Skip to main content

Origin and functional role of antisense transcription in endogenous and exogenous retroviruses

Abstract

Most proteins expressed by endogenous and exogenous retroviruses are encoded in the sense (positive) strand of the genome and are under the control of regulatory elements within the 5’ long terminal repeat (LTR). A number of retroviral genomes also encode genes in the antisense (negative) strand and their expression is under the control of negative sense promoters within the 3’ LTR. In the case of the Human T-cell Lymphotropic Virus 1 (HTLV-1), the antisense protein HBZ has been shown to play a critical role in the virus lifecycle and in the pathogenic process, while the function of the Human Immunodeficiency Virus 1 (HIV-1) antisense protein ASP remains unknown. However, the expression of 3’ LTR-driven antisense transcripts is not always demonstrably associated with the presence of an antisense open reading frame encoding a viral protein. Moreover, even in the case of retroviruses that do express an antisense protein, such as HTLV-1 and the pandemic strains of HIV-1, the 3’ LTR-driven antisense transcript shows both protein-coding and noncoding activities. Indeed, the ability to express antisense transcripts appears to be phylogenetically more widespread among endogenous and exogenous retroviruses than the presence of a functional antisense open reading frame within these transcripts. This suggests that retroviral antisense transcripts may have originated as noncoding molecules with regulatory activity that in some cases later acquired protein-coding function. Here, we will review examples of endogenous and exogenous retroviral antisense transcripts, and the ways through which they benefit viral persistence in the host.

Introduction

Retroviruses (Retroviridae) are a family of positive-sense single stranded RNA (ssRNA) viruses that infect vertebrates [1]. They are classified into two subfamilies (Orthoretrovirinae and Spumavirinae) and several genera. The defining feature of retroviruses is that they encode a reverse transcriptase (RT, an RNA- and DNA-dependent DNA polymerase) that converts the ssRNA genome into a dsDNA molecule, and an integrase (IN) that inserts the dsDNA molecule into the host genome.

All retroviral genomes include five essential genetic elements: two identical long terminal repeats (LTR), and three essential genes: the structural genes gag and env, and the enzymatic genes pol/pro. The two LTRs are identical direct repeat sequences located at the 5’ and 3’ ends of the genome, they are arranged in a tail-to-head orientation and have regulatory roles: the 5’ LTR contains the promoter driving expression of the viral genes, and the 3’ LTR contains the polyadenylation signal. Retroviral genera that contain only these five elements (alpha-, beta-, and gammaretroviruses) are said to have a “simple” genome. On the other hand, deltaretroviruses, epsilonretroviruses, lentiviruses, and all spumaviruses encode additional regulatory and accessory proteins, and thus have a “complex” genome. In many cases, these genes overlap each other and are encoded in different reading frames. Their expression involves two or three classes of transcripts produced from a single RNA molecule encompassing the entire viral genome: full length and singly spliced transcripts for retroviruses with simple genome, plus multiply spliced transcripts for retroviruses with complex genomes [1].

Over millions of years, now-extinct ancestral exogenous retroviruses infected and colonized their hosts and persist to this day as endogenous retroviruses (ERV) [2,3,4]. While present-day exogenous retroviruses only infect somatic cells, ancestral forms may have also been capable of infecting cells of the germline, which ensured that they may be transmitted vertically and become fixed in the population [4, 5]. Today, ERVs occupy a large fraction of vertebrate genomes. In humans, endogenous retroviruses (HERV) account for ~ 8% of the genome with ~ 700,000 loci [6, 7]. HERVs are broadly grouped into Class I (similar to gamma- and epsilonretroviruses, Class II (similar to betaretroviruses), and Class III (similar to spumaviruses). While most ERVs present a simple genome that includes the five basic genetic elements [4], some (such as HERV-K) can under certain circumstances express additional proteins [8, 9].

Retroviral LTRs as bidirectional promoters

The retroviral 5’ and 3’ LTRs play regulatory roles in viral expression. Specifically, the U3 region of the 5’ LTR – organized in promoter, enhancer, and modulatory domains – contains binding motifs for a wide array of transcription factors that promote and regulate expression of the viral genes. Retroviruses such as HTLV-1 and HIV-1 also express regulatory proteins that function as transactivators and augment transcription from their own 5’ LTR through positive-feedback loops. The HTLV-1 Tax protein promotes transcription by recruiting to the 5’ LTR transcription factors of the ATF/CREB family as well as p300/CBP [10]. The HIV-1 Tat protein recruits P-TEFb to the 5’ LTR, which in turn increases processivity of RNA polymerase II and promotes transcription elongation [11]. At the other end of the proviral genome, the 3’ LTR contains the motifs required for polyadenylation of the pre-mRNA, which ensures its proper processing, nuclear export, stability, and translation [12].

As discussed above, 5’ and 3’ LTRs are identical direct repeats and contain the same transcriptional regulatory elements. While the presence of a polyadenylation signal in the 3’ LTR is essential to proper processing the viral transcripts, the presence of a polyadenylation signal in the 5’ LTR could impact viral expression. The location of the polyadenylation signals separates retroviruses into two groups [13]. On one hand, viruses such as HTLV-1, HTLV-2, BLV, RSV, and MMTV use bipartite polyadenylations signals: an AAUAAA sequence (recognized by the cleavage and polyadenylation specific factor, CPSF) located in the U3 region upstream of the transcription start site (U3-R boundary), and a GU/U-rich sequence (recognized by the cleavage stimulation factor, CstF) positioned in the beginning of the U5 region. Therefore, the full polyadenylation signal is present only at the 3’ end of the viral transcripts (3’ LTR). Further, in RSV and MMTV, the R region is very short, thus keeping the two signals at a functional distance. However, in HTLV-1, the R region is 275 bp, and functional proximity of the two polyadenylation signals is achieved via secondary structure (looping) of the viral transcript [13]. In the second group of retroviruses (e.g., HIV-1, HIV-2, EIAV, MoMLV) both polyadenylation signals (AAUAAA and GU/U-rich sequences) are located in the R region, and therefore are transcribed at both ends of the viral RNA molecule. In the case of HIV-1, two mechanisms ensure utilization exclusively of the polyadenylation signals at the 3’ end of the transcripts: the presence of U3-derived sequences with “polyadenylation enhancer” activity (which are not present at the 5’ end) [14,15,16,17], and the formation of secondary structures that suppress the polyadenylation activity of the signals at the 5’ end of the transcript [18,19,20]. Suppression of the 5’ polyadenylation signal has also been shown to involve binding the of U1 snRNP splicing factor to the major splice donor site at the 5’ end of the viral RNA [14, 21, 22].

Several studies have provided mounting evidence that retroviral LTRs are capable of bidirectional transcription. An early report by Larocca and colleagues showed that the 3’ LTR of HTLV-1 can direct antisense transcription in a Tax-independent manner [23]. A more recent study showed that sense and antisense transcription across the HTLV-1 provirus do not interfere with each other and with the expression of Tax or the HTLV-1 antisense protein, HBZ [24]. The same study also showed that, in the absence of Tax, antisense transcription predominates [24]. A negative sense promoter has also been identified within the 3’ LTR of HIV-1 [25,26,27]. Antisense transcription driven by the LTR of both HTLV-1 and HIV-1 is independent of – or inhibited by – their respective transactivators, Tax and Tat [25,26,27]. Interestingly, the ability to drive bidirectional transcription is not limited to complex retroviruses. Indeed, the LTRs of the simple retrovirus, murine leukemia virus (MLV) and also endogenous retroviruses can direct sense and antisense transcription [26, 28,29,30,31].

Altogether, bidirectional transcriptional activity is widespread among simple and complex exogenous retroviruses as well as endogenous retroviruses. This suggests that this property has an ancestral origin, it has been conserved in present-day retroviruses, and thus it serves a purpose that benefits viral persistence or spread.

Antisense transcription in exogenous retroviruses

HIV-1 and other lentiviruses. The first retroviral antisense gene to be described was the antisense protein (asp) gene of HIV-1 [32]. The asp ORF was identified through sequence analysis of twelve HIV-1 viral isolates, and it maps in the same genomic region as the env gene straddling the gp120/gp41 junction (Fig. 1). The product of this gene – the antisense protein (ASP) – is a polypeptide of ~ 190 residues with high content of hydrophobic amino acids, which suggests an association with cellular membranes [32]. While the original study did not include experimental evidence that the asp ORF encodes an actual protein, it provided clues in support of that conclusion. First was the evidence that the ORF is longer than 100 codons, which is uncommon in DNA strands complementary to known genes [33]. Second, the sequences analyzed showed the presence of signals necessary for production of an antisense mRNA transcript, such as the promoter, poly-A addition signal and site, and the downstream G and T domains. Finally, conserved sequences necessary for protein translation were also detected, including canonical start and stop codons, and a codon periodicity of ‘G-nonG-N’ [32].

Fig. 1
figure 1

The HIV-1 antisense gene, asp. The figure shows a schematic representation of the HIV-1 proviral genome with structural and enzymatic genes (gag, pol, env), regulatory genes (tat and rev), and accessory genes (vif, vpr, vpu, nef) expressed from the proviral 5’ LTR. The antisense gene asp is expressed from a negative sense promoter in the U3 region of the 3’ LTR in a manner independent of the viral transactivator, Tat. The negative sense promoter contains binding sites for USF, Ets-1, LEF-1, Sp1 and NF-κB. The antisense transcript Ast is a bifunctional RNA with both noncoding and protein-coding activities. The former is carried out in the nucleus: Ast acts as a lncRNA that promotes epigenetic silencing of HIV-1 by recruiting the histone methyltransferase (PRC2) to the 5’ LTR leading to trimethylation of lysine 27 on histone H3 (H3K27me3), which leads to assembly of the nucleosome Nuc-1, and inhibition of transcription. In addition, Ast is translocated to the cytoplasm where it functions as a mRNA and leads to the expression of the antisense protein ASP. In non-productively infected cells, ASP accumulates in the nucleus, whereas in productively infected cells ASP localizes in the cytoplasm and on the cell membrane in close proximity of the ENV. Further, upon viral budding and release, ASP is also detectable on the viral envelope

Experimental evidence of antisense transcription in the HIV-1 genome first came in 1990 through the use of Northern blot analysis of poly-A + RNA extracted from acutely infected cells [34] (Table 1). Subsequently, the use of RT-PCR allowed to prove HIV-1 antisense transcription in chronically infected T- and myeloid-derived cell lines, and also in clinical samples from early-stage, asymptomatic people living with HIV-1 (PLWH) [25, 35]. Further, the introduction of strand-specific RT-qPCR assays able to avoid artifacts due to endogenous and/or self-priming provided stronger evidence of antisense transcription in the HIV-1 proviral genome [36,37,38]. Antisense transcription was also detected in studies employing high-throughput sequencing methods [39].

Table 1 Antisense transcription activity in endogenous and exogenous retroviruses

The structure of the HIV-1 antisense transcripts and the mechanisms that regulate the expression were the focus of several reports. The earliest identified an antisense RNA of 2242 nt that originated in the R region of the 3’ LTR and terminated in a poly-A tract [25]. These results were confirmed by a later report [40]. However, Landry et al. identified multiple transcription start sites in the U3 region as well as in the nef and env genes, and also a polyadenylation signal in the pol gene [36]. A more in-depth analysis by Kobayashi-Ishihara and colleagues described a major antisense transcript (ASP-L or Ast) of 2574 nt with start site in the U3 region of the 3’LTR and a termination site in the env gene (Fig. 1) [37]. Interestingly, the same study demonstrated that a large fraction of HIV-1 antisense transcripts has a predominantly nuclear localization [37].

The location of the 5’ terminus of the antisense transcripts suggested that their expression is directed by a negative sense promoter (NSP) within the 3’LTR [25], which was shown to have 3- to 9-fold lower activity than that of the HIV-1 positive sense promoter (PSP), and it was inhibited by Tat expression, possibly by directing the transcriptional machinery to the PSP [25, 27]. The report by Michael et al. showed that NSP is a TATA-less promoter, and that the NF-κB and USF binding sites are critical for its activity [25], and a subsequent report identified an Sp1 binding site that is essential for NSP function [41]. Bentley et al. described regions of the 3’LTR with moderate, profound, and variable impact on NSP activity [27]. The segment of the 3’LTR with profound impact on NSP activity was mapped in the U3 region, and it contains binding sites for Sp1, NF-κB, LEF-1, Ets-1, and USF (Fig. 1) [27]. Disruption of the TATA box in the positive strand of the U3 region in the 3’ LTR increased NSP activity, which supports the notion that NSP is a TATA-less promoter and suggests that antisense transcription is under the control of an initiator element (InR) [27, 37]. Indeed, two putative InRs were later identified within the U3 region of the 3’ LTR [42], and a third one within in the R region (Fig. 1) [40]. A recent study from the Matsuoka group showed that HIV-1 antisense transcripts are inefficiently polyadenylated, which promotes their nuclear retention [43]. Our group investigated additional mechanisms involved in regulating the expression and possibly the function of HIV-1 antisense transcripts. First, we reported that the activity of the NSP within the HIV-1 3’ LTR is under epigenetic regulation [44]. In particular, we found the presence of a nucleosome over the U3 region and in close proximity of the nef-3’ LTR boundary (Fig. 1). Assembly and disassembly of this nucleosome is under the control of epigenetic modifications of lysine 9 and 27 on histone H3: acetylation of these residues increases transcriptional activity of NSP and promotes antisense transcription, whereas di- and trimethylation of these residues has the opposite effects [44]. In addition, we identified and precisely mapped post-transcriptional (epigenetic) base modifications deposited on HIV-1 antisense transcripts, which include primarily ribose methylation at multiple adenosine and guanosine residues, and pseudouridylation [45]. Studies are underway that seek to address whether dynamic addition/removal of these modification contributes to regulate the stability, sub-cellular localization, interaction with binding partners, and functional activity of HIV-1 antisense transcripts.

In line with that, studies from several groups including our own have shown that HIV-1 antisense transcripts act as bifunctional RNAs (Fig. 1). In addition to serving as mRNA for the expression of the HIV-1 antisense protein ASP, these transcripts also function as noncoding RNAs that regulate the expression of HIV-1 sense transcripts. An early report showed HIV-1 antisense transcripts reduce the expression HIV-1 Gag RNA, the levels of HIV-1 proviral DNA, and viral production in the culture supernatant [37]. Subsequently, Kevin Morris’ group provided evidence that HIV-1 antisense transcripts promote HIV-1 latency via epigenetic silencing of HIV-1 transcription [46]. In that report, Saayman et al. showed that knockdown of HIV-1 antisense transcripts resulted in a reduction in suppressive epigenetic marks (H3K9me2 and H3K27me3) at the 5’LTR, and they also demonstrated that HIV-1 antisense transcripts interact with the DNA methyltransferases [46]. Our group reported that ectopic overexpression of HIV-1 antisense transcripts lacking protein-coding capacity suppressed basal HIV-1 transcription during latency, inhibited latency reversal, and accelerated re-establishment of latency [38]. In addition, overexpression of HIV-1 antisense RNA maintained high levels of PRC2 and the suppressive epigenetic mark H3K27me3 at the HIV-1 5’LTR even after treatment with LRAs. We also showed that these effects involve interaction with subunits of the epigenetic silencer PRC2 (Fig. 1) [38]. A more in-depth discussion can be found in a recent review by Li et al. [47].

Sequence analyses and computer modeling suggest that the protein encoded by HIV-1 antisense transcripts (ASP) includes intracellular N- and C-termini, two transmembrane domains and an intervening extracellular loop [48, 49]. Additional features of interest are two closely spaced cysteine triplets in the N-terminal portion of the protein and a highly conserved PxxPxxP motif located between residues 40–50 of the protein (Fig. 1). Thorough and systematic experimental analyses are needed to assess the functional role of these and possibly other yet unidentified ASP domains [50]. A very recent study utilized molecular modeling and dynamics simulation to predict the 3D-structures of ASP [51]. The in silico analyses described in that study identified three possible functional domains in ASP, namely the Von Willebrand Factor Domain-A (VWFA), the Integrin subunit alpha-X (ITGSX), and the ETV6-Transcriptional repressor [51]. Wet lab molecular studies are required to confirm the validity of these findings, and to ascertain the role these domains play in the mechanisms of action of ASP. Despite being first identified more than 30 years ago, the role of ASP in the virus lifecycle remains largely a mystery, which is in part due to its low expression levels. Additionally, the hydrophobic properties of ASP make it exceptionally challenging to raise antibodies able to reliably detect it. Nevertheless, expression of ASP has been demonstrated in several cell systems. An early study used electron microscopy to show that ASP associates with plasma, mitochondrial and nuclear membranes [139]. A later report found that ASP localizes to the plasma membrane (with both polarized and unpolarized expression patterns) as well as with cell surface protrusion [49]. More recently, we used flow cytometry and confocal microscopy to study ASP expression in several chronically infected lymphoid and myeloid cell lines and also in acutely infected primary human CD4 + T cells and monocyte-derived macrophages (MDM) [48]. Using a mouse monoclonal antibody against an epitope mapping in the extracellular loop of ASP, we found that ASP displays a polarized nuclear distribution in unstimulated, non-productively infected lymphoid and myeloid cell lines [48]. After cell stimulation and reactivation of productive infection, ASP was transported into the cytoplasm and to the plasma membrane where it colocalizes with the HIV-1 envelope glycoprotein ENV (Fig. 1). We also showed that upon budding and release from infected cells, ASP is present on the viral envelope (Fig. 1) [48].

The role of ASP in viral replication has been explored to a much lesser degree. Clerc and colleagues generated HIV-1 molecular clones with a premature stop codon in the asp ORF, but they were not able to show any appreciable difference in viral replication compared to the wildtype clone [49]. Two reports by the Barbeau group showed that ASP expression can induce autophagy in infected cells [52, 53]. Many viruses utilize autophagy to their advantage during their replication cycle [54]. In the case of HIV-1, autophagy is required for GAG processing during infection of macrophages, and it significantly increases viral production [55]. Therefore, these findings provide at least one possible role of ASP in HIV-1 replication.

Antisense transcription has also been shown to occur in other lentiviruses, including feline immunodeficiency virus (FIV) and bovine immunodeficiency virus (BIV) [56, 57] (Table 1). However, these and other lentiviruses do not express antisense proteins. Indeed, while a full-length asp ORF is present in the majority of HIV-1 strains of group M (responsible for the pandemic), it is absent in HIV-1 strains that belong to non-pandemic groups N, O, and P [58, 59]. Moreover, the frequency of a full-length asp ORF among strains of each group-M HIV-1 clade correlates with the prevalence of the clade [59]. The asp ORF is also absent in HIV-2 as well as in all species of lentiviruses infecting non-human primates (SIV infecting Cercopithecidae and African apes) [58, 59]. While many of these viral strains contain the canonical start and stop codons of the asp ORF, they also display one or more premature stops that prevent expression of a full-length ASP.

This suggests that the asp ORF originated recently, when group M diverged from its most immediate progenitor, SIVcpz_Ptt. This is supported by evidence we reported recently showing that the average number of internal stop codons in the genomic region of the asp ORF steadily declines in viral strains that are progressively closer to HIV-1 group M [58]. It also suggests the possibility that the creation and conservation of a full-length may facilitate viral spread or pathogenesis. On the other hand, this leaves open the question about the role that antisense transcription plays in the context of lentiviruses that do not express an antisense protein.

HTLV-1 and other deltaretroviruses. Although evidence supporting the presence of antisense genes in the HIV-1 and HTLV-1 genomes came around the same time [23, 25, 34], much more progress has been made in understanding the role of the HTLV-1 antisense gene. Sense transcription of the HTLV-1 proviral genome is driven by the 5’ LTR. The HTLV-1 Tax transactivator induces sense transcription by interacting with phosphorylated CREB-2 when bound to Tax Response Elements (TRE) in the 5’ LTR (Fig. 2), and by recruiting the CBP/p300 coactivator [60]. Antisense transcription of HTLV-1 was shown to be driven by a TATA-less negative sense promoter in the proviral 3’ LTR that relies on Sp1 binding sites [23, 26, 61]. In addition to Sp1, the transcription factor MEF-2 has also been shown to promote HTLV-1 antisense transcription [62], which was found to start from multiple positions in the R and U5 regions of the 3’ LTR (Fig. 2) [63]. As reported for HIV-1, antisense transcripts of HTLV-1 show primarily nuclear distribution due to inefficient polyadenylation [43].

Fig. 2
figure 2

The HTLV-1 antisense gene, hbz. The figure shows a schematic representation of the HTLV-1 proviral genome with structural and enzymatic genes (gag, pol/pro, env), regulatory genes (tax and rex), and accessory genes (p12, p13, p30) expressed from the proviral 5’ LTR. The antisense gene hbz is expressed from the proviral 3’ LTRs in a manner independent of the viral transactivator, Tax. The hbz promoter contains binding sites for Sp1 and MEF-2. HTLV-1 produces spliced (s-hbz) and unspliced (us-hbz) antisense RNA variants, which encode two forms of the HBZ protein. Hbz transcripts have both noncoding and protein-coding activities. They act as lncRNAs that increase expression of Survivin and E2F1, which promote cell survival and proliferation, respectively. In addition, Hbz transcripts act as mRNAs that encode for the HBZ protein. HBZ interacts with CREB-2, it prevents the formation of the CREB-2/Tax heterodimer and binding to TRE sites in the HTLV-1 5’ LTR, and ultimately it downregulates HTLV-1 expression

Demonstration that HTLV-1 encodes an antisense protein came in 2002 in a seminal paper by Jean Michel Mesnard and colleagues [64]. In their search for novel CREB-2 binding partners, the authors screened a cDNA library from the HTLV-1 infected MT2 cell line and isolated the cDNA of a previously unknown protein encoded in the negative strand of HTLV-1. Sequence analyses showed that the cDNA encodes a 31-kDa protein with features typical of bZIP transcription factors, hence its name: HTLV-1 bZIP factor, or HBZ [64]. Subsequent reports demonstrated that HTLV-1 also expresses both spliced and unspliced forms of the Hbz RNA (Fig. 2) [65,66,67], which differ in their 5′ untranslated regions and in the 5′ region of coding sequences resulting in slightly different N-terminal sequence of the two protein products (for a comprehensive review, see [68]). The protein product of the spliced Hbz RNA is expressed at higher levels than the product of the unspliced mRNA, therefore most studies have focused on the former (Table 1).

In their seminal report, Gaudray and colleagues showed that HBZ displays nuclear localization, that it interacts with CREB-2, and that it prevents binding of CREB-2 and Tax to TRE sites in the HTLV-1 5’ LTR, which results in downregulation of HTLV-1 expression (Fig. 2) [64]. Interaction of HBZ with CREB-2 and other members of the AP1 superfamily including c-Jun, JunB, JunD, CREB, MafB, and ATF3 was shown to require the bZIP domain in the C-terminus of HBZ [64, 69,70,71,72,73,74,75]. The N-terminal portion of the protein contains an activation domain that recruits the transcriptional coactivator CBP/p300 [64]. In addition to downregulating HTLV-1 expression, HBZ was shown to have a number of other properties. HBZ contributes to HTLV-1 latency by inhibiting the effects of Tax on activation of the canonical NF-κB pathway, which causes cell senescence [76, 77], and also by reducing Rex-mediated nuclear export of HTLV-1 unspliced or partially spliced transcripts [78]. HBZ has also been shown to promote cell proliferation via its interaction with members of the AP1 superfamily, ATF3 and JunD [69, 75]. The effects of HBZ on cell proliferation also involve its ability to induce the noncanonical Wtn ligand Wtn5 [79], and to upregulate expression of brain-derived neurotropic factor (BDNF) that promotes proliferation of leukemia cells [80]. HBZ reduces apoptosis and autophagy by inhibiting the expression of the pro-apoptotic molecule, Bim [81] and by inducing the mTOR pathway [82]. Additionally, HBZ affects genomic integrity [83] and telomere length [84], and it promotes inflammation [85,86,87]. Therefore, HBZ contributes to HTLV-1 persistence through multiple mechanisms. The simian counterpart of HTLV-1 (STLV-1) was shown to express spliced sense and antisense transcripts that encode Tax and SBZ (STLV-1 bZIP) proteins with functions very similar to the HTLV-1 factors [88] (Table 1).

As observed in the case of HIV-1, the antisense transcript of HTLV-1 is also bifunctional and has activities that are independent of its protein-coding role. Indeed, the HBZ RNA was shown to promote cell proliferation by increasing expression of E2F1 [66]. Moreover, the HTLV-1 antisense transcript has anti-apoptotic activities that involve increased expression of Survivin (Fig. 2) [89].

Among other members of the Deltaretrovirus genus, the human T-lymphotropic viruses 2, 3, and 4 (HTLV-2, 3, and 4) express antisense proteins called APH-2, 3, and 4 [90,91,92] (Table 1), although their role in viral replication is not as well characterized. HTLV-2 expresses a spliced antisense transcript from several start sites in the 3’ LTR, which encodes APH-2 [63, 93]. This protein lacks the bZIP domain found in HBZ, but can interact with the Tax protein of HTLV-2 (Tax2) and with CREB [94]. APH-2 shows nuclear, but not nucleolar, distribution. Similarly, HTLV-3 and 4 express spliced antisense transcripts that encode the APH-3 and APH-4 proteins, respectively [91]. Both of them lack a bZIP domain but suppress 5’ LTR-driven transcription, and they show different subcellular localizations: APH-3 is found in both nucleus and cytoplasm, whereas APH-4 only in the nucleus [91]. Again, the simian counterparts of HTLV-2, 3 and 4 (called STLV-2, 3, and 4) also encode antisense proteins [95].

Finally, recent studies have shown that the Bovine Leukemia Virus (BLV) expresses two antisense transcripts (AS1 and AS2) both in leukemic and in asymptomatic samples [96] (Table 1). Expression of these transcripts is under the control of IRF binding sites and E-box in the 3’ LTR [96]. Both transcripts initiate in the U5 region of the 3’ LTR and are spliced. AS1 undergoes alternative polyadenylation and generates two forms: a short one (AS1-S) of ~ 600nt, and a long one (AS1-L) of ~ 2200 nt. The AS2 transcript is present in a single form of ~ 400 nt. While AS1 and AS2 include a 264-nt ORF, their actual protein coding potential is unlikely based on analysis using Coding Potential Assessment Tools and also due to their nuclear localization, which suggests a predominantly regulatory role [96]. Moreover, ablation of BLV noncoding antisense transcripts has been shown to reduce proliferation of infected cells, suggesting that these molecules play a role in the leukemic process [97].

In summary, Deltaretroviruses – just like Lentiviruses – express antisense transcripts from the 3’ LTR. In the case of HTLVs, antisense transcripts are bifunctional molecules with both protein-coding and noncoding roles. However, in the case of BLV, the antisense transcripts have only regulatory, noncoding activity.

Gammaretroviruses. To date, the expression of antisense transcripts or proteins has not been reported for any members of this genus of simple retroviruses. However, two studies have shown that the LTR of the murine leukemia virus (MLV) is capable of driving bidirectional transcription (Table 1). Studies by the Pedersen group showed that in T- and B-cell lymphomas, the MLV 5’ LTR can direct antisense transcription from multiple start sites within the U3 region [28, 98]. These 5’ LTR-driven transcripts extend into neighboring host genome and produce chimeric RNA molecules expressing proto-oncogenes such as Bach2 and Jdp2, which cause cell transformation. The authors confirmed these results using a knock-in mouse model homozygous for a single copy of the MLV LTR inserted in a genomic region that activates Nras in B-cell lymphomas [28]. Subsequently, Arpin-André et al. reported molecular studies that confirmed the ability of the MLV LTR to drive bidirectional transcription [26]. These studies also showed that in the absence of their respective viral transactivators, the LTRs of HIV-1 and HTLV-1 showed much weaker sense transcription than the LTR of MLV (which does not encode a transactivator). However, the three retroviral LTRs displayed similar antisense transcription activity [26].

These studies demonstrate that, although simple retroviruses have not been shown to express antisense transcripts or proteins, their LTRs do have properties of bidirectional promoters, and therefore are potentially capable of driving antisense transcription.

Antisense transcription in endogenous retroviruses

As discussed above, vertebrate genomes contain a high proportion of transposable elements, many of which have retroviral origin [3, 99]. Indeed, sequencing of the human genome has revealed that ~ 8% is occupied by various forms of endogenous retroviruses (ERV), including full-length viral genomes, partial genomes, and as many as 25,000 isolated (solo) LTRs [6, 100]. These are sequences that originated from integration of now-extinct exogenous retroviruses over millions of years [2, 3]. Ancestral retroviruses may have been capable of infecting cells of the germ line, which ensured transmission to the offspring. The mechanism through which ERV formed multiple integration sites is similar to the one used by exogenous retroviruses, but it is confined to the same cell. It entails expression of the genomic RNA, reverse transcription into dsDNA, and then integration into a different site of the host genome.

To date, there is no evidence that endogenous retroviruses encode antisense proteins in the proviral minus strand in a manner similar to primate T-lymphotropic viruses (bZIP proteins) and group M HIV-1 strains (ASP). However, multiple studies have demonstrated that ERVs, much like exogenous retroviruses, express antisense transcripts (Table 1). Analysis of HERV-K subtype HML-2 expression in prostate cancer cells revealed that this ERV is capable of both sense and antisense transcription [101]. The same report also provided evidence that solo LTRs are transcribed in both directions [101]. A similar study focusing on a model of mammary epithelial cell transformation detected expression of HERV-K (HML-2) in both directions with the majority of transcription being in the antisense orientation [102]. A more recent study showed that differentiation of monocytes into macrophages in response to ionizing radiation is associated with reactivation of > 600 retroelements, especially several clades of HERVs [103]. In the case of HERVs, this involved induction of both sense and antisense transcription, with the latter expressed at 3- to 5-fold higher levels than the former [103]. Expression of HERVs has also been investigated in the context of HIV-1 infection [104, 105]. These studies found the presence of HERV-K (HML-2) RNA and proteins in plasma of PLWH, and some of the RNA sequences derived from newly discovered proviruses [105]. Also, these RNA molecules had been expressed from intact proviruses as well as proviruses lacking the 5’ LTR, suggesting that they originated from antisense transcription [105].

There is also ample evidence that ERV LTRs are able to drive bidirectional transcription from studies that investigated the expression of ERV-host chimeric transcripts (for a comprehensive review see [106]. For instance, in the Down Syndrome Critical Region (DSCR), the HERV1 LTR acts as a bidirectional promoter leading to the expression of DSCR4 and DSCR8 transcripts [30]. In another example, in K562 cells the bidirectionally active HERV9 LTR has been shown to direct the expression of an isoform of the CADM2 transcript that is shorter than in other cell types [106]. Evidence of bidirectional LTR activity has also been shown in mice [107]. The murine ERV-L (MuERV-L) and intracisternal A particle (IAP) are activated in the preimplantation mouse embryo, and sense and antisense transcripts are expressed at the 2- and 8-cell stage embryo [107]. Interestingly, host-mediated RNA interference limits the expression of these two retroelements as a way to protect genomic integrity at this early stage of development [107].

Therefore, the LTRs of endogenous retroviruses function as bidirectional promoters and direct antisense transcription despite the absence of antisense open reading frames and even in case of partial or complete deletion of the proviral genome (solo LTR). This ability is in line with what has been discussed above in regard to exogenous retroviruses, and it appears to be a feature of all retroelements.

Origin of bifunctional transcripts

Bifunctional RNAs (also known as “coding-noncoding RNAs”, or cncRNAs) are defined as transcripts that have both protein-coding and noncoding (regulatory) activities. They have been described in all living organisms, including bacteria (RNAIII in S. aureus, SR1 in B. subtilis, and SgrS in E. coli), plants (ENOD40 in M. trunculata), insects (oskar in D. melanogaster), and vertebrates (squint in D. rerio, vegt in X. levis, Ube3a in R. norvegicus, Sra1 in M. musculus) [108, 109]. Humans are no exception, and many examples of bifunctional transcripts have been also described in Homo sapiens [110].

Studies conducted over the last few years have shown that ~ 40% of transcripts considered to be “true” long noncoding RNAs (functional transcripts without protein coding activities) are associated with ribosomes in the cytoplasm and may be translated [111,112,113]. Some lncRNAs have actually been shown to contain short open reading frames that are translated into peptides with biological roles. For instance, the lncRNAs LINC00948, LINC00116, and LOC100507537 express the micropeptides Myoregulin (46 aa), Mitoregulin (56 aa) and Dwarf (34 aa) [114,115,116]. A more recent study identified 129 additional examples of lncRNAs that contain small ORF translated into short polypeptides with biological function [117]. At the other end of the spectrum, we find transcripts with known protein-coding function and that also have regulatory roles. Two examples are the mRNAs SRA, which interacts with and activates the transcription factor, MyoD [118], and p53, which interacts with and inactivates the E3 ubiquitin-protein ligase mdm2 [119].

The antisense transcripts HBZ of HTLV-1 and Ast of HIV-1 have both protein-coding and non-coding activities [50, 68]. Did these two bifunctional transcripts originate as lncRNAs that later developed protein-coding capacity or, vice versa, as mRNAs that have acquired regulatory activity? Which is the ancestral and which is the acquired function? While this question remains to be addressed conclusively, two lines of evidence may indicate a possible answer.

First, as discussed above, our current understanding indicates that antisense transcription is a common feature among retroviruses – both endogenous and exogenous, and simple and complex (Table 1). In many cases these antisense transcripts are not translated into a protein product. We also have evidence that retroviral LTRs have an intrinsic ability to direct bidirectional transcription even in the case of ERVs that lack antisense ORFs or in the case of solo LTRs.

Second, a recent study from the Matsuoka group has demonstrated that even antisense transcripts with protein-coding capacity, such as HBZ and Ast, display a predominantly nuclear distribution due to inefficient polyadenylation [43]. In their study, the authors presented evidence that this is not the consequence of weak polyadenylation signals, but more likely of the weak native promoters in the 3’ LTRs [43].

Thus, the evidence that expression of antisense transcripts is phylogenetically more widespread among retroviruses than expression of antisense proteins along with the predominantly nuclear distribution of bifunctional antisense transcripts that favors regulatory over protein-coding roles suggest that retroviral antisense transcripts may have originated as lncRNAs, which in some cases have acquired the ability to function as mRNAs. Confirmation of this hypothesis requires further studies, including a direct demonstration that non-pandemic HIV-1 strains and SIV strains – which do not encode antisense proteins – are capable of antisense transcription.

Regulatory activity of natural antisense transcripts

Antisense transcripts can modulate the expression of their cognate protein-coding sense genes [120]. Most often, antisense transcripts inhibit expression of the sense gene, but cases have been reported where they increase expression of the sense transcript by protecting it from nuclease degradation [121]. The regulatory function of antisense transcripts is mediated by the RNA molecule itself and/or by the act of antisense transcription itself. This can occur at various, non-mutually exclusive steps during the expression of the sense RNA molecule: initiation, processing, transport, stability, and translation [122]. For comprehensive reviews, see [47, 123,124,125].

Antisense RNAs inhibit transcription of their cognate sense gene through at least four mechanisms (collectively, transcriptional interference or TI) [125]:

  • promoter competition: sense and antisense RNA are expressed from a bidirectional promoter. Assembly of the transcriptional machinery expressing the antisense RNA blocks or prevents expression of the sense RNA.

  • binding site occlusion: the RNA polymerase complex expressing the antisense transcript prevents chromatin access to transcription factors required for expression of the sense transcript.

  • RNA polymerase collision: the antisense transcriptional machinery displaces the machinery assembled onto the promoter of the sense transcript’s (‘sitting duck’), or it stalls the incoming sense transcriptional machinery (‘roadblock’).

  • epigenetic silencing: the antisense transcript recruits chromatin remodeling factors at the promoter region of the sense transcript, leading to a closed chromatin status and transcriptional repression. We showed this to be one of the mechanisms through which the HIV-1 antisense transcript Ast promotes silencing of the proviral 5’ LTR and viral latency [38].

Antisense transcripts can also regulate the expression of their paired sense transcript post-transcriptionally through three mechanisms involving formation of double stranded RNA complexes:

  • RNA masking: formation of a sense-antisense duplex that blocks the interaction of the sense transcript with factors (proteins and miRNAs) that regulate its splicing, stability, transport, and translation [124, 125].

  • RNA interference: recognition of the RNA duplex by Dicer, with subsequent cleavage and formation of ‘endo-siRNAs’ [125]. Alternatively, the RNA duplex is recognized by Protein Kinase R (PKR), which undergoes dimerization and autophosphorylation, suppresses protein expression, and ultimately triggers IFNα/β innate immune responses.

  • RNA editing: recognition of duplex RNA by members of the ADAR protein family that deaminate adenosine residues into inosine, which results in amino acid changes [124].

As discussed above, antisense transcription has been documented among many genera of endogenous and exogenous retroviruses even in the absence of protein-coding function (Table 1). What selective advantage does the noncoding function of antisense transcripts afford to the virus? What purpose does it serve in the virus life cycle?

Bimodal (ON-OFF) gene expression requires the establishment of thresholds that ensure stability of the system in the two states (bistability) and that regulate the switch between the two states. In protein-based regulatory systems, threshold generation is achieved through cooperativity [126,127,128]. For example, in the case of HIV-1, the choice between ON (viral expression) and OFF states (viral latency) is thought to be regulated entirely by the HIV-1 transactivator Tat through a mechanism of positive-feedback loop. However, since Tat is a monomeric transactivator that does not bind the TAR sequence cooperatively, classical models would predict an instability of the OFF state (viral latency) [128]. Moreover, Tat is not sequestered or neutralized by natural inhibitors, and thus it is not possible to invoke a threshold-linear mechanism of transcriptional silencing to achieve and stabilize viral latency [129, 130]. Given this evidence, how does HIV-1 achieve stable proviral latency?

Ultrasensitivity is another mechanism of threshold generation that governs the switch between two alternative transcription states [131,132,133]. It stabilizes the OFF state by ignoring weak stimuli, but it allows a rapid, sigmoid-like switch to the ON state in response to strong signals [133]. Therefore, in systems where the protein transactivator lacks cooperativity or inhibitors (such as the HIV-1 transactivator Tat), ultrasensitive responses contribute to achieve bistability.

A number of reports have shown that some antisense transcripts can establish an ultrasensitive ON-OFF switch regulating the expression of their cognate sense transcript (Fig. 3) [123]. When antisense transcription precedes that of the sense transcript, the antisense RNA buffers weak stimuli (transient, spurious signals) and sets a threshold dampening stochastic variations in the expression of sense RNA [134, 135]. Moreover, repression of sense transcription is more robust in the case of antisense transcripts that exert their regulatory role both transcriptionally and post-transcriptionally [136]. However, when activating stimuli are sufficiently strong to overcome the threshold set by the antisense transcript, then sense transcription is switched on reaching maximal levels very rapidly (sigmoid curve). In the ON state, the continued presence of antisense RNAs can lead to transcriptional bursting of the sense gene and increase cell-to-cell variability [137]. Finally, when the activating stimuli decline and cease, antisense RNAs contribute to a faster return to baseline levels of the sense RNA [137]. Therefore, antisense transcripts establish ultrasensitive thresholds that dampen stochastic expression of the sense gene in conditions of sub-optimal signals, allow rapid expression of the sense gene when activating stimuli are sufficiently strong, and promote a faster return to baseline levels when activating stimuli subside. It also worth noting that these thresholds are established independently for each locus or cell, which increases variability and adaptation [138].

Fig. 3
figure 3

Dynamics of ON/OFF sense transcription in the absence (left) and presence (right) of cognate regulatory antisense transcript. Left panel (in the absence of regulatory antisense RNA): at low levels of activating stimuli, OFF state of sense transcription is unstable and noisy, and it is induced stochastically in response to transient fluctuations of the signal. As activating stimuli strengthen, sense transcription slowly increases and gradually reaches maximum levels along a hyperbolic curve (ON). Similarly, return to baseline levels of sense transcription occurs gradually with the waning of the activating stimuli. Right panel (in the presence of regulatory antisense RNA): at low levels of activating stimuli (OFF), the antisense transcript buffers spurious activating stimuli, thus dampening noisy expression of sense transcripts and achieving a stable OFF state. When activating stimuli are sufficiently strong to overcome the threshold set by the antisense RNA, sense transcription is induced and rapidly reaches maximum levels along a sigmoid curve (ON). As activating stimuli weaken below the ultrasensitive threshold, expression of sense transcripts rapidly drops to baseline levels following an inverse sigmoid curve and returns to a stable OFF state

In the case of retroviruses, it is intriguing to speculate that antisense transcription might provide a mechanism to regulate sense gene expression through the establishment of ultrasensitive thresholds. This strategy allows the virus to stabilize the latency phase, and to control the switch to and from replication. Moreover, this can be achieved without the need to encode yet another open reading frame in small and densely packed viral genomes. Importantly, HIV-1 and HTLV-1 express antisense transcript relying primarily on ubiquitous cellular transcription factors independently of the viral transactivators, Tax and Tat [23,24,25, 27], which ensures constitutive expression. Studies on the non-protein coding function of HIV-1 antisense transcripts seem to support this interpretation, and they suggest that HIV-1 antisense transcripts suppress sense transcription via epigenetic silencing of the 5’ LTR [37, 38, 46].

This regulatory mechanism affords the virus a greater control over its destiny because it offers the ability to choose between replication and latency in response to the cell environment. Under conditions that are not ideal for production of viral progeny, expression of viral proteins would be a risk without benefit, because it would unnecessarily expose the infected cell to recognition and elimination by the immune system. At the same time, as we discussed above, positive-feedback loops where the viral transactivator (e.g., HIV-1 Tat) lacks cooperativity or natural inhibitors cannot ensure stability of viral latency (OFF state). Using antisense transcription as a means to establish an ultrasensitive threshold allows the virus to stabilize latency when the cell does not provide the ideal environment for viral replication, thus avoiding unnecessary immune recognition, ensuring viral persistence, and increasing the chances for spread.

Conclusions

Antisense transcription is a property of many endogenous and exogenous retroviruses. This is not only reflected in the expression of antisense RNA molecules, but also in the intrinsic ability of retroviral LTRs to direct bidirectional transcription. Moreover, retroviruses are capable of antisense transcription even in the absence of open reading frames encoded in the minus strand of the proviral genome. The broad phylogenetic distribution of antisense transcription across the Retroviridae family suggests important roles in the virus lifecycle that precede and go beyond protein-coding functions. Among other plausible roles, antisense transcripts may have evolved as a strategy to stabilize viral latency when the cell environment does not offer optimal conditions to sustain viral replication, while at the same time allowing rapid reversal of latency and maximal viral expression in response to strong activating stimuli.

Data availability

Not applicable.

Abbreviations

ASP:

antisense protein

BIV:

bovine immunodeficiency virus

BLV:

bovine leukemia virus

cncRNA:

coding-noncoding RNA

FIV:

feline immunodeficiency virus

SIV:

simian immunodeficiency virus

ERV:

endogenous retrovirus

HBZ:

HTLV-1 bZIP

HERV:

human endogeneous retrovirus

IAP:

intracisternal A particle

lncRNA:

long noncoding RNA

LTR:

Long Terminal Repeat

HIV-1:

Human Immunodeficiency Virus 1

HTLV-1/2/3/4:

Human T-lymphotropic Virus 1/2/3/4

MLV:

murine leukemia virus

ncRNA:

noncoding RNA

NSP:

negative sense promoter

PLWH:

people living with HIV-1

PSP:

positive sense promoter

STLV:

simian T-lymphotropic virus

References

  1. Coffin J, Blomberg J, Fan H, Gifford R, Hatziioannou T, Lindemann D, Mayer J, Stoye J, Tristem M, Johnson W et al. ICTV Virus Taxonomy Profile: Retroviridae 2021. J Gen Virol 2021, 102(12).

  2. Bannert N, Kurth R. The evolutionary dynamics of human endogenous retroviral families. Annu Rev Genomics Hum Genet. 2006;7:149–73.

    Article  CAS  PubMed  Google Scholar 

  3. Vargiu L, Rodriguez-Tome P, Sperber GO, Cadeddu M, Grandi N, Blikstad V, Tramontano E, Blomberg J. Classification and characterization of human endogenous retroviruses; mosaic forms are common. Retrovirology. 2016;13:7.

    Article  PubMed  PubMed Central  Google Scholar 

  4. Grandi N, Tramontano E. Human endogenous retroviruses are ancient acquired elements still shaping Innate Immune responses. Front Immunol. 2018;9:2039.

    Article  PubMed  PubMed Central  Google Scholar 

  5. Mayer J, Meese E. Human endogenous retroviruses in the primate lineage and their influence on host genomes. Cytogenet Genome Res. 2005;110(1–4):448–56.

    Article  CAS  PubMed  Google Scholar 

  6. Lander ES, Linton LM, Birren B, Nusbaum C, Zody MC, Baldwin J, Devon K, Dewar K, Doyle M, FitzHugh W, et al. Initial sequencing and analysis of the human genome. Nature. 2001;409(6822):860–921.

    Article  CAS  PubMed  Google Scholar 

  7. Magiorkinis G, Blanco-Melo D, Belshaw R. The decline of human endogenous retroviruses: extinction and survival. Retrovirology. 2015;12:8.

    Article  PubMed  PubMed Central  Google Scholar 

  8. Magin C, Lower R, Lower J. cORF and RcRE, the Rev/Rex and RRE/RxRE homologues of the human endogenous retrovirus family HTDV/HERV-K. J Virol. 1999;73(11):9496–507.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Armbruester V, Sauter M, Krautkraemer E, Meese E, Kleiman A, Best B, Roemer K, Mueller-Lantzsch N. A novel gene from the human endogenous retrovirus K expressed in transformed cells. Clin Cancer Res. 2002;8(6):1800–7.

    CAS  PubMed  Google Scholar 

  10. Polakowski N, Gregory H, Mesnard JM, Lemasson I. Expression of a protein involved in bone resorption, Dkk1, is activated by HTLV-1 bZIP factor through its activation domain. Retrovirology. 2010;7:61.

    Article  PubMed  PubMed Central  Google Scholar 

  11. Mbonye U, Karn J. The molecular basis for human immunodeficiency virus latency. Annu Rev Virol. 2017;4(1):261–85.

    Article  CAS  PubMed  Google Scholar 

  12. Schrom EM, Moschall R, Schuch A, Bodem J. Regulation of retroviral polyadenylation. Adv Virus Res. 2013;85:1–24.

    Article  CAS  PubMed  Google Scholar 

  13. Furger A, Monks J, Proudfoot NJ. The retroviruses human immunodeficiency virus type 1 and Moloney murine leukemia virus adopt radically different strategies to regulate promoter-proximal polyadenylation. J Virol. 2001;75(23):11735–46.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Ashe MP, Griffin P, James W, Proudfoot NJ. Poly(A) site selection in the HIV-1 provirus: inhibition of promoter-proximal polyadenylation by the downstream major splice donor site. Genes Dev. 1995;9(23):3008–25.

    Article  CAS  PubMed  Google Scholar 

  15. Valsamakis A, Zeichner S, Carswell S, Alwine JC. The human immunodeficiency virus type 1 polyadenylylation signal: a 3’ long terminal repeat element upstream of the AAUAAA necessary for efficient polyadenylylation. Proc Natl Acad Sci U S A. 1991;88(6):2108–12.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Gilmartin GM, Fleming ES, Oetjen J, Graveley BR. CPSF recognition of an HIV-1 mRNA 3’-processing enhancer: multiple sequence contacts involved in poly(A) site definition. Genes Dev. 1995;9(1):72–83.

    Article  CAS  PubMed  Google Scholar 

  17. Graveley BR, Gilmartin GM. A common mechanism for the enhancement of mRNA 3’ processing by U3 sequences in two distantly related lentiviruses. J Virol. 1996;70(3):1612–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Berkhout B, Klaver B, Das AT. A conserved hairpin structure predicted for the poly(A) signal of human and simian immunodeficiency viruses. Virology. 1995;207(1):276–81.

    Article  CAS  PubMed  Google Scholar 

  19. Das AT, Klaver B, Berkhout B. A hairpin structure in the R region of the human immunodeficiency virus type 1 RNA genome is instrumental in polyadenylation site selection. J Virol. 1999;73(1):81–91.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Klasens BI, Thiesen M, Virtanen A, Berkhout B. The ability of the HIV-1 AAUAAA signal to bind polyadenylation factors is controlled by local RNA structure. Nucleic Acids Res. 1999;27(2):446–54.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Ashe MP, Pearson LH, Proudfoot NJ. The HIV-1 5’ LTR poly(A) site is inactivated by U1 snRNP interaction with the downstream major splice donor site. EMBO J. 1997;16:5752–63.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Ashe MP, Furger A, Proudfoot NJ. Stem-loop 1 of the U1 snRNP plays a critical role in the suppression of HIV-1 polyadenylation. RNA. 2000;6(2):170–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Larocca D, Chao LA, Seto MH, Brunck TK. Human T-cell leukemia virus minus strand transcription in infected T-cells. Biochem Biophys Res Commun. 1989;163(2):1006–13.

    Article  CAS  PubMed  Google Scholar 

  24. Laverdure S, Polakowski N, Hoang K, Lemasson I. Permissive sense and antisense transcription from the 5’ and 3’ long terminal repeats of human T-Cell leukemia virus type 1. J Virol. 2016;90(7):3600–10.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Michael NL, Vahey MT, d’Arcy L, Ehrenberg PK, Mosca JD, Rappaport J, Redfield RR. Negative-strand RNA transcripts are produced in human immunodeficiency virus type 1-infected cells and patients by a novel promoter downregulated by Tat. J Virol. 1994;68(2):979–87.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Arpin-Andre C, Laverdure S, Barbeau B, Gross A, Mesnard JM. Construction of a reporter vector for analysis of bidirectional transcriptional activity of retrovirus LTR. Plasmid. 2014;74:45–51.

    Article  CAS  PubMed  Google Scholar 

  27. Bentley K, Deacon N, Sonza S, Zeichner S, Churchill M. Mutational analysis of the HIV-1 LTR as a promoter of negative sense transcription. Arch Virol. 2004;149(12):2277–94.

    Article  CAS  PubMed  Google Scholar 

  28. Rasmussen MH, Ballarin-Gonzalez B, Liu J, Lassen LB, Fuchtbauer A, Fuchtbauer EM, Nielsen AL, Pedersen FS. Antisense transcription in gammaretroviruses as a mechanism of insertional activation of host genes. J Virol. 2010;84(8):3780–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Feuchter A, Mager D. Functional heterogeneity of a large family of human LTR-like promoters and enhancers. Nucleic Acids Res. 1990;18(5):1261–70.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Dunn CA, Romanish MT, Gutierrez LE, van de Lagemaat LN, Mager DL. Transcription of two human genes from a bidirectional endogenous retrovirus promoter. Gene. 2006;366(2):335–42.

    Article  CAS  PubMed  Google Scholar 

  31. Leupin O, Attanasio C, Marguerat S, Tapernoux M, Antonarakis SE, Conrad B. Transcriptional activation by bidirectional RNA polymerase II elongation over a silent promoter. EMBO Rep. 2005;6(10):956–60.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Miller RH. Human immunodeficiency virus may encode a novel protein on the genomic DNA plus strand. Science. 1988;239(4846):1420–2.

    Article  CAS  PubMed  Google Scholar 

  33. Casino A, Cipollaro M, Guerrini AM, Mastrocinque G, Spena A, Scarlato V. Coding capacity of complementary DNA strands. Nucleic Acids Res. 1981;9(6):1499–518.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Bukrinsky MI, Etkin AF. Plus strand of the HIV provirus DNA is expressed at early stages of infection. AIDS Res Hum Retroviruses. 1990;6(4):425–6.

    Article  CAS  PubMed  Google Scholar 

  35. Vanhee-Brossollet C, Thoreau H, Serpente N, D’Auriol L, Levy JP, Vaquero C. A natural antisense RNA derived from the HIV-1 env gene encodes a protein which is recognized by circulating antibodies of HIV + individuals. Virology. 1995;206(1):196–202.

    Article  CAS  PubMed  Google Scholar 

  36. Landry S, Halin M, Lefort S, Audet B, Vaquero C, Mesnard JM, Barbeau B. Detection, characterization and regulation of antisense transcripts in HIV-1. Retrovirology. 2007;4:71.

    Article  PubMed  PubMed Central  Google Scholar 

  37. 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.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Zapata JC, Campilongo F, Barclay RA, DeMarino C, Iglesias-Ussel MD, 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.

    Article  CAS  PubMed  Google Scholar 

  39. Lefebvre G, Desfarges S, Uyttebroeck F, Munoz M, Beerenwinkel N, Rougemont J, Telenti A, Ciuffi A. Analysis of HIV-1 expression level and sense of transcription by high-throughput sequencing of the infected cell. J Virol. 2011;85(13):6205–11.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Ludwig LB, Ambrus JL Jr, Krawczyk KA, Sharma S, Brooks S, Hsiao CB, Schwartz SA. Human immunodeficiency virus-type 1 LTR DNA contains an intrinsic gene producing antisense RNA and protein products. Retrovirology. 2006;3:80.

    Article  PubMed  PubMed Central  Google Scholar 

  41. Peeters A, Lambert PF, Deacon NJ. A fourth Sp1 site in the human immunodeficiency virus type 1 long terminal repeat is essential for negative-sense transcription. J Virol. 1996;70(10):6665–72.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Kugel JF, Goodrich JA. Finding the start site: redefining the human initiator element. Genes Dev. 2017;31(1):1–2.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Ma G, Yasunaga JI, 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 U S A 2021, 118(17).

  44. Li R, Caico I, Xu Z, Iqbal MS, Romerio F. Epigenetic Regulation of HIV-1 Sense and Antisense Transcription in Response to Latency-Reversing Agents. Noncoding RNA 2023, 9(1).

  45. Estevez M, Li R, Paul B, Daneshvar K, Mullen AC, Romerio F, Addepalli B. Identification and mapping of post-transcriptional modifications on the HIV-1 antisense transcript Ast in human cells. RNA 2022.

  46. Saayman S, Ackley A, Turner AM, Famiglietti M, Bosque A, Clemson M, Planelles V, Morris KV. An HIV-encoded antisense long noncoding RNA epigenetically regulates viral transcription. Mol therapy: J Am Soc Gene Therapy. 2014;22(6):1164–75.

    Article  CAS  Google Scholar 

  47. Li R, Sklutuis R, Groebner JL, Romerio F. HIV-1 Natural Antisense Transcription and Its Role in Viral Persistence. Viruses 2021, 13(5).

  48. Affram Y, Zapata JC, Gholizadeh Z, Tolbert WD, Zhou W, Iglesias-Ussel MD, Pazgier M, Ray K, Latinovic OS, Romerio F. The HIV-1 Antisense Protein ASP Is a Transmembrane Protein of the Cell Surface and an Integral Protein of the Viral Envelope. J Virol 2019, 93(21).

  49. Clerc I, Laverdure S, Torresilla C, Landry S, Borel S, Vargas A, Arpin-Andre C, Gay B, Briant L, Gross A, et al. Polarized expression of the membrane ASP protein derived from HIV-1 antisense transcription in T cells. Retrovirology. 2011;8:74.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Gholizadeh Z, Iqbal MS, Li R, Romerio F. The HIV-1 Antisense Gene ASP: The New Kid on the Block. Vaccines (Basel) 2021, 9(5).

  51. Sathiyamani B, Daniel EA, Ansar S, Esakialraj BH, Hassan S, Revanasiddappa PD, Keshavamurthy A, Roy S, Vetrivel U, Hanna LE. Structural analysis and molecular dynamics simulation studies of HIV-1 antisense protein predict its potential role in HIV replication and pathogenesis. Front Microbiol. 2023;14:1152206.

    Article  PubMed  PubMed Central  Google Scholar 

  52. Torresilla C, Larocque E, Landry S, Halin M, Coulombe Y, Masson JY, Mesnard JM, Barbeau B. Detection of the HIV-1 minus-strand-encoded antisense protein and its association with autophagy. J Virol. 2013;87(9):5089–105.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Liu Z, Torresilla C, Xiao Y, Nguyen P, Caté C, Barbosa K, Rassart E, Cen S, Bourgault S, Barbeaub B. HIV-1 antisense protein of different clades induces Autophagy and Associates with the Autophagy factor p62. J Virol. 2019;93(2):e01757–01718.

    Article  PubMed  PubMed Central  Google Scholar 

  54. Dreux M, Chisari FV. Viruses and the autophagy machinery. Cell Cycle. 2010;9(7):1295–307.

    Article  CAS  PubMed  Google Scholar 

  55. Kyei GB, Dinkins C, Davis AS, Roberts E, Singh SB, Dong C, Wu L, Kominami E, Ueno T, Yamamoto A, et al. Autophagy pathway intersects with HIV-1 biosynthesis and regulates viral yields in macrophages. J Cell Biol. 2009;186(2):255–68.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. Briquet S, Richardson J, Vanhee-Brossollet C, Vaquero C. Natural antisense transcripts are detected in different cell lines and tissues of cats infected with feline immunodeficiency virus. Gene. 2001;267(2):157–64.

    Article  CAS  PubMed  Google Scholar 

  57. Liu B, Zhao X, Shen W, Kong X. Evidence for the antisense transcription in the proviral R29-127 strain of bovine immunodeficiency virus. Virol Sin. 2015;30(3):224–7.

    Article  PubMed  PubMed Central  Google Scholar 

  58. Pavesi A, Romerio F. Extending the Coding Potential of Viral Genomes with Overlapping Antisense ORFs: A Case for the De Novo Creation of the Gene Encoding the Antisense Protein ASP of HIV-1. Viruses 2022, 14(1).

  59. Cassan E, Arigon-Chifolleau AM, Mesnard JM, Gross A, Gascuel O. Concomitant emergence of the antisense protein gene of HIV-1 and of the pandemic. Proc Natl Acad Sci U S A. 2016;113(41):11537–42.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. Bangham CRM, Miura M, Kulkarni A, Matsuoka M. Regulation of latency in the human T cell leukemia virus, HTLV-1. Annu Rev Virol. 2019;6(1):365–85.

    Article  CAS  PubMed  Google Scholar 

  61. Yoshida M, Satou Y, Yasunaga J, Fujisawa J, Matsuoka M. Transcriptional control of spliced and unspliced human T-cell leukemia virus type 1 bZIP factor (HBZ) gene. J Virol. 2008;82(19):9359–68.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  62. Jain P, Lavorgna A, Sehgal M, Gao L, Ginwala R, Sagar D, Harhaj EW, Khan ZK. Myocyte enhancer factor (MEF)-2 plays essential roles in T-cell transformation associated with HTLV-1 infection by stabilizing complex between tax and CREB. Retrovirology. 2015;12:23.

    Article  PubMed  PubMed Central  Google Scholar 

  63. Lin E, Panfil AR, Sandel G, Jain P. Novel perspectives on antisense transcription in HIV-1, HTLV-1, and HTLV-2. Front Microbiol. 2022;13:1042761.

    Article  PubMed  PubMed Central  Google Scholar 

  64. Gaudray G, Gachon F, Basbous J, Biard-Piechaczyk M, Devaux C, Mesnard JM. The complementary strand of the human T-cell leukemia virus type 1 RNA genome encodes a bZIP transcription factor that down-regulates viral transcription. J Virol. 2002;76(24):12813–22.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  65. Cavanagh MH, Landry S, Audet B, Arpin-Andre C, Hivin P, Pare ME, Thete J, Wattel E, Marriott SJ, Mesnard JM, et al. HTLV-I antisense transcripts initiating in the 3’LTR are alternatively spliced and polyadenylated. Retrovirology. 2006;3:15.

    Article  PubMed  PubMed Central  Google Scholar 

  66. Satou Y, Yasunaga J, Yoshida M, Matsuoka M. HTLV-I basic leucine zipper factor gene mRNA supports proliferation of adult T cell leukemia cells. Proc Natl Acad Sci U S A. 2006;103(3):720–5.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  67. Murata K, Hayashibara T, Sugahara K, Uemura A, Yamaguchi T, Harasawa H, Hasegawa H, Tsuruda K, Okazaki T, Koji T, et al. A novel alternative splicing isoform of human T-cell leukemia virus type 1 bZIP factor (HBZ-SI) targets distinct subnuclear localization. J Virol. 2006;80(5):2495–505.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. Ma G, Yasunaga J, Matsuoka M. Multifaceted functions and roles of HBZ in HTLV-1 pathogenesis. Retrovirology. 2016;13:16.

    Article  PubMed  PubMed Central  Google Scholar 

  69. Gazon H, Lemasson I, Polakowski N, Cesaire R, Matsuoka M, Barbeau B, Mesnard JM, Peloponese JM Jr. Human T-cell leukemia virus type 1 (HTLV-1) bZIP factor requires cellular transcription factor JunD to upregulate HTLV-1 antisense transcription from the 3’ long terminal repeat. J Virol. 2012;86(17):9070–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  70. Basbous J, Arpin C, Gaudray G, Piechaczyk M, Devaux C, Mesnard JM. The HBZ factor of human T-cell leukemia virus type I dimerizes with transcription factors JunB and c-Jun and modulates their transcriptional activity. J Biol Chem. 2003;278(44):43620–7.

    Article  CAS  PubMed  Google Scholar 

  71. Matsumoto J, Ohshima T, Isono O, Shimotohno K. HTLV-1 HBZ suppresses AP-1 activity by impairing both the DNA-binding ability and the stability of c-Jun protein. Oncogene. 2005;24(6):1001–10.

    Article  CAS  PubMed  Google Scholar 

  72. Thebault S, Basbous J, Hivin P, Devaux C, Mesnard JM. HBZ interacts with JunD and stimulates its transcriptional activity. FEBS Lett. 2004;562(1–3):165–70.

    Article  CAS  PubMed  Google Scholar 

  73. Lemasson I, Lewis MR, Polakowski N, Hivin P, Cavanagh MH, Thebault S, Barbeau B, Nyborg JK, Mesnard JM. Human T-cell leukemia virus type 1 (HTLV-1) bZIP protein interacts with the cellular transcription factor CREB to inhibit HTLV-1 transcription. J Virol. 2007;81(4):1543–53.

    Article  CAS  PubMed  Google Scholar 

  74. Ohshima T, Mukai R, Nakahara N, Matsumoto J, Isono O, Kobayashi Y, Takahashi S, Shimotohno K. HTLV-1 basic leucine-zipper factor, HBZ, interacts with MafB and suppresses transcription through a maf recognition element. J Cell Biochem. 2010;111(1):187–94.

    Article  CAS  PubMed  Google Scholar 

  75. Hagiya K, Yasunaga J, Satou Y, Ohshima K, Matsuoka M. ATF3, an HTLV-1 bZip factor binding protein, promotes proliferation of adult T-cell leukemia cells. Retrovirology. 2011;8:19.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  76. Zhao T, Yasunaga J, Satou Y, Nakao M, Takahashi M, Fujii M, Matsuoka M. Human T-cell leukemia virus type 1 bZIP factor selectively suppresses the classical pathway of NF-kappaB. Blood. 2009;113(12):2755–64.

    Article  CAS  PubMed  Google Scholar 

  77. Zhi H, Yang L, Kuo YL, Ho YK, Shih HM, Giam CZ. NF-kappaB hyper-activation by HTLV-1 tax induces cellular senescence, but can be alleviated by the viral anti-sense protein HBZ. PLoS Pathog. 2011;7(4):e1002025.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  78. Philip S, Zahoor MA, Zhi H, Ho YK, Giam CZ. Regulation of human T-lymphotropic virus type I latency and reactivation by HBZ and Rex. PLoS Pathog. 2014;10(4):e1004040.

    Article  PubMed  PubMed Central  Google Scholar 

  79. Ma G, Yasunaga J, Fan J, Yanagawa S, Matsuoka M. HTLV-1 bZIP factor dysregulates the wnt pathways to support proliferation and migration of adult T-cell leukemia cells. Oncogene. 2013;32(36):4222–30.

    Article  CAS  PubMed  Google Scholar 

  80. Polakowski N, Terol M, Hoang K, Nash I, Laverdure S, Gazon H, Belrose G, Mesnard JM, Cesaire R, Peloponese JM, et al. HBZ stimulates brain-derived neurotrophic factor/TrkB autocrine/paracrine signaling to promote survival of human T-cell leukemia virus type 1-Infected T cells. J Virol. 2014;88(22):13482–94.

    Article  PubMed  PubMed Central  Google Scholar 

  81. Tanaka-Nakanishi A, Yasunaga J, Takai K, Matsuoka M. HTLV-1 bZIP factor suppresses apoptosis by attenuating the function of FoxO3a and altering its localization. Cancer Res. 2014;74(1):188–200.

    Article  CAS  PubMed  Google Scholar 

  82. Mukai R, Ohshima T. HTLV-1 HBZ positively regulates the mTOR signaling pathway via inhibition of GADD34 activity in the cytoplasm. Oncogene. 2014;33(18):2317–28.

    Article  CAS  PubMed  Google Scholar 

  83. Vernin C, Thenoz M, Pinatel C, Gessain A, Gout O, Delfau-Larue MH, Nazaret N, Legras-Lachuer C, Wattel E, Mortreux F. HTLV-1 bZIP factor HBZ promotes cell proliferation and genetic instability by activating OncomiRs. Cancer Res. 2014;74(21):6082–93.

    Article  CAS  PubMed  Google Scholar 

  84. Kuhlmann AS, Villaudy J, Gazzolo L, Castellazzi M, Mesnard JM, Duc Dodon M. HTLV-1 HBZ cooperates with JunD to enhance transcription of the human telomerase reverse transcriptase gene (hTERT). Retrovirology. 2007;4:92.

    Article  PubMed  PubMed Central  Google Scholar 

  85. Satou Y, Yasunaga J, Zhao T, Yoshida M, Miyazato P, Takai K, Shimizu K, Ohshima K, Green PL, Ohkura N, et al. HTLV-1 bZIP factor induces T-cell lymphoma and systemic inflammation in vivo. PLoS Pathog. 2011;7(2):e1001274.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  86. Yamamoto-Taguchi N, Satou Y, Miyazato P, Ohshima K, Nakagawa M, Katagiri K, Kinashi T, Matsuoka M. HTLV-1 bZIP factor induces inflammation through labile Foxp3 expression. PLoS Pathog. 2013;9(9):e1003630.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  87. Mitagami Y, Yasunaga J, Kinosada H, Ohshima K, Matsuoka M. Interferon-gamma promotes inflammation and development of T-Cell lymphoma in HTLV-1 bZIP factor transgenic mice. PLoS Pathog. 2015;11(8):e1005120.

    Article  PubMed  PubMed Central  Google Scholar 

  88. Miura M, Yasunaga J, Tanabe J, Sugata K, Zhao T, Ma G, Miyazato P, Ohshima K, Kaneko A, Watanabe A, et al. Characterization of simian T-cell leukemia virus type 1 in naturally infected japanese macaques as a model of HTLV-1 infection. Retrovirology. 2013;10:118.

    Article  PubMed  PubMed Central  Google Scholar 

  89. Mitobe Y, Yasunaga J, Furuta R, Matsuoka M. HTLV-1 bZIP factor RNA and protein impart distinct functions on T-cell proliferation and survival. Cancer Res. 2015;75(19):4143–52.

    Article  CAS  PubMed  Google Scholar 

  90. Panfil AR, Dissinger NJ, Howard CM, Murphy BM, Landes K, Fernandez SA, Green PL. Functional comparison of HBZ and the related APH-2 protein provides insight into human T-Cell leukemia virus type 1 pathogenesis. J Virol. 2016;90(7):3760–72.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  91. Larocque E, Halin M, Landry S, Marriott SJ, Switzer WM, Barbeau B. Human T-cell lymphotropic virus type 3 (HTLV-3)- and HTLV-4-derived antisense transcripts encode proteins with similar tax-inhibiting functions but distinct subcellular localization. J Virol. 2011;85(23):12673–85.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  92. Switzer WM, Salemi M, Qari SH, Jia H, Gray RR, Katzourakis A, Marriott SJ, Pryor KN, Wolfe ND, Burke DS, et al. Ancient, independent evolution and distinct molecular features of the novel human T-lymphotropic virus type 4. Retrovirology. 2009;6:9.

    Article  PubMed  PubMed Central  Google Scholar 

  93. Halin M, Douceron E, Clerc I, Journo C, Ko NL, Landry S, Murphy EL, Gessain A, Lemasson I, Mesnard JM, et al. Human T-cell leukemia virus type 2 produces a spliced antisense transcript encoding a protein that lacks a classic bZIP domain but still inhibits Tax2-mediated transcription. Blood. 2009;114(12):2427–38.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  94. Barbeau B, Mesnard JM. Making sense out of antisense transcription in human T-cell lymphotropic viruses (HTLVs). Viruses. 2011;3(5):456–68.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  95. Sintasath DM, Wolfe ND, Zheng HQ, LeBreton M, Peeters M, Tamoufe U, Djoko CF, Diffo JL, Mpoudi-Ngole E, Heneine W, et al. Genetic characterization of the complete genome of a highly divergent simian T-lymphotropic virus (STLV) type 3 from a wild Cercopithecus mona monkey. Retrovirology. 2009;6:97.

    Article  PubMed  PubMed Central  Google Scholar 

  96. Durkin K, Rosewick N, Artesi M, Hahaut V, Griebel P, Arsic N, Burny A, Georges M, Van den Broeke A. Characterization of novel bovine leukemia virus (BLV) antisense transcripts by deep sequencing reveals constitutive expression in tumors and transcriptional interaction with viral microRNAs. Retrovirology. 2016;13(1):33.

    Article  PubMed  PubMed Central  Google Scholar 

  97. Safari R, Jacques JR, Brostaux Y, Willems L. Ablation of non-coding RNAs affects bovine leukemia virus B lymphocyte proliferation and abrogates oncogenesis. PLoS Pathog. 2020;16(5):e1008502.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  98. Sokol M, Wabl M, Ruiz IR, Pedersen FS. Novel principles of gamma-retroviral insertional transcription activation in murine leukemia virus-induced end-stage tumors. Retrovirology. 2014;11:36.

    Article  PubMed  PubMed Central  Google Scholar 

  99. Escalera-Zamudio M, Greenwood AD. On the classification and evolution of endogenous retrovirus: human endogenous retroviruses may not be ‘human’ after all. APMIS. 2016;124(1–2):44–51.

    Article  PubMed  Google Scholar 

  100. Domansky AN, Kopantzev EP, Snezhkov EV, Lebedev YB, Leib-Mosch C, Sverdlov ED. Solitary HERV-K LTRs possess bi-directional promoter activity and contain a negative regulatory element in the U5 region. FEBS Lett. 2000;472(2–3):191–5.

    Article  CAS  PubMed  Google Scholar 

  101. Agoni L, Guha C, Lenz J. Detection of human endogenous Retrovirus K (HERV-K) transcripts in human prostate Cancer cell lines. Front Oncol. 2013;3:180.

    Article  PubMed  PubMed Central  Google Scholar 

  102. Montesion M, Bhardwaj N, Williams ZH, Kuperwasser C, Coffin JM. Mechanisms of HERV-K (HML-2) Transcription during Human Mammary Epithelial Cell Transformation. J Virol 2018, 92(1).

  103. Mikhalkevich N, O’Carroll IP, Tkavc R, Lund K, Sukumar G, Dalgard CL, Johnson KR, Li W, Wang T, Nath A, et al. Response of human macrophages to gamma radiation is mediated via expression of endogenous retroviruses. PLoS Pathog. 2021;17(2):e1009305.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  104. Contreras-Galindo R, Kaplan MH, Leissner P, Verjat T, Ferlenghi I, Bagnoli F, Giusti F, Dosik MH, Hayes DF, Gitlin SD, et al. Human endogenous retrovirus K (HML-2) elements in the plasma of people with lymphoma and breast cancer. J Virol. 2008;82(19):9329–36.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  105. Contreras-Galindo R, Kaplan MH, Contreras-Galindo AC, Gonzalez-Hernandez MJ, Ferlenghi I, Giusti F, Lorenzo E, Gitlin SD, Dosik MH, Yamamura Y, et al. Characterization of human endogenous retroviral elements in the blood of HIV-1-infected individuals. J Virol. 2012;86(1):262–76.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  106. Sokol M, Jessen KM, Pedersen FS. Utility of next-generation RNA-sequencing in identifying chimeric transcription involving human endogenous retroviruses. APMIS. 2016;124(1–2):127–39.

    Article  CAS  PubMed  Google Scholar 

  107. Svoboda P, Stein P, Anger M, Bernstein E, Hannon GJ, Schultz RM. RNAi and expression of retrotransposons MuERV-L and IAP in preimplantation mouse embryos. Dev Biol. 2004;269(1):276–85.

    Article  CAS  PubMed  Google Scholar 

  108. Kumari P, Sampath K. cncRNAs: bi-functional RNAs with protein coding and non-coding functions. Semin Cell Dev Biol. 2015;47–48:40–51.

    Article  PubMed  PubMed Central  Google Scholar 

  109. Nam JW, Choi SW, You BH. Incredible RNA: dual functions of Coding and Noncoding. Mol Cells. 2016;39(5):367–74.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  110. Dhamija S, Menon MB. Non-coding transcript variants of protein-coding genes - what are they good for? RNA Biol. 2018;15(8):1025–31.

    PubMed  PubMed Central  Google Scholar 

  111. Ruiz-Orera J, Messeguer X, Subirana JA, Alba MM. Long non-coding RNAs as a source of new peptides. eLife. 2014;3:e03523.

    Article  PubMed  PubMed Central  Google Scholar 

  112. Ingolia NT, Lareau LF, Weissman JS. Ribosome profiling of mouse embryonic stem cells reveals the complexity and dynamics of mammalian proteomes. Cell. 2011;147(4):789–802.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  113. Zeng C, Fukunaga T, Hamada M. Identification and analysis of ribosome-associated lncRNAs using ribosome profiling data. BMC Genomics. 2018;19(1):414.

    Article  PubMed  PubMed Central  Google Scholar 

  114. Anderson DM, Anderson KM, Chang CL, Makarewich CA, Nelson BR, McAnally JR, Kasaragod P, Shelton JM, Liou J, Bassel-Duby R, et al. A micropeptide encoded by a putative long noncoding RNA regulates muscle performance. Cell. 2015;160(4):595–606.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  115. Nelson BR, Makarewich CA, Anderson DM, Winders BR, Troupes CD, Wu F, Reese AL, McAnally JR, Chen X, Kavalali ET, et al. A peptide encoded by a transcript annotated as long noncoding RNA enhances SERCA activity in muscle. Science. 2016;351(6270):271–5.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  116. Stein CS, Jadiya P, Zhang X, McLendon JM, Abouassaly GM, Witmer NH, Anderson EJ, Elrod JW, Boudreau RL. Mitoregulin: a lncRNA-Encoded microprotein that supports mitochondrial Supercomplexes and respiratory efficiency. Cell Rep. 2018;23(13):3710–3720e3718.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  117. van Heesch S, Witte F, Schneider-Lunitz V, Schulz JF, Adami E, Faber AB, Kirchner M, Maatz H, Blachut S, Sandmann CL, et al. Translational Landsc Hum Heart Cell. 2019;178(1):242–260e229.

    Google Scholar 

  118. Hube F, Velasco G, Rollin J, Furling D, Francastel C. Steroid receptor RNA activator protein binds to and counteracts SRA RNA-mediated activation of MyoD and muscle differentiation. Nucleic Acids Res. 2011;39(2):513–25.

    Article  CAS  PubMed  Google Scholar 

  119. Candeias MM, Malbert-Colas L, Powell DJ, Daskalogianni C, Maslon MM, Naski N, Bourougaa K, Calvo F, Fahraeus R. P53 mRNA controls p53 activity by managing Mdm2 functions. Nat Cell Biol. 2008;10(9):1098–105.

    Article  CAS  PubMed  Google Scholar 

  120. Faghihi MA, Modarresi F, Khalil AM, Wood DE, Sahagan BG, Morgan TE, Finch CE, St Laurent G 3rd, Kenny PJ, Wahlestedt C. Expression of a noncoding RNA is elevated in Alzheimer’s disease and drives rapid feed-forward regulation of beta-secretase. Nat Med. 2008;14(7):723–30.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  121. Uchida T, Rossignol F, Matthay MA, Mounier R, Couette S, Clottes E, Clerici C. Prolonged hypoxia differentially regulates hypoxia-inducible factor (HIF)-1alpha and HIF-2alpha expression in lung epithelial cells: implication of natural antisense HIF-1alpha. J Biol Chem. 2004;279(15):14871–8.

    Article  CAS  PubMed  Google Scholar 

  122. Lapidot M, Pilpel Y. Genome-wide natural antisense transcription: coupling its regulation to its different regulatory mechanisms. EMBO Rep. 2006;7(12):1216–22.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  123. Pelechano V, Steinmetz LM. Gene regulation by antisense transcription. Nat Rev Genet. 2013;14(12):880–93.

    Article  CAS  PubMed  Google Scholar 

  124. Wight M, Werner A. The functions of natural antisense transcripts. Essays Biochem. 2013;54:91–101.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  125. Zinad HS, Natasya I, Werner A. Natural antisense transcripts at the interface between host genome and Mobile Genetic Elements. Front Microbiol. 2017;8:2292.

    Article  PubMed  PubMed Central  Google Scholar 

  126. Weinberger LS, Shenk T. An HIV feedback resistor: auto-regulatory circuit deactivator and noise buffer. PLoS Biol. 2007;5(1):e9.

    Article  PubMed  Google Scholar 

  127. Razooky BS, Cao Y, Hansen MMK, Perelson AS, Simpson ML, Weinberger LS. Nonlatching positive feedback enables robust bimodality by decoupling expression noise from the mean. PLoS Biol. 2017;15(10):e2000841.

    Article  PubMed  PubMed Central  Google Scholar 

  128. Aull KH, Tanner EJ, Thomson M, Weinberger LS. Transient thresholding: a mechanism enabling noncooperative transcriptional circuitry to form a switch. Biophys J. 2017;112(11):2428–38.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  129. Levine E, Zhang Z, Kuhlman T, Hwa T. Quantitative characteristics of gene regulation by small RNA. PLoS Biol. 2007;5(9):e229.

    Article  PubMed  PubMed Central  Google Scholar 

  130. Chen D, Arkin AP. Sequestration-based bistability enables tuning of the switching boundaries and design of a latch. Mol Syst Biol. 2012;8:620.

    Article  PubMed  PubMed Central  Google Scholar 

  131. Ferrell JE Jr, Ha SH. Ultrasensitivity part III: cascades, bistable switches, and oscillators. Trends Biochem Sci. 2014;39(12):612–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  132. Ferrell JE Jr, Ha SH. Ultrasensitivity part II: multisite phosphorylation, stoichiometric inhibitors, and positive feedback. Trends Biochem Sci. 2014;39(11):556–69.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  133. Ferrell JE Jr, Ha SH. Ultrasensitivity part I: michaelian responses and zero-order ultrasensitivity. Trends Biochem Sci. 2014;39(10):496–503.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  134. Legewie S, Dienst D, Wilde A, Herzel H, Axmann IM. Small RNAs establish delays and temporal thresholds in gene expression. Biophys J. 2008;95(7):3232–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  135. Duhring U, Axmann IM, Hess WR, Wilde A. An internal antisense RNA regulates expression of the photosynthesis gene isiA. Proc Natl Acad Sci U S A. 2006;103(18):7054–8.

    Article  PubMed  PubMed Central  Google Scholar 

  136. Shimoni Y, Friedlander G, Hetzroni G, Niv G, Altuvia S, Biham O, Margalit H. Regulation of gene expression by small non-coding RNAs: a quantitative view. Mol Syst Biol. 2007;3:138.

    Article  PubMed  PubMed Central  Google Scholar 

  137. Mehta P, Goyal S, Wingreen NS. A quantitative comparison of sRNA-based and protein-based gene regulation. Mol Syst Biol. 2008;4:221.

    Article  PubMed  PubMed Central  Google Scholar 

  138. Xu Z, Wei W, Gagneur J, Clauder-Munster S, Smolik M, Huber W, Steinmetz LM. Antisense expression increases gene expression variability and locus interdependency. Mol Syst Biol. 2011;7:468.

    Article  PubMed  PubMed Central  Google Scholar 

  139. Sylvie, Briquet Catherine, Vaquero (2002) Immunolocalization Studies of an Antisense Protein in HIV-1-Infected Cells and Viral Particles Virology 292(2) 177-184 https://doi.org/10.1006/viro.2001.1224

Download references

Acknowledgements

The author acknowledges colleagues at Johns Hopkins University and other institutions for helpful discussions.

Funding

This work was supported by National Institutes of Health grants AI144983 and AI172665 to FR.

Author information

Authors and Affiliations

Authors

Contributions

F.R. conceived, wrote and edited the manuscript; F.R. prepared figures and tables.

Corresponding author

Correspondence to Fabio Romerio.

Ethics declarations

Competing interests

The author declares that he has no competing interests

Ethics approval and consent to participate

Not applicable.

Consent to publish

Not applicable.

Additional information

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Romerio, F. Origin and functional role of antisense transcription in endogenous and exogenous retroviruses. Retrovirology 20, 6 (2023). https://doi.org/10.1186/s12977-023-00622-x

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s12977-023-00622-x