Virology and Emerging Diseases - Sci Forschen

Full Text

REVIEW ARTICLE
DNA unwinding by Viral Protein R Initializes Complicated Cellular Responses in HIV-1 Infection: Defining the Viper’s First Bite

  Kenta Iijima      Yukihito Ishizaka*   

Department of Intractable Diseases, National Center for Global health and Medicine, 1-21-1 Toyama, Shinjuku-ku, Tokyo, 162-8655, Japan

*Corresponding author: Yukihito Ishizaka, Department of Intractable Diseases, National Center for Global health and Medicine, 1-21-1 Toyama, Shinjuku-ku, Tokyo, 162-8655, Japan, E-mail: zakay@ri.ncgm.go.jp


Abstract

Vpr, an accessory gene of human immunodeficiency virus (HIV) type1 that encodes a virion-associated protein, induces multiple cellular responses, e.g. transcriptional regulation, chromatin modulation and the DNA damage response with cell-cycle checkpoint activation. Vpr may promote HIV-1 infection of quiescent cells, including resting macrophages, but it remained elusive how Vpr contributes to viral infection into these cells. With the object of clarifying the role of Vpr, we first summarize the pleiotropic function of Vpr in the HIV-1 life cycle and subsequently the recently identified mode of Vpr-induced DNA damage triggered by unwinding of DNA in association with viral DNA integration. Finally, we discuss the role of Vpr in HIV-associated diseases, which are important issues for HIV-1–infected patients in the post-antiretroviral therapy era.

Keywords

HIV-1, Vpr, DNA damage, Integration, Macrophage


Introduction

Combined antiretroviral therapy (cART) for human immunodeficiency virus (HIV)-1 can suppress viral replication, such that viral DNA is at a non-detectable level in the blood of HIV-1–positive patients, and can prevent the immunodeficiency caused by T-cell depletion. After interruption of the cART regimen, however, viral replication readily resumes in long-lived reservoir cells [1]. Because macrophages are the major cell type involved in the formation of viral reservoirs, it is important to understand the mode of viral infection of resting macrophages [2,3].

Viral protein R (Vpr), an accessory gene of HIV-1, encodes a virion-associated nuclear protein of 96 amino acids and is proposed to facilitate viral infection of macrophages [4-8]. Structural analyses revealed that Vpr contains three α-helices with self-dimerisation/ oligomerisation activity and a flexible carboxyl (C)-terminal region that includes a stretch of basic amino acids with DNA-binding activity [9-11]. Vpr possesses pleiotropic functions for stimulating viral replication in multiple steps [12,13]. Following viral entry into a target cell, the viral RNA is reverse transcribed into linear doublestranded DNA (dsDNA) in the cytoplasm [14]. The synthesized viral genomic DNA (vDNA) is wrapped in viral proteins to form a preintegration complex (PIC), subsequently translocated to the nucleus and integrated into the host chromosome. The integration step of the vDNA is catalyzed by integrase (IN), which assembles at both ends of vDNA to form a strand–transfer complex, which facilitates establishment of a lasting infection [15,16]. The integrated proviral DNA serves as a template for the transcription of viral genes and is propagated along with the host genome.

Vpr induces a cell-cycle abnormality during the G2 /M phase as well as a variety of DNA damage responses (DDR) involving phosphorylation of histone H2AX and focus formation by multiple DNA repair proteins, and activation of ataxia telangiectasia mutated (ATM), ATM and Rad3-related (ATR); central kinases response to DNA double-strand break (DSB) and single-strand break (SSB)/ replication stress, respectively [17-23]. Regarding to the mechanism by which Vpr-induced G2 /M arrest promotes viral infection, it was proposed that Vpr provides a cellular environment suitable for viral gene transcription and/or delays cell death during viral replication [12,13,24]. In contrast, the biological importance of Vpr in resting (G0 /G1 phase) macrophages is unclear. Smith et al. and others reported that DSBs promote HIV-1 infection and/or integration of vDNA [8,25,26]. We also demonstrated that Vpr-induced DSBs provide sites for vDNA integration [27]. Although the mode of Vprinduced DSBs has been unknown, we recently discovered that Vpr alters DNA structure and activates both the ATM and ATR pathways [27]. The findings revealed that unwinding of dsDNA by Vpr leads to supercoiling of DNA, accumulation of topoisomerase 1 (Topo1) and DSBs. Of note, unwinding of dsDNA recruits replication protein A (RPA) 70, a single-stranded DNA (ssDNA)-binding subunit of the RPA complex, which activates ATR-dependent DDR [28]. Because DSBs may induce multiple cellular responses, including apoptosis and immune activation, it is likely that Vpr-induced DSB is the molecular basis of multiple acquired immunodeficiency syndrome (AIDS)- related symptoms.

Here we summarize the biological functions of Vpr related to DSBs and discuss the potential of Vpr as a target of candidate anti–HIV-1 compounds for treating HIV-1–positive patients.

1. Vpr regulates Viral Infection at Multiple Steps

1a. Reverse transcription

HIV particles enclose viral proteins that include the nucleocapsid (NC), matrix (MA), IN, reverse transcriptase (RT), Vif, Vpr and Nef proteins [12,13]. Vpr is incorporated into viral particles through its association with the p6 domain of Gag [29]. After entry into a target cell, the viral RNA is reverse-transcribed in the cytoplasm into a linear dsDNA by the reverse transcription complex, which contains two copies of viral genomic RNA, as well as the RT, IN, NCp7, MA, Vpr and cellular proteins [15]. Vpr associates with the reverse transcription complex and modulates the quantity and quality of HIV-1 RT, an error-prone RNA-dependent DNA polymerase [15]. Without Vpr, the newly synthesized vDNA has a high frequency of mutations [30,31]. Quality control by Vpr is in part dependent on uracil DNA glycosylase (UNG) 2, which excises uracil bases from DNA [32] (Figure 1). UNG2 is required for vDNA synthesis, because uracil bases in DNA can result from misincorporation of dUTP or deamination of cytosine. Notably, Vpr associates with and incorporates UNG2 into viral particles. The involvement of UNG2 in Vpr-dependent mutation suppression was demonstrated using a VprW54R-UNG2 fusion protein [32]. The tryptophan residue of Vpr at amino acid 54 is essential for the interaction with UNG2 and its incorporation into viral particles. When the VprW54R-UNG2 fusion protein was incorporated into viral particles, the mutation rate of vDNA was decreased with comparable to that of Vpr wild-type virus. Interestingly, however, a catalytically inactive mutant of UNG2 fused to VprW54R also suppressed mutation in cells infected with the Vpr-defective virus. Further analysis revealed that p32, a subunit of the RPA complex that associates with UNG2, prevents mutation of vDNA by a mechanism different from that of UNG2 [32]. Although the precise role of p32 is unclear, it may protect vDNA from host nucleases.

Figure 1: Involvement of Vpr in viral replication at multiple steps.
After the entry of HIV-1, viral RNA (blue wavy line) associates with Vpr is reverse-transcribed in double strand vDNA (black and red wavy lines). Vpr facilitates the incorporation of vDNA through interaction with importin-α (Imp-α). Vpr promotes the viral genes (or genomic RNA) transcription from LTR by utilizing cellular transcriptional factors. Transcribed genomic RNA is incorporated to new virions. At this time, Vpr transfers UNG2 (and RPA32) to new virions, which suppresses frequency of mutation rate in subsequent reverse-transcription. On the contrary, Vpr promotes proteasome-mediated degradation of UNG2. Extracellular Vpr stimulates TLR4/MyD88 pathway, which leads to activation of IKK, followed by NFκB nuclear translocation. Besides to association with TLR4, Vpr directly activates by interacting with TAK1, leading to NFκB activation. By binding to ANT/VDAC, Vpr induces the release of cyt c from mitochondrial inter-membrane space to cytoplasm. Complex of Apaf-1 and cyt c (apoptosome) activates capspase-9,-3 dependent apoptotic cell death. This intrinsic apoptosis induction is facilitated by Bax expression, while Vpr is known to downregulated Bax by inhibiting p53 activity. A part of figure was generated through the use of QIAGEN’s Ingenuity® Pathway Analysis (IPA®, QIAGEN Redwood City, www.qiagen.com/ingenuity).

In contrast, UNG2 suppresses vDNA synthesis [33-35]. UNG2 converts APOBEC3G-induced deaminated cytosines to a basic site, leading to cleavage by apurinic/apyrimidinic endonuclease. Thus, Vpr induces proteasomal degradation of UNG2. The roles of UNG2 in viral replication are controversial, but Vpr contributes to vDNA synthesis by modulating both functions of UNG2.

Gleenberg et al. reported that Vpr interferes with reverse transcription by directly interacting with RT [36], whereas Lyonnais et al. observed that Vpr induced the bridging and condensation of synthesized DNA to promote its nuclear import [37]. Although the biological importance of Vpr and RT is not fully understood, the close functional associations among Vpr, RT and IN may link reverse transcription and nuclear import, followed by chromosomal integration.

1b. Nuclear import and transactivation of the long terminal repeat

Vpr may promote HIV-1 replication in resting macrophages, but it is dispensable for virus replication in proliferating cells [4-8]. In resting macrophages, vDNA must be imported into the nucleus independently of cell division and without disrupting the nuclear membrane. vDNA associates with both viral and host proteins to form the PIC. Among the four viral proteins present in the PIC (RT, IN, MA and Vpr), Vpr reportedly facilitates nuclear import of the PIC by directly binding to importin-α [38,39] (Figure 1). Alternatively, Vpr may shuttle between the cytoplasm and nucleus using an exportin-1-dependent nuclear export signal [40]. Nucleocytoplasmic shuttling by Vpr is required not only for incorporation of Vpr in newly produced virions but also for efficient replication of HIV-1 in tissue macrophages. Vpr associates with the nuclear pore complex, particularly the nucleoporins p54, p58 and hCG1, and accumulates in the nuclear envelope [41,42]. Interestingly, integration of HIV-1 DNA occurs in the nucleus close to nuclear pores, where a series of cellular genes are transcriptionally active [43,44]. It is important to determine whether Vpr regulates vDNA integration into transcriptionally active chromatin by interacting with the nuclear pore complex, because the chromatin status at the time of viral integration is a critical determinant of the life cycle of HIV-1 [45].

Vpr transactivates the long terminal repeat (LTR) of HIV-1, which promotes viral replication and pathogenesis [12,13,46] (Figure 1). Vpr was initially reported to reactivate latent proviral DNA, and several activating elements—NF-AT, glucocorticoid response elements (GREs), NF-κB, Sp1, transactivation response element and a TATA box—were subsequently identified in the U3 region of the HIV-1 LTR [46]. Vpr interacts with the Sp1–DNA complex and may transactivate the Sp1-bound promoter by stabilising Sp1–promoter complexes [47]. Sawaya et al. reported that Vpr transactivates LTR in an Sp1- dependent manner, whereas p53 suppresses transactivation of the LTR [48]. Moreover, Vpr inhibited the transcriptional activity of p53, which was proposed to promote replication of HIV-1–infected cells without inducing apoptosis [48]. Vpr binds and augments the activity of several transcription factors, e.g. glucocorticoid receptor (GR), transcription factor IIB and p300 [12,13,46]. Vpr complexes with GR and transactivates the HIV-1 LTR and GRE-containing promoters in a manner dependent on the LXXLL GR-binding motif at amino acids 64–68 (helix-2) of Vpr [49]. Additionally, association of Vpr with p300 and TFIIH cooperatively induces GRE activation [50]. Furthermore, Vpr forms a tertiary complex of Vpr–GR–poly (ADP-ribose) polymerase-1 (PARP-1), and suppresses NF-κB by preventing the nuclear localisation of PARP-1, a co-activator of NF-κB [51]. However, Vpr reportedly activates NF-κB by enhancing the phosphorylation of TAK1 and IKK. These findings imply that Vpr modulates NF-κB activity to promote viral replication [52,53].

Vpr activates LTR throughout viral infection and replication. Hoshino et al. showed that extracellular Vpr activates NF-κB in a TLR4/MyD88-dependent manner and induces IL-6 production with subsequent reactivation of latent HIV-1 [54]. Vpr induces reactivation of LTR in latently infected cells by promoting proteasomal degradation of histone deacetylases 1 and 3, which leads to silencing of viral genes [55,56].

Additionally, Vpr in cooperation with Tat transactivates the HIV1 LTR by concomitantly occupying the multiple binding sites within the LTR [57]. Notably, Vpr but not Tat enhances Nef expression from unintegrated vDNA, indicating a role for Vpr during the early stage of viral infection [58].

1c. Induction of apoptosis

Vpr disrupts mitochondrial function and activates the intrinsic apoptosis pathway by associating via its C-terminus with two candidate factors of the permeability transition pore complex (PTPC): adenine nucleotide translocator (ANT) and voltage-dependent anion channel (VDAC) [59,60] (Figure 1). Jacotot et al. reported that synthetic Vpr induced mitochondrial membrane permeabilisation by the PTPC [59]. Furthermore, the C-terminus of Vpr cooperates with Bax to increase the permeability of ANT-containing liposomes. The Vpr-induced mitochondrial membrane permeabilisation leads to release of cytochrome c (cyt c) from the intermembrane space. The resulting apoptosome composed of cyt c, dATP, Apaf-1 and procaspse-9 activates procaspase-9 and -3, which mediate apoptotic DNA fragmentation. Although a clear apoptotic cascade was proposed based on in vitro data [59], Vpr may play different roles in vivo because Vpr-induced apoptosis was prevented by preincubation of Vpr with DNA or RNA [59]. Anderson et al. reported that Vprinduced apoptotic cell death is dependent on Bax, but not on ANT [60]. Moreover, Bax activation requires DDR activation, which involves ATR-mediated phosphorylation of BRCA1 and subsequent upregulation of GADD45α [61]. However, whether Bax-mediated apoptosis is coupled to Vpr-induced DDR activation is controversial. In contrast, Conti et al. reported that low levels of Vpr exerted an anti-apoptotic effect in CD4+ T-cell lines by upregulating Bcl-2 and downregulating Bax [62]. Thus, the role of Vpr in different cell types warrants further investigation.

1d. Vpr modulates the immune response

Vpr modulates the immune response directly and indirectly. Exogenous Vpr secreted by infected cells inhibits the immune response by inducing apoptosis and cell-cycle arrest [12,13]. Regarding the indirect effects, Vpr increases the expression of TNF-α in dendritic cells, which leads to induction of apoptosis in CD8+ T-cells [63]. Additionally, Vpr upregulates unique-long 16 binding proteins 1 and 2 (ULBP1/2), ligands for natural killer group 2 member D (NKG2D), in CD4+ T cells with concomitant activation of ATR kinase [64-66] (Figure 2). Upregulation of ULBP1/2 increases the cytolytic activity of NK cells and promotes clearance of infected cells. In contrast, the increased and decreased TGF-β and IL-12 production, respectively, in Vpr-affected CD4+ T cells suppresses the activity of NK cells by downregulating CD107α and interferon (IFN)-γ [67,68] (Figure 2). Vpr exerts complex effects on the immune system, and its influence on clearance of infected cells should be further investigated to understand the role of Vpr in T-cell depletion.

Figure 2: Vpr modulates both intra- and extra-cellular environment
HIV-1 infection to macrophages evokes cGAS/STING and NFκB activation, which is enhanced by Vpr. Resulting upregulation of ISGs facilities tissue inflammation. Furthermore, Vpr induces dysregulation of glucose and glutamate metabolism. Increased concentration of extracellular glutamate shows neurotoxic effect and induces neural cell death. By the infection of HIV-1 to CD4+ T-cells, ATR activity is elevated in Vpr dependent manner. ATR activation induces upregulation of ULBP1/2, ligands of NKG2D, by which cytotoxic activity of NK-cells is increased, while Vpr dysregulates the production of cytokines from HIV-1–infected CD4+ T-cells. Thus, Vpr suppresses the activity of NK-cells, which is characterized by the reduced expression of CD107α and IFNγ. TGF-β, an immune suppressive cytokine; IL-12, an immune stimulative cytokine. A part of figure was generated through the use of QIAGEN’s Ingenuity® Pathway Analysis (IPA®, QIAGEN Redwood City, www.qiagen.com/ingenuity).

Transcriptomic analyses of monocyte-derived macrophages and dendritic cells revealed that Vpr affects the cellular immune response by upregulating IFN-stimulatory genes (ISGs): MX1, MX2, ISG15, ISG20, IFIT1, IFIT2, IFIT3, IFI27, IFI44L and TNFSF10 [69,70] (Figure 2). Importantly, in CD4+ cells, Vpr increased the expression of type I IFNs and ISGs induced by cGAS-dependent sensing of HIV-1 DNA, which could contribute to hyperimmune activation and the progression of AIDS [71]. Interestingly, treatment of mouse macrophage-like cells with recombinant Vpr induced IFN-β production by activating the cGAS/STING pathway and mobilizing long interspersed element-1 (LINE-1), an abundant endogenous retrotransposon [72]. Proteomic analysis of Vpr-regulated proteins in monocyte-derived macrophages revealed that Vpr disturbs glucose and glutamate metabolism [73]. Vpr modulates the glycolytic pathway by increasing the expression of pyruvate kinase muscle type 2, which phosphorylates tyrosine 705 of STAT3, which can subsequently upregulate the HIV-1 LTR [73]. Notably, dysregulation of glutamate metabolism, including downregulation of glutamate dehydrogenase 2 and upregulation of glutaminase C, may be linked to HIV-associated neurocognitive disorder (HAND) via the neurotoxic effects of glutamate [74] (Figure 2). Together with its pro-apoptotic effect on neurons, these results suggest Vpr to be a target for novel therapeutic strategies for HAND [75-80].

2. Cell-Cycle Abnormality during the G2/M Phase and Activation of the DNA Damage Response

Cells are equipped with sophisticated system responding to DNA damages to counteract cytotoxic effects by exogenous and endogenous DNA damages, including DNA single- or double-strand breaks, chemical modifications and replication stress. Among enormous numbers of players in DDR, ATM, ATR and DNA-PKcs (DNAdependent protein kinase catalytic subunit) function as the most upstream kinases and activate the downstream events including repair of DNA damages and activation of cell-cycle checkpoint [23]. The association between retroviral integration and DDR was initially investigated by Daniel et al. [81,82]. IN catalyses vDNA integration by transferring the 3’-terminus of vDNA to nicked host genomic DNA, because the IN-induced flanking nicks on both strands of DNA are recognised as DSBs by the cellular DNA repair machinery [15,81,82]. Given that the ligation of vDNA to host DNA requires processing of DNA ends, retroviral DNA integration must involve the host DNA repair pathway, as observed in the process of removing flapped DNA strands by FEN1 [83]. Additionally, numbers of evidence indicate the indispensable role of cellular DSB repair pathway. DSB repair by DNAPK-dependent non-homologous end joining is required for vDNA integration and the survival of infected cells [84]. Lau et al. reported that the activity of ATM is required for survival of HIV-1–infected CD4+ T cells [85]. Data indicate that microhomology-mediated end joining (i.e. using short patch of homologous DNA sequences at the ends of DSBs) is involved in vDNA integration or the repair of INgenerated DSBs [86].

Importantly, DNA damage induced by external factors increases viral infectivity, particularly in quiescent cells including resting macrophages [8,25,26]. The HIV-1 LTR is more active during the G2 phase, implying that Vpr-induced cell-cycle arrest optimizes the cellular environment for viral replication [12,13,24]. However, Vif, an HIV-1 accessory protein, induces G2 /M cell-cycle arrest, suggesting that promotion of viral infection via induction of DDR is not unique to Vpr [87]. Additionally, the biological importance of Vpr-induced DDR in resting macrophages is unclear, because these cells are in G0/G1 phase and express low levels of ATR and Chk1 [21]. Smith et al. and others reported that DSBs promote HIV-1 infection and/or integration of vDNA [8,25,26]. We also reported that DNA damaging agents promote viral infection of resting cells [8,27]. Moreover, DSB sites could facilitate vDNA integration by providing a platform for vDNA integration [8,27]. Notably, even the IN-D64A mutant virus, which is deficient in IN activity, could integrate into DSB sites. Although DSB-directed viral integration is a small population of the whole viral infection, it may hamper complete virus elimination by cART [8,27].

2a. ATR activation with Vpr-induced ubiquitination

Vpr was initially reported to inhibit the proliferation of HIV-1– infected T cells at the G2/M phase by hyperphosphorylating CDC2, a cyclin-dependent kinase complexed with cyclin B [88]. Vpr-induced cell-cycle arrest is dependent on ATR kinase and accompanied by activation of a typical DDR, including phosphorylation of histone H2AX and formation of BRCA1, 53BP1, RAD51 and FANCD2 foci [17-22,89]. Although Vpr induced a DDR including focus formation of DSB repair proteins and accumulation of RPA [21,27], ATR has been mostly highlighted for approaching the mode of Vpr-induced DDR, because ATM is not essential for Vpr-induced G2/M arrest [21].

Vpr-induced DDR requires ubiquitination of cellular protein(s) by Vpr-binding protein (VprBP)/ DNA damage-binding protein 1 (DDB1), components of the cullin E3-ubiquitin ligase complex [90-93]. Depletion of VprBP or DDB1 abrogates Vpr-induced DDR activation, and the Q65R mutant of Vpr, which is unable to bind to VprBP, did not induce a DDR [90,91]. Belzile et al. reported that Vpr-induced DDR required K48-linked ubiquitination of cellular proteins, indicating a pivotal role for proteasomal degradation [93]. Despite much effort, the factors responsible for Vpr-induced DDR and authentic targets of Vpr-dependent ubiquitination are not known. UNG2 was first identified as a DDR-associated factor that is degraded by Vpr-induced ubiquitination [94]. Because UNG2 mediates baseexcision repair, Vpr-induced degradation of UNG2 may increase the magnitude of DNA damage [94]. This scenario fits to helicaselike transcription factor (HTLF), a translocase involved in repair of damaged replication forks [95]. Klockow et al. reported that Dicer, an endonuclease subunit of RNA-induced silencing complex (RISC) that may suppress viral replication, is a candidate target of Vpr-induced ubiquitination [96]. Small RNAs are generated by RISC immediately after DNA damage and activate the DDR [97-99]. Thus, DNA damage would be increased if the RISC system is impaired by Vpr. Additionally, telomerase reverse transcriptase, an RT subunit of telomerase, may be a target of Vpr-induced ubiquitination [100]. If so, Vpr may destabilize telomeric DNA. It is possible that such Vpr-induced telomeric DNA instability is linked to the chromosomal breakage/fusion/bridge cycle, as observed in Vpr-expressing cells [101].

2b. Involvement of the replication fork and SLX4 in the Vprinduced DDR

Our understanding of Vpr function was recently advanced by the identification of SLX4 as a Vpr-binding factor [89]. SLX4 is a structurespecific endonuclease that complexes with Mus81/Eme1 and plays roles in resolving DNA replication and recombination intermediates [102,103]. Consequently, downregulation of SLX4 induces the DDR as a result of accumulation of DNA damage. In contrast, Laguette et al. proposed that the SLX4 complex is prematurely activated by Vpr and directly induces the DDR by cleaving DNA strands [89]. In addition to DDR activation, Vpr-activated SLX4 degrades excess vDNA to prevent the activation of cellular immune sensing, promoting viral replication [89].

DNA replication may be involved in the Vpr-induced DDR because the latter resembles the DDR caused by the DNA-synthesis inhibitors hydroxyurea and aphidicolin [21]. Indeed, Vpr-induced DDR depends on ATR, a kinase activated by RPA-coated ssDNA during DNA replication. Interestingly, Romani et al. identified minichromosome maintenance (MCM) 10, a component of the MCM complex with roles in DNA replication initiation and strand elongation, as a novel target of Vpr [104]. They demonstrated that Vpr induces proteasomal degradation of MCM10 as a required step of Vpr-induced cell-cycle arrest.

3. Integrated Mode of the Vpr-Induced DDR and DSBs

Vpr induces cell-cycle arrest at the G2/M phase by ATR-dependent activation of the DDR [17,21], DDB1/VprBP-dependent ubiquitination [90-93], stalled replication folk [21], and recruitment of RPA and SLX4 [21,27,89]. However, the critical molecular event responsible for these cellular responses is unclear. We discovered that Vpr alters the structure of dsDNA, which may uncover the links among these factors [27]. An overview of our findings follows (Figure 3).

Figure 3: Mode of Vpr-induced DDR and DSB
At first, Vpr binds and unwinds dsDNA. Partial unwinding of dsDNA promotes loading of RPA70, which activates ATR-dependent DDR. Then, chromatin remodeling is provoked by Vpr-induced ubiquitination of histone H2B, and promotes the loading of RPA70, leading to full activation of ATR pathway. On the other hand, unwinding of dsDNA accumulates supercoiling of DNA, which recruits Topo1 for relaxation. Excess accumulation of Topo1 induces covalent complex of Topo1 and DNA, where replication and/or transcription machineries collide and DSB is induced. Vpr-induced DSB is targeted by vDNA integration. SLX4 is involved in the resolution of stalled replication fork and the excision Topo1-bound DNA, therefore causes DDR. Additionally, SLX4 is required for degradation of excess amounts of vDNA to escape from cGAS/STING-immune sensing. A part of figure was generated through the use of QIAGEN’s Ingenuity® Pathway Analysis (IPA®, QIAGEN Redwood City, www.qiagen.com/ingenuity).

3a. Unwinding of dsDNA is the initial step in the Vprinduced DDR

First, atomic force microscopy and supercoiling 294 assays revealed that Vpr induces morphological changes in supercoiled DNA [27]. A mutational analysis indicated that the four arginines in the C-terminal stretch of Vpr are required for the topological changes. Moreover, Vpr induces loading of RPA70 onto dsDNA, by unwinding of dsDNA (Figure 3). These findings support our hypothesis that the structural alternation of dsDNA by Vpr triggers RPA-dependent activation of ATR.

To demonstrate that Vpr plays a similar role in cells, we used a LacO/ LacI system in human cells [105,106]. Using this system, LacR-fused Vpr can be tethered to the chromosome region containing the LacO array [27]. We first analyzed whether Vpr unwinds the corresponding region using psoralen, a DNA-intercalating cross-linker that preferentially binds to negatively supercoiled DNA [107,108]. LacRfused Vpr increased the unwinding of dsDNA at the LacO repeats in a C-terminal stretch-dependent manner. Next, we analyzed RPA70 accumulation on the LacO repeats by chromatin immunoprecipitation assay. The results showed that RPA70 accumulation is also dependent on the C-terminal stretch of Vpr. Because formation of RPA-coated ssDNA triggers ATR activation, Vpr-induced unwinding of dsDNA could lead to DDR activation.

3b. Mechanism of DSB induction by Vpr

In vitro, Vpr increased the supercoiling of DNA. Accumulation of supercoiled DNA recruits Topo1, which relaxes supercoiled DNA by cleavage/ligation cycle for regulating the DNA replication, transcription and repair [109]. Excess accumulation of Topo1 leads to stable formation of the covalent complex of Topo1 and DNA (Topo1- cc), which induces DSBs by colliding with the replication fork or transcriptional machinery [110]. In quiescent cells, DSBs are likely induced by the latter mechanism.

As expected, Topo1-cc accumulated in Vpr-expressing cells (Figure 3). Consistently, the number of DSBs was increased by Vpr expression in a Topo1-dependent manner [27]. Using the LacO/LacI system, we demonstrated that Topo1-cc and DSBs accumulate at Vpr-tethered LacO repeats. Moreover, vDNA was integrated at Vprinduced DSB sites, indicating a role for Vpr in providing sites for vDNA integration. The importance of Vpr-directed viral integration is unknown because it accounts for only small fraction of overall HIV-1 infection. Vpr-directed viral integration may influence the selection of integration sites. Thus, Vpr accumulation at centromeres through its interaction with HP1 is interesting [111], because the heterochromatic status of these regions likely contributes to latent viral infection [112,113].

Consistent with previous reports that Topo1 dysfunction leads to formation of DNA/RNA hybrids (R-loop) during transcription [114], Vpr expression promoted accumulation of R-loops [27]. Furthermore, the overexpression of RNaseH1, which degrades R-loops, reduced DDR activation by Vpr. Although the mechanism of Vpr-induced accumulation of Topo1-cc is unknown, spontaneous generation of Topo1-cc might be augmented by Vpr-induced negative supercoiling of DNA and/or production of reactive oxygen species due to mitochondrial dysfunction. Oxidative modification of DNA traps the Topo1–DNA complex at the ligation step [115].

3c. Histone H2B is a candidate target of Vpr-dependent ubiquitination

We analyzed the mobility of histone H2B in chromatin using GFPtagged histone H2B by a fluorescence recovery after photobleaching assay [27,116]. Expression of wild-type, but not a ubiquitinationdeficient mutant (Q65R), Vpr significantly increased the mobility of histone H2B (Figure 3). Interestingly, Q65R was defective in RPA70 loading, but competent in unwinding dsDNA. Consistently, Vpr induced ubiquitination of histone H2B at lysine 120, a target residue of DDB1/VprBP-dependent modification [27,117]. Finally, treatment with trichostatin A, an inducer of chromatin relaxation, recovered the defective loading of RPA70 in cells expressing the Vpr Q65R mutant [27,118]. These data suggest that histone H2B is a target of Vpr-dependent ubiquitination, and chromatin remodelling by ubiquitination of histone H2B facilitates RPA loading of chromatin.

3d. SXL4 is involved in Vpr-induced DSB, and Topo1 is required for vDNA integration in resting macrophages

Formation of Topo1-cc is followed by activation of SLX4, which resolves the stalled replication fork and removes Topo1-cc via its endonuclease activity [102,103] (Figure 3). The R-loops produced due to dysregulation of Topo1 induce DSBs and the ATM-dependent DDR via nucleotide excision repair [109,110,114,119]. This also occurs in resting macrophages, which do not express ATR/Chk1 [21]. Indeed, downregulation of Topo1 in monocyte-derived macrophages diminished the Vpr-mediated enhancement of viral infectivity [27]. These observations imply that Topo1-cc is related to the effects of Vpr in resting macrophages. Although Vpr induces DSBs and the ATMdependent DDR [8,11,22,27,120], involvement of ATM activation has been ignored from Vpr-induced DDR, because ATR activity in Vpr-induced DDR is much more potent, and Vpr-induced cell-cycle abnormality is attenuated by downregulation of ATR [21]. Current data clearly indicate that DSBs and ATM activation are important for understanding the function of Vpr [27].

Future Perspectives

In the post-ART era, the progression of AIDS can be suppressed, but HIV-1–infected individuals are still susceptible to related various complications, e.g. HAND and malignancies [121-123]. This is due to the existence of cART-refractory sanctuaries that include gutassociated lymphoid tissue, lymph nodes, tissue macrophages and the central nervous system [2,3]. In these sanctuaries, viral replication is abortive but sustained, and several viral proteins, including Vpr, might be continuously produced. Indeed, the serum of patients on cART reportedly contains up to hundreds ng/mL Vpr [124-126]. Notably, soluble Vpr in serum is biologically active and induces retrotransposition of LINE-1 [126]. Thus, Vpr induces DSBs in a manner involving Topo1, and soluble Vpr induces formation of Topo1-cc and DSBs [27].

Topo1-mediated DNA damage could be responsible for several neurodegenerative disorders [127], suggesting Vpr to be involved in HAND development. Additionally, HIV-1–positive patients have a high incidence of lymphomas [121]. Because Vpr in serum induces DSBs in blood cells, it may be a risk factor for the development of malignancies. Interestingly, expression of NKG2D ligands in CD4+ T cells was induced by Vpr [64-66]. Moreover, the DDR upregulates the expression of NKG2D ligands [128], implying that soluble Vpr is involved in perturbation of cell-mediated immunity. The expression of immune-checkpoint molecules, including PD-L1, is upregulated by DNA damage in a ATM/ATR/Chk1-dependent manner [129], supporting a role for Vpr in HIV-associated cancer development. Although Vpr is dispensable for viral integration in proliferating lymphocytes [5-8], its suppression of apoptosis and the immune response could prevent elimination of infected cells and lead to malignant transformation [12,13]. Therefore, Vpr is a target for development of novel anti–HIV-1 agents with the aim of improving the quality of life of HIV-infected individuals.

Acknowledgments

This work was supported in parts by JSPS KAKENHI Grant Number JP26860313, Grants-in-Aid for Research from the National Center for Global Health and Medicine (25A-108) and (21A-129).

Conflict of Interest

The authors declare that they have no conflict of interest

Author Contribution

K.I. and Y.I. wrote the paper. All authors read and approved the final manuscript.

Methodology Used to Search and Select Information

PubMed was almost used to search literatures including term of “Vpr”


References

  1. Trono D, Van Lint C, Rouzioux C, Verdin E, Barré-Sinoussi F, et al. (2010) HIV persistence and the prospect of long-term drug-free remissions for HIV-infected individuals. Science 329: 174–180. [Ref.]
  2. Iglesias-Ussel MD, Romerio F (2011) HIV Reservoirs: The New Frontier. AIDS Rev 13: 13–29. [Ref.]
  3. Abbas W, Tariq M, Iqbal M, Kumar A, Herbein G (2015) Eradication of HIV-1 from the macrophage reservoir: an uncertain goal? Viruses 7: 1578–1598. [Ref.]
  4. Cohen EA, Dehni G, Sodroski JG, Haseltine WA (1990) Human immunodeficiency virus vpr product is a virion-associated regulatory protein. J Virol 64: 3097–3099. [Ref.]
  5. Balliet JW, Kolson DL, Eiger G, Kim FM, McGann KA, et al. (1994) Distinct effects in primary macrophages and lymphocytes of the human immunodeficiency virus type 1 accessory genes vpr, vpu, and nef: mutational analysis of a primary HIV-1 isolate. Virology 200: 623–631. [Ref.]
  6. Connor RI, Chen BK, Choe S, Landau NR (1995) Vpr is required for efficient replication of human immunodeficiency virus type-1 in mononuclear phagocytes. Virology 206: 935–944. [Ref.]
  7. Eckstein DA, Sherman MP, Penn ML, Chin PS, de Noronha CM, et al. (2001) HIV-1 Vpr enhances viral burden by facilitating infection of tissue macrophages but not nondividing CD4+ T cells. J Exp Med 194: 1407–1419. [Ref.]
  8. Koyama T, Sun B, Tokunaga K, Tatsumi M, Ishizaka Y (2013) DNA damage enhances integration of HIV-1 into macrophages by overcoming integrase inhibition. Retrovirology 10: 21. [Ref.]
  9. Zhang S, Pointer D, Singer G, Feng Y, Park K, et al. (1998) Direct binding to nucleic acids by Vpr of human immunodeficiency virus type 1. Gene 212: 157–166. [Ref.]
  10. Morellet N, Bouaziz S, Petitjean P, Roques BP (2003) NMR structure of the HIV-1 regulatory protein VPR. J Mol Biol 327: 215–227. [Ref.]
  11. Tachiwana H, Shimura M, Nakai-Murakami C, Tokunaga K, Takizawa Y, et al. (2006) HIV-1 Vpr induces DNA double-strand breaks. Cancer Res 66: 627–631. [Ref.]
  12. Kogan M, Rappaport J (2011) HIV-1 accessory protein Vpr: relevance in the pathogenesis of HIV and potential for therapeutic intervention. Retrovirology 8: 25. [Ref.]
  13. Guenzel CA, Hérate C, Benichou S (2014) HIV-1 Vpr-a still “enigmatic multitasker”. Front Microbiol 5: 127. [Ref.]
  14. Munir S, Thierry S, Subra F, Deprez E, Delelis O (2013) Quantitative analysis of the time-course of viral DNA forms during the HIV-1 life cycle. Retrovirology 10: 87. [Ref.]
  15. Craigie R, Bushman FD (2012) HIV DNA integration. Cold Spring Harb Perspect Med 2: a006890. [Ref.]
  16. Krishnan L, Engelman A (2012) Retroviral integrase proteins and HIV1 DNA integration. J Biol Chem 287: 40858–40866. [Ref.]
  17. Roshal M, Kim B, Zhu Y, Nghiem P, Planelles V (2003) Activation of the ATR-mediated DNA damage response by the HIV-1 viral protein R. J Biol Chem 278: 25879–25886. [Ref.]
  18. Zimmerman ES, Chen J, Andersen JL, Ardon O, Dehart JL, et al. (2004) Human immunodeficiency virus type 1 Vpr-mediated G2 arrest requires Rad17 and Hus1 and induces nuclear BRCA1 and gamma-H2AX focus formation. Mol Cell Biol 24: 9286–9294. [Ref.]
  19. Groschel B, Bushman F (2005) Cell cycle arrest in G2 /M promotes early steps of infection by human immunodeficiency virus. J Virol 79: 5695–5704. [Ref.]
  20. Lai M, Zimmerman ES, Planelles V, Chen J (2005) Activation of the ATR pathway by human immunodeficiency virus type 1 Vpr involves its direct binding to chromatin in vivo. J Virol 79: 15443–15451. [Ref.]
  21. Zimmerman ES, Sherman MP, Blackett JL, Neidleman JA, Kreis C, et al. (2006) Human immunodeficiency virus type 1 Vpr induces DNA replication stress in vitro and in vivo. J Virol 80: 10407–10418. [Ref.]
  22. Nakai-Murakami C, Shimura M, Kinomoto M, Takizawa Y, Tokunaga K, et al. (2007) HIV-1 Vpr induces ATM-dependent cellular signal with enhanced homologous recombination. Oncogene 26: 477–486. [Ref.]
  23. Blackford AN, Jackson SP (2017) ATM, ATR, and DNA-PK: The Trinity at the Heart of the DNA Damage Response. Mol Cell 66: 801–817. [Ref.]
  24. Sherman MP, de Noronha CM, Williams SA, Greene WC (2002) Insights into the biology of HIV-1 viral protein R. DNA Cell Biol 21: 679–688. [Ref.]
  25. Smith JA, Daniel R (2011) Up-regulation of HIV-1 transduction in nondividing cells by double-strand DNA break-inducing agents. Biotechnol Lett 33: 243–252. [Ref.]
  26. Ebina H, Kanemura Y, Suzuki Y, Urata K, Misawa N, et al. (2012) Integrase-independent HIV-1 infection is augmented under conditions of DNA damage and produces a viral reservoir. Virology 427: 44–50. [Ref.]
  27. Iijima K, Kobayashi J, Ishizaka Y (2018) Structural alteration of DNA induced by viral protein R of HIV-1 triggers the DNA damage response. Retrovirology 15: 8. [Ref.]
  28. Zou L, Elledge SJ (2003) Sensing DNA damage through ATRIP recognition of RPA-ssDNA complexes. Science 300: 1542–1548. [Ref.]
  29. Paxton W, Connor RI, Landau NR (1993) Incorporation of Vpr into human immunodeficiency virus type 1 virions: requirement for the p6 region of gag and mutational analysis. J Virol 67: 7229–7237. [Ref.]
  30. Mansky LM (1996) The mutation rate of human immunodeficiency virus type 1 is influenced by the vpr gene. Virology 222: 391–400. [Ref.]
  31. Mansky LM, Le Rouzic E, Benichou S, Gajary LC (2003) Influence of reverse transcriptase variants, drugs, and Vpr on human immunodeficiency virus type 1 mutant frequencies. J Virol 77: 2071–2080. [Ref.]
  32. Chen R, Le Rouzic E, Kearney JA, Mansky LM, Benichou S (2004) Vprmediated incorporation of UNG2 into HIV-1 particles is required to modulate the virus mutation rate and for replication in macrophages. J Biol Chem 279: 28419–28425. [Ref.]
  33. Schröfelbauer B, Yu Q, Zeitlin SG, Landau NR (2005) Human immunodeficiency virus type 1 Vpr induces the degradation of the UNG and SMUG uracil-DNA glycosylases. J Virol 79: 10978–10987. [Ref.]
  34. Yang B, Chen K, Zhang C, Huang S, Zhang H (2007) Virion-associated uracil DNA glycosylase-2 and apurinic/apyrimidinic endonuclease are involved in the degradation of APOBEC3G-edited nascent HIV-1 DNA. J Biol Chem 282: 11667–11675. [Ref.]
  35. Guenzel CA, Hérate C, Le Rouzic E, Maidou-Peindara P, Sadler HA, et al. (2012) Recruitment of the nuclear form of uracil DNA glycosylase into virus particles participates in the full infectivity of HIV-1. J Virol 86: 2533–2544. [Ref.]
  36. Gleenberg IO, Herschhorn A, Hizi A (2007) Inhibition of the activities of reverse transcriptase and integrase of human immunodeficiency virus type-1 by peptides derived from the homologous viral protein R (Vpr). J Mol Biol 369: 1230–1243. [Ref.]
  37. Lyonnais S, Gorelick RJ, Heniche-Boukhalfa F, Bouaziz S, Parissi V, et al. (2013) A protein ballet around the viral genome orchestrated by HIV-1 reverse transcriptase leads to an architectural switch: from nucleocapsid-condensed RNA to Vpr-bridged DNA. Virus Res 171: 287–303. [Ref.]
  38. Popov S, Rexach M, Zybarth G, Reiling N, Lee MA, et al. (1998) Viral protein R regulates nuclear import of the HIV-1 pre-integration complex. EMBO J 17: 909–917. [Ref.]
  39. Nitahara-Kasahara Y, Kamata M, Yamamoto T, Zhang X, Miyamoto Y, et al. (2007) Novel nuclear import of Vpr promoted by importin alpha is crucial for human immunodeficiency virus type 1 replication in macrophages. J Virol 81: 5284–5293. [Ref.]
  40. Sherman MP, de Noronha CM, Eckstein LA, Hataye J, Mundt P, et al. (2003) Nuclear export of Vpr is required for efficient replication of human immunodeficiency virus type 1 in tissue macrophages. J Virol 77: 7582–7589. [Ref.]
  41. Fouchier RA, Meyer BE, Simon JH, Fischer U, Albright AV, et al. (1998) Interaction of the human immunodeficiency virus type 1 Vpr protein with the nuclear pore complex. J Virol 72: 6004–6013. [Ref.]
  42. Le Rouzic E, Mousnier A, Rustum C, Stutz F, Hallberg E, et al. (2002) Docking of HIV-1 Vpr to the nuclear envelope is mediated by the interaction with the nucleoporin hCG1. J Biol Chem 277: 45091– 45098. [Ref.]
  43. Marini B, Kertesz-Farkas A, Ali H, Lucic B, Lisek K, et al. (2015) Nuclear architecture dictates HIV-1 integration site selection. Nature 521: 227–231. [Ref.]
  44. Lelek M, Casartelli N, Pellin D, Rizzi E, Souque P, et al. (2015) Chromatin organization at the nuclear pore favours HIV replication. Nat Commun 6: 6483. [Ref.]
  45. Demeulemeester J, De Rijck J, Gijsbers R, Debyser Z (2015) Retroviral integration: Site matters: Mechanisms and consequences of retroviral integration site selection. Bioessays 37: 1202–1214. [Ref.]
  46. Kilareski EM, Shah S, Nonnemacher MR, Wigdahl B (2009) Regulation of HIV-1 transcription in cells of the monocyte-macrophage lineage. Retrovirology 6: 118. [Ref.]
  47. Wang L, Mukherjee S, Jia F, Narayan O, Zhao LJ (1995) Interaction of virion protein Vpr of human immunodeficiency virus type 1 with cellular transcription factor Sp1 and trans-activation of viral long terminal repeat. J Biol Chem 270: 25564–25569. [Ref.]
  48. Sawaya BE, Khalili K, Mercer WE, Denisova L, Amini S (1998) Cooperative actions of HIV-1 Vpr and p53 modulate viral gene transcription. J Biol Chem 273: 20052–20057. [Ref.]
  49. Sherman MP, de Noronha CM, Pearce D, Greene WC (2000) Human immunodeficiency virus type 1 Vpr contains two leucine-rich helices that mediate glucocorticoid receptor coactivation independently of its effects on G(2) cell cycle arrest. J Virol 74: 8159–8165. [Ref.]
  50. Kino T, Gragerov A, Slobodskaya O, Tsopanomichalou M, Chrousos GP, et al. (2002) Human immunodeficiency virus type 1 (HIV1) accessory protein Vpr induces transcription of the HIV-1 and glucocorticoid-responsive promoters by binding directly to p300/ CBP coactivators. J Virol 76: 9724–9734. [Ref.]
  51. Muthumani K, Choo AY, Zong WX, Madesh M, Hwang DS, et al. (2006) The HIV-1 Vpr and glucocorticoid receptor complex is a gainof-function interaction that prevents the nuclear localization of PARP-1. Nat Cell Biol 8: 170–179. [Ref.]
  52. Liu R, Lin Y, Jia R, Geng Y, Liang C, et al. (2014) HIV-1 Vpr stimulates NF-κB and AP-1 signaling by activating TAK1. Retrovirology 11: 45. [Ref.]
  53. Liang Z, Liu R, Lin Y, Liang C, Tan J, et al. (2015) HIV-1 Vpr protein activates the NF-κB pathway to promote G2/M cell cycle arrest. Virol Sin 30: 441–448. [Ref.]
  54. Hoshino S, Konishi M, Mori M, Shimura M, Nishitani C, et al. (2010) HIV-1 Vpr induces TLR4/MyD88-mediated IL-6 production and reactivates viral production from latency. J Leukoc Biol 87: 1133– 1143. [Ref.]
  55. Romani B, Baygloo NS, Hamidi-Fard M, Aghasadeghi MR, Allahbakhshi E (2016) HIV-1 Vpr Protein Induces Proteasomal Degradation of Chromatin-associated Class I HDACs to Overcome Latent Infection of Macrophages. J Biol Chem 291: 2696–2711. [Ref.]
  56. Romani B, Kamali Jamil R, Hamidi-Fard M, Rahimi P, Momen SB, et al. (2016) HIV-1 Vpr reactivates latent HIV-1 provirus by inducing depletion of class I HDACs on chromatin. Sci Rep 6: 31924. [Ref.]
  57. Sawaya BE, Khalili K, Gordon J, Taube R, Amini S (2000) Cooperative interaction between HIV-1 regulatory proteins Tat and Vpr modulates transcription of the viral genome. J Biol Chem 275: 35209–35214. [Ref.]
  58. Poon B, Chang MA, Chen IS (2007) Vpr is required for efficient Nef expression from unintegrated human immunodeficiency virus type 1 DNA. J Virol 81: 10515–10523. [Ref.]
  59. Jacotot E, Ravagnan L, Loeffler M, Ferri KF, Vieira HL, et al. (2000) The HIV-1 viral protein R induces apoptosis via a direct effect on the mitochondrial permeability transition pore. J Exp Med 191: 33–46. [Ref.]
  60. Andersen JL, DeHart JL, Zimmerman ES, Ardon O, Kim B, et al. (2006) HIV-1 Vpr-induced apoptosis is cell cycle dependent and requires Bax but not ANT. PLoS Pathog 2: e127. [Ref.]
  61. Andersen JL, Zimmerman ES, DeHart JL, Murala S, Ardon O, et al. (2005) ATR and GADD45alpha mediate HIV-1 Vpr-induced apoptosis. Cell Death Differ 12: 326–334. [Ref.]
  62. Conti L, Rainaldi G, Matarrese P, Varano B, Rivabene R, et al. (1998) The HIV-1 vpr protein acts as a negative regulator of apoptosis in a human lymphoblastoid T cell line: possible implications for the pathogenesis of AIDS. J Exp Med 187: 403–413. [Ref.]
  63. Majumder B, Venkatachari NJ, Schafer EA, Janket ML, Ayyavoo V (2007) Dendritic cells infected with vpr-positive human immunodeficiency virus type 1 induce CD8+ T-cell apoptosis via upregulation of tumor necrosis factor alpha. J Virol 81: 7388–7399. [Ref.]
  64. Ward J, Davis Z, DeHart J, Zimmerman E, Bosque A, et al. (2009) HIV-1 Vpr triggers natural killer cell-mediated lysis of infected cells through activation of the ATR-mediated DNA damage response. PLoS Pathog 5: e1000613. [Ref.]
  65. Richard J, Sindhu S, Pham TN, Belzile JP, Cohen EA (2010) HIV-1 Vpr up-regulates expression of ligands for the activating NKG2D receptor and promotes NK cell-mediated killing. Blood 115: 1354–1363. [Ref.]
  66. Richard J, Pham TN, Ishizaka Y, Cohen EA (2013) Viral protein R upregulates expression of ULBP2 on uninfected bystander cells during HIV-1 infection of primary CD4+ T lymphocytes. Virology 443: 248–256. [Ref.]
  67. Majumder B, Venkatachari NJ, O’Leary S, Ayyavoo V (2008) Infection with Vpr-positive human immunodeficiency virus type 1 impairs NK cell function indirectly through cytokine dysregulation of infected target cells. J Virol 82: 7189–7200. [Ref.]
  68. Sowrirajan B, Barker E (2011) The natural killer cell cytotoxic function is modulated by HIV-1 accessory proteins. Viruses 3: 1091–1111. [Ref.]
  69. Zahoor MA, Xue G, Sato H, Murakami T, Takeshima SN, et al. (2014) HIV-1 Vpr induces interferon-stimulated genes in human monocytederived macrophages. PLoS One 9: e106418. [Ref.]
  70. Zahoor MA, Xue G, Sato H, Aida Y (2015) Genome-wide transcriptional profiling reveals that HIV-1 Vpr differentially regulates interferonstimulated genes in human monocyte-derived dendritic cells. Virus Res 208: 156–163. [Ref.]
  71. Vermeire J, Roesch F, Sauter D, Rua R, Hotter D, et al. (2016) HIV Triggers a cGAS-Dependent, Vpu- and Vpr-Regulated Type I Interferon Response in CD4+ T Cells. Cell Rep 17: 413–424. [Ref.]
  72. Doi A, Iijima K, Kano S, Ishizaka Y (2015) Viral protein R of HIV type1 induces retrotransposition and upregulates glutamate synthesis by the signal transducer and activator of transcription 1 signaling pathway. Microbiol Immunol 59: 398–409. [Ref.]
  73. Barrero CA, Datta PK, Sen S, Deshmane S, Amini S, et al. (2013) HIV1 Vpr modulates macrophage metabolic pathways: a SILAC-based quantitative analysis. PLoS One 8: e68376. [Ref.]
  74. Datta PK, Deshmane S, Khalili K, Merali S, Gordon JC, et al. (2016) Glutamate metabolism in HIV-1 infected macrophages: Role of HIV-1 Vpr. Cell Cycle 15: 2288–2298. [Ref.]
  75. Jones GJ, Barsby NL, Cohen EA, Holden J, Harris K, et al. (2007) HIV1 Vpr causes neuronal apoptosis and in vivo neurodegeneration. J Neurosci 27: 3703–3711. [Ref.]
  76. Power C, Hui E, Vivithanaporn P, Acharjee S, Polyak M (2012) Delineating HIV-associated neurocognitive disorders using transgenic models: the neuropathogenic actions of Vpr. J Neuroimmune Pharmacol 7: 319–331. [Ref.]
  77. Gangwani MR, Noel RJ Jr, Shah A, Rivera-Amill V, Kumar A (2013) Human immunodeficiency virus type 1 viral protein R (Vpr) induces CCL5 expression in astrocytes via PI3K and MAPK signaling pathways. J Neuroinflammation 10: 136. [Ref.]
  78. Gangwani MR, Kumar A (2015) Multiple Protein Kinases via Activation of Transcription Factors NF-κB, AP-1 and C/EBP-δ Regulate the IL-6/ IL-8 Production by HIV-1 Vpr in Astrocytes. PLoS One 10: e0135633. [Ref.]
  79. Boyd S, Akpamagbo Y, Kashanchi F (2016) HIV Vpr controls CNS metabolism. Cell Cycle 15: 2389–2390. [Ref.]
  80. Dampier W, Antell GC, Aiamkitsumrit B, Nonnemacher MR, Jacobson JM, et al. (2017) Specific amino acids in HIV-1 Vpr are significantly associated with differences in patient neurocognitive status. J Neurovirol 23: 113–124. [Ref.]
  81. Daniel R, Katz RA, Skalka AM (1999) A role for DNA-PK in retroviral DNA integration. Science 284: 644–647. [Ref.]
  82. Daniel R, Ramcharan J, Rogakou E, Taganov KD, Greger JG, et al. (2004) Histone H2AX is phosphorylated at sites of retroviral DNA integration but is dispensable for postintegration repair. J Biol Chem 279: 45810–45814. [Ref.]
  83. Faust EA, Triller H (2002) Stimulation of human flap endonuclease 1 by human immunodeficiency virus type 1 integrase: possible role for flap endonuclease 1 in 5’-end processing of human immunodeficiency virus type 1 integration intermediates. J Biomed Sci 9: 273–287. [Ref.]
  84. Daniel R, Greger JG, Katz RA, Taganov KD, Wu X, et al. (2004) Evidence that stable retroviral transduction and cell survival following DNA integration depend on components of the nonhomologous end joining repair pathway. J Virol 78: 8573–8581. [Ref.]
  85. Lau A, Swinbank KM, Ahmed PS, Taylor DL, Jackson SP, et al. (2005) Suppression of HIV-1 infection by a small molecule inhibitor of the ATM kinase. Nat Cell Biol 7: 493–500. [Ref.]
  86. Rahal EA, Henricksen LA, Li Y, Williams RS, Tainer JA, et al. (2010) ATM regulates Mre11-dependent DNA end-degradation and microhomology-mediated end joining. Cell Cycle 9: 2866–2877. [Ref.]
  87. Sakai K, Dimas J, Lenardo MJ (2006) The Vif and Vpr accessory proteins independently cause HIV-1-induced T cell cytopathicity and cell cycle arrest. Proc Natl Acad Sci USA 103: 3369–3374. [Ref.]
  88. He J, Choe S, Walker R, Di Marzio P, Morgan DO, et al. (1995) Human immunodeficiency virus type 1 viral protein R (Vpr) arrests cells in the G2 phase of the cell cycle by inhibiting p34cdc2 activity. J Virol 69: 6705–6711. [Ref.]
  89. Laguette N, Brégnard C, Hue P, Basbous J, Yatim A, et al. (2014) Premature activation of the SLX4 complex by Vpr promotes G2/M arrest and escape from innate immune sensing. Cell 156: 134–145. [Ref.]
  90. DeHart JL, Zimmerman ES, Ardon O, Monteiro-Filho CM, Argañaraz ER, et al. (2007) HIV-1 Vpr activates the G2 checkpoint through manipulation of the ubiquitin proteasome system. Virol J 4: 57. [Ref.]
  91. Belzile JP, Duisit G, Rougeau N, Mercier J, Finzi A, et al. (2007) HIV-1 Vpr-mediated G2 arrest involves the DDB1-CUL4AVPRBP E3 ubiquitin ligase. PLoS Pathog 3: e85. [Ref.]
  92. Tan L, Ehrlich E, Yu XF (2007) DDB1 and Cul4A are required for human immunodeficiency virus type 1 Vpr-induced G2 arrest. J Virol 81: 10822–10830. [Ref.]
  93. Belzile JP, Richard J, Rougeau N, Xiao Y, Cohen EA (2010) HIV-1 Vpr induces the K48-linked polyubiquitination and proteasomal degradation of target cellular proteins to activate ATR and promote G2 arrest. J Virol 84: 3320–3330. [Ref.]
  94. Eldin P, Chazal N, Fenard D, Bernard E, Guichou JF, et al. (2014) Vpr expression abolishes the capacity of HIV-1 infected cells to repair uracilated DNA. Nucleic Acids Res 42: 1698–1710. [Ref.]
  95. Lahouassa H, Blondot ML, Chauveau L, Chougui G, Morel M, et al. (2016) HIV-1 Vpr degrades the HLTF DNA translocase in T cells and macrophages. Proc Natl Acad Sci USA 113: 5311–5316. [Ref.]
  96. Klockow LC, Sharifi HJ, Wen X, Flagg M, Furuya AK, et al. (2013) The HIV-1 protein Vpr targets the endoribonuclease Dicer for proteasomal degradation to boost macrophage infection. Virology 444: 191–202. [Ref.]
  97. Francia S, Michelini F, Saxena A, Tang D, de Hoon M, et al. (2012) Site-specific DICER and DROSHA RNA products control the DNAdamage response. Nature 488: 231–235. [Ref.]
  98. Francia S, Cabrini M, Matti V, Oldani A, d’Adda di Fagagna F (2016) DICER, DROSHA and DNA damage response RNAs are necessary for the secondary recruitment of DNA damage response factors. J Cell Sci 129: 1468–1476. [Ref.]
  99. Burger K, Schlackow M, Potts M, Hester S, Mohammed S, et al. (2017) Nuclear phosphorylated Dicer processes double-stranded RNA in response to DNA damage. J Cell Biol 216: 2373–2389. [Ref.]
  100. Wang X, Singh S, Jung HY, Yang G, Jun S, et al. (2013) HIV-1 Vpr protein inhibits telomerase activity via the EDD-DDB1-VPRBP E3 ligase complex. J Biol Chem 288: 15474–15480. [Ref.]
  101. Shimura M, Onozuka Y, Yamaguchi T, Hatake K, Takaku F, et al. (1999) Micronuclei formation with chromosome breaks and gene amplification caused by Vpr, an accessory gene of human immunodeficiency virus. Cancer Res 59: 2259–2264. [Ref.]
  102. Deng C, Brown JA, You D, Brown JM (2005) Multiple endonucleases function to repair covalent topoisomerase I complexes in Saccharomyces cerevisiae. Genetics 170: 591–600. [Ref.]
  103. Kim Y, Spitz GS, Veturi U, Lach FP, Auerbach AD, et al. (2013) Regulation of multiple DNA repair pathways by the Fanconi anemia protein SLX4. Blood 121: 54–63. [Ref.]
  104. Romani B, Shaykh Baygloo N, Aghasadeghi MR, Allahbakhshi E (2015) HIV-1 Vpr Protein Enhances Proteasomal Degradation of MCM10 DNA Replication Factor through the Cul4-DDB1[VprBP] E3 Ubiquitin Ligase to Induce G2/M Cell Cycle Arrest. J Biol Chem 290: 17380–17389. [Ref.]
  105. Janicki SM, Tsukamoto T, Salghetti SE, Tansey WP, Sachidanandam R, et al. (2004) From silencing to gene expression: real-time analysis in single cells. Cell 116: 683–698. [Ref.]
  106. Piccinno R, Cipinska M, Roukos V (2017) Studies of the DNA Damage Response by Using the Lac Operator/Repressor (LacO/LacR) Tethering System. Methods Mol Biol 1599: 263–275. [Ref.]
  107. Naughton C, Avlonitis N, Corless S, Prendergast JG, Mati IK, et al. (2013) Transcription forms and remodels supercoiling domains unfolding large-scale chromatin structures. Nat Struct Mol Biol 20: 387–395. [Ref.]
  108. Anders L, Guenther MG, Qi J, Fan ZP, Marineau JJ, et al. (2014) Genome-wide localization of small molecules. Nat Biotechnol 32: 92–96. [Ref.]
  109. Pommier Y, Sun Y, Huang SN, Nitiss JL (2016) Roles of eukaryotic topoisomerases in transcription, replication and genomic stability. Nat Rev Mol Cell Biol 17: 703–721. [Ref.]
  110. Sordet O, Redon CE, Guirouilh-Barbat J, Smith S, Solier S, et al. (2009) Ataxia telangiectasia mutated activation by transcriptionand topoisomerase I-induced DNA double-strand breaks. EMBO Rep 10: 887–893. [Ref.]
  111. Shimura M, Toyoda Y, Iijima K, Kinomoto M, Tokunaga K, et al. (2011) Epigenetic displacement of HP1 from heterochromatin by HIV-1 Vpr causes premature sister chromatid separation. J Cell Biol 194: 721–735. [Ref.]
  112. Jordan A, Bisgrove D, Verdin E (2003) HIV reproducibly establishes a latent infection after acute infection of T cells in vitro. EMBO J 22: 1868–1877. [Ref.]
  113. Kameoka M, Nukuzuma S, Itaya A, Tanaka Y, Ota K, et al. (2005) Poly(ADP-ribose)polymerase-1 is required for integration of the human immunodeficiency virus type 1 genome near centromeric alphoid DNA in human and murine cells. Biochem Biophys Res Commun 334: 412–417. [Ref.]
  114. Sollier J, Cimprich KA (2015) Breaking bad: R-loops and genome integrity. Trends Cell Biol 25: 514–522. [Ref.]
  115. Pommier Y, Barcelo JM, Rao VA, Sordet O, Jobson AG, et al. (2006) Repair of topoisomerase I-mediated DNA damage. Prog Nucleic Acid Res Mol Biol 81: 179–229. [Ref.]
  116. Kimura H, Cook PR (2001) Kinetics of core histones in living human cells: little exchange of H3 and H4 and some rapid exchange of H2B. J Cell Biol 153: 1341–1353. [Ref.]
  117. Zeng M, Ren L, Mizuno K, Nestoras K, Wang H, et al. (2016) CRL4(Wdr70) regulates H2B monoubiquitination and facilitates Exo1-dependent resection. Nat Commun 7: 11364. [Ref.]
  118. Tóth KF, Knoch TA, Wachsmuth M, Frank-Stöhr M, Stöhr M, et al. (2004) Trichostatin A-induced histone acetylation causes decondensation of interphase chromatin. J Cell Sci 117: 4277–4287. [Ref.]
  119. Sollier J, Stork CT, García-Rubio ML, Paulsen RD, Aguilera A, et al. (2014) Transcription-coupled nucleotide excision repair factors promote R-loop-induced genome instability. Mol Cell 56: 777–785. [Ref.]
  120. Taneichi D, Iijima K, Doi A, Koyama T, Minemoto Y, et al. (2011) Identification of SNF2h, a chromatin-remodeling factor, as a novel binding protein of Vpr of human immunodeficiency virus type 1. J. Neuroimmune Pharmacol 6: 177–187. [Ref.]
  121. Shiels MS, Pfeiffer RM, Gail MH, Hall HI, Li J, et al. (2011) Cancer burden in the HIV-infected population in the United States. J Natl Cancer Inst 103: 753–762. [Ref.]
  122. Kaplan LD (2012) HIV-associated lymphoma. Best Pract Res Clin Haematol 25: 101–117. [Ref.]
  123. Saylor D, Dickens AM, Sacktor N, Haughey N, Slusher B, et al. (2016) HIV-associated neurocognitive disorder--pathogenesis and prospects for treatment. Nat Rev Neurol 12: 234–248. [Ref.]
  124. Levy DN, Refaeli Y, MacGregor RR, Weiner DB. (1994) Serum Vpr regulates productive infection and latency of human immunodeficiency virus type 1. Proc Natl Acad Sci USA 91: 10873– 10877. [Ref.]
  125. Hoshino S, Sun B, Konishi M, Shimura M, Segawa T, et al. (2007) Vpr in plasma of HIV type 1-positive patients is correlated with the HIV type 1 RNA titers. AIDS Res Hum Retroviruses 23: 391–397. [Ref.]
  126. Iijima K, Okudaira N, Tamura M, Doi A, Saito Y, et al. (2013) Viral protein R of human immunodeficiency virus type-1 induces retrotransposition of long interspersed element-1. Retrovirology 10: 83. [Ref.]
  127. Katyal S, Lee Y, Nitiss KC, Downing SM, Li Y, et al. (2014) Aberrant topoisomerase-1 DNA lesions are pathogenic in neurodegenerative genome instability syndromes. Nat Neurosci 17: 813–821. [Ref.]
  128. Gasser S, Orsulic S, Brown EJ, Raulet DH (2005) The DNA damage pathway regulates innate immune system ligands of the NKG2D receptor. Nature 436: 1186–1190. [Ref.]
  129. Sato H, Niimi A, Yasuhara T, Permata TBM, Hagiwara Y, et al. (2017) DNA double-strand break repair pathway regulates PD-L1 expression in cancer cells. Nat Commun 8: 1751. [Ref.]

Download Provisional PDF Here

 

Article Information

Aritcle Type: REVIEW ARTICLE

Citation: Iijima K, Ishizaka Y (2018) DNA unwinding by Viral Protein R Initializes Complicated Cellular Responses in HIV-1 Infection: Defining the Viper’s First Bite. J Emerg Dis Virol 4(1): dx.doi.org/10.16966/2473-1846.141

Copyright: © 2018 Iijima K, et al. This is an open-access article distributed under the terms of the Creative Commons Attribution License, which permits unrestricted use, distribution, and reproduction in any medium, provided the original author and source are credited.

Publication history: 

  • Received date: 29 Jun, 2018

  • Accepted date: 17 Jul, 2018

  • Published date: 23 Jul, 2018