAP1-AS1 inhibits FNβ1 transcription, ISRE activation, and ISG expression.Wang et al., 2019 [62]
Lnc-MxAIAVLnc-MxA inhibits IFNβ transcription by binding to its promoter and enhances viral replication.Li et al., 2019 [61]
VINIAV, VSVVIN increases virus replication and viral gene expression. Molecular mechanisms are unknown.Winterling et al., 2014 [29]
EGOTHCV, IAV, SFVEGOT inhibits the expression of several ISGs and enhances viral replication. Molecular mechanisms are unknown.Carnero et al., 2016 [31]
LetheHCVLethe inhibits RelA-mediated DNA-binding; inhibits expression of antiviral factors, protein kinase R (PKR), 2′,5′-oligoadenylate synthetase (OAS) proteins, and Interferon Regulatory Factor 1 (IRF1), and enhances HCV replicationRapicavoli et al., 2013 [66]; Xiong et al., 2015 [71]
LncRNA RP11- 288L9.4HCVTSPOAP1-AS1 inhibits expression of IFNα-inducible protein 6 (IFI6) by histone modification and enhances HCV replication.Liu et al., 2019 [72]
NRIRHCVNRIR inhibits transcription of several interferon-stimulated genes (ISG) and enhances HCV replication.Kambara et al., 2014 [65]
Lnc_000641pseudorabies virus (PRV)Lnc_000641 inhibits IFNα transcription, phosphorylation of transcription factors (Jak and STAT1), and increases PRV replication.Fang et al., 2021 [73]
NEAT1HSV-1NEAT1 recruits STAT3 to viral gene promoters to increase viral gene expression.Wang et al., 2017 [74]
Antiviral lncRNA
IVRPIEIAVIVRPIE upregulates IFNβ and several ISGs, including IRF1, IFIT1, IFIT3, Mx1, ISG15, and IFI44L, by affecting histone modification of these genes.Zhao et al., 2020 [63]
OASL-IT1ZKIVOASL-IT1 enhances expression of IFN-β, Mx1, IFITM1 and inhibits ZKIV replication.Wang et al., 2021 [75]
LUARISEMCV,
HBV, HCV
LUARIS upregulated the level of IFN-stimulated genes through interactions with hnRNPU and ATF2 and suppressed EMCV, HBV, and HCV.Nishitsuji et al., 2016 [76]
NEAT1Hantaan
virus
NEAT1 relocates SFPQ to paraspeckles, increases RIG-I and DDX60 transcription, increases IFN- γ, and inhibits virus.Ma et al., 2017 [56]
LncRNAs influence the long-term persistence of the virus
NRONHIVNRON mediates degradation of HIV Tat protein.Li et al., 2016 [77]
MALAT1HIVMALAT1 promotes HIV reactivation from latent provirus.Qu et al., 2019 [78]
7SKHIV7SK promotes HIV latency by inactivating p-TEFb.Nguyen et al., 2001 [79] Contreras et al., 2007 [80]; Budhiraja et al.,2013 [81]; Eilebrecht et al., 2017 [82]
uc002yug.2HIVuc002yug.2 promotes viral reactivation by inhibition of Transcription Repressor RUNX1.Huan et al., 2018 [83]
lincRNA-p21HIVlincRNA-p21 inhibits DSB-induced cell death, promotes viral persistence.Barichievy et al., 2018 [84]
HEALHIVHEAL promotes viral reactivation by recruiting histone acetyltransferase p300 to HIV-1 promoter region.Chao et al.,2019 [85]
NEAT1HIVNEAT1 sequesters unspliced HIV transcripts in nuclear paraspeckle bodies promoting long-term persistence of HIV.Zhang et al., 2013 [51]
HIV antisense lncRNAHIVHIV antisense lncRNA recruits chromatin remodeling proteins such as DNMT3a, the enhancer of Zeste 2 (EZH2), and histone deacetylase 1 (HDAC-1) to HIV 5′long terminal repeat. These proteins bring about H3K9 dimethylation, H3K27 trimethylation, and histone deacetylation, resulting in epigenetic silencing of viral transcription.Saayman et al., 2014 [86]
KSHV-encoded
PAN RNA
KSHVPAN RNA binds lysine demethylases UTX and JMJD3, and the lysine methyltransferase MLL2 facilitates the recruitment of histone demethylases to the viral chromatin.Rossetto et al., 2012 [87] Rossetto, 2013 [88]
Rossetto, 2016 [89]
HCMV-encoded RNA4.9HCMVRNA4.9 tethers the components of the polycomb repression complex (PRC) to the major immediate early promoter region (MIEP) and represses viral transcription.Rosseto., 2013 [90]
EBV-encoded
BART lncRNA
EBV-associated epithelial tumorsBART lncRNAs downregulate the expression of the tumor suppressor gene RASA1 and unfolded protein response (UPR) genes. BART lncRNAs regulate host gene expressions through chromatin modification.Marquitz.,2015 [91];
Verhoeven, 2019 [92]
EBV-encoded lncRNA BHLF1EBV-associated epithelial tumorsBHLF1 localizes at the surface of the viral replication compartment and forms an RNA–DNA hybrid at the site of virus transcription.Park & Miller, 2018 [93]; Rennekamp & Lieberman, 2011 [94]
Table 2. Methods to modulate lncRNA expression for functional studies.
Table 2. Methods to modulate lncRNA expression for functional studies.
MethodUseAdvantagesLimitation
siRNA/shRNAKnockdownInexpensive, cost-effective for large-scale screeningNuclear lncRNAs cannot be targeted efficiently by siRNA; structural constraints limit accessibility, large-scale off-target cleavage, and knockdown may be short-lived.
Antisense Oligo (ASO)KnockdownEfficient degradation of nuclear lncRNAStructural constraints limit accessibility, large-scale off-target cleavage, and knockdown may be short-lived.
CRISPR/Cas9Gene knockout or knock-inEasily programmable to target genes of interest, most definitive CRISPR/Cas9-mediated frameshift mutations are not helpful for most lncRNAs as their functional sequence motifs are unknown. CRISPR/Cas9 excision of the entire lncRNA gene may disrupt overlapping coding or noncoding RNA region.
CRISPRiInhibition of transcriptionEasily programmable to target genes of interestCRISPRi may deregulate overlapping coding or noncoding RNA region, the functions of lncRNA transcript from those of promoter or enhancer element encoded within the lncRNA locus or small peptide encoded by the transcript.
CRISPR/Cas13dKnockdownEasily programmable, independent of PAM, superior RNA knockdown efficiency and dramatically higher specificity than currently available methods, stable long-term expressionCannot decipher the function of enhancer element encoded within the lncRNA locus or small peptide encoded by the transcript
CRISPRaActivation of transcriptionEasily programmable, enhanced lncRNA expression from the endogenous lociDependence on protospacer-adjacent motif (PAM); may deregulate overlapping coding or noncoding RNA region; and the functions of lncRNA transcript cannot be distinguished from those of promoter or enhancer element encoded within the lncRNA locus or small peptide encoded by the transcript.
CRISPR-display Easily programmable, allows site-specific delivery of lncRNA transcript to desired genomic loci; this method can be used to test both cis and trans effects of lncRNA transcripts and distinguish them from the act of lncRNA transcription. Limited by the number of available functional RNA motifs and RNA-binding protein functions
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Kulkarni, V.; Jayakumar, S.; Mohan, M.; Kulkarni, S. Aid or Antagonize: Nuclear Long Noncoding RNAs Regulate Host Responses and Outcomes of Viral Infections. Cells 2023, 12, 987. https://doi.org/10.3390/cells12070987

AMA Style

Kulkarni V, Jayakumar S, Mohan M, Kulkarni S. Aid or Antagonize: Nuclear Long Noncoding RNAs Regulate Host Responses and Outcomes of Viral Infections. Cells. 2023; 12(7):987. https://doi.org/10.3390/cells12070987

Chicago/Turabian Style

Kulkarni, Viraj, Sahana Jayakumar, Mahesh Mohan, and Smita Kulkarni. 2023. "Aid or Antagonize: Nuclear Long Noncoding RNAs Regulate Host Responses and Outcomes of Viral Infections" Cells 12, no. 7: 987. https://doi.org/10.3390/cells12070987

APA Style

Kulkarni, V., Jayakumar, S., Mohan, M., & Kulkarni, S. (2023). Aid or Antagonize: Nuclear Long Noncoding RNAs Regulate Host Responses and Outcomes of Viral Infections. Cells, 12(7), 987. https://doi.org/10.3390/cells12070987

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