Next Article in Journal
Structural Impact of the Interaction of the Influenza A Virus Nucleoprotein with Genomic RNA Segments
Next Article in Special Issue
The ISG15-Protease USP18 Is a Pleiotropic Enhancer of HIV-1 Replication
Previous Article in Journal
Herpes Simplex Virus Bronchopneumonitis in Critically Ill Patients with Acute on Chronic Liver Failure: A Retrospective Analysis
Previous Article in Special Issue
IFNα Subtypes in HIV Infection and Immunity
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

HIV-1 Vpr Functions in Primary CD4+ T Cells

by
Carlos Alberto Vanegas-Torres
and
Michael Schindler
*
Institute for Medical Virology and Epidemiology of Viral Diseases, University Hospital Tuebingen, 72076 Tuebingen, Germany
*
Author to whom correspondence should be addressed.
Submission received: 31 January 2024 / Revised: 5 March 2024 / Accepted: 5 March 2024 / Published: 9 March 2024
(This article belongs to the Special Issue Innate Sensing and Restriction of Retroviruses)

Abstract

:
HIV-1 encodes four accesory proteins in addition to its structural and regulatory genes. Uniquely amongst them, Vpr is abundantly present within virions, meaning it is poised to exert various biological effects on the host cell upon delivery. In this way, Vpr contributes towards the establishment of a successful infection, as evidenced by the extent to which HIV-1 depends on this factor to achieve full pathogenicity in vivo. Although HIV infects various cell types in the host organism, CD4+ T cells are preferentially targeted since they are highly permissive towards productive infection, concomitantly bringing about the hallmark immune dysfunction that accompanies HIV-1 spread. The last several decades have seen unprecedented progress in unraveling the activities Vpr possesses in the host cell at the molecular scale, increasingly underscoring the importance of this viral component. Nevertheless, it remains controversial whether some of these advances bear in vivo relevance, since commonly employed cellular models significantly differ from primary T lymphocytes. One prominent example is the “established” ability of Vpr to induce G2 cell cycle arrest, with enigmatic physiological relevance in infected primary T lymphocytes. The objective of this review is to present these discoveries in their biological context to illustrate the mechanisms whereby Vpr supports HIV-1 infection in CD4+ T cells, whilst identifying findings that require validation in physiologically relevant models.

1. Introduction

Similarly to all other retroviruses, lentiviruses, including HIV-1, exhibit the prototypical gag, pol, and env genes, encoding the structural, catalytic, and binding elements of their viral progeny, respectively. Additionally, expression of the two essential regulatory proteins Tat and Rev ensures the correct spatiotemporal transcription and production of viral proteins during the viral life cycle. However, a hallmark of HIV-1 and other lentiviruses is the presence of accessory genes (nef, vif, vpu, and vpx/vpr) [1,2]. These refer to different virulence factors and they are termed “accessory” proteins, given that they have been described as largely dispensable for infection and viral replication in most cell lines but important for full in vivo pathogenicity, high viral loads, and persistence [3]. Although accessory proteins do not exert any known inherent catalytic activity, they collectively excel at counteracting restriction factors (RFs), hijacking metabolic pathways, and mediating immune evasion, consequently contributing towards successful viral replication, spread, and transmission. All this is achieved as a result of their capacity to engage in highly pleiotropic interactions with a diverse range of host cell components [1,4].
One such factor is Vpr: this 14 kDa, 96 aa protein presents an unstructured N-terminal domain followed by three amphipathic α-helices. Vpr is the HIV-1 accessory protein with the highest abundance in the viral particles, as it specifically interacts with the structural peptide p6 and is therefore encapsidated into the assembling viral particle. As a consequence, it is present in de novo infected cells after uncoating and thought to exert multiple biological effects directly thereupon [5,6]. Its tertiary structure is stabilized by a lipophilic core composed of several sterically convergent hydrophobic residues from helices α1 and α3 [7]. These three helical motifs together comprise the contact sites Vpr employs to recruit protein targets, whilst the unstructured domain is involved in the stabilization of various such contacts, as well as the multimerization of this protein in mature virions [7,8]. Despite extensive efforts being directed towards unveiling the precise role(s) Vpr plays in subverting the host’s physiology, no defined clear functions have been found to explain why vpr is maintained within the lentiviral genome and why it is apparently essential for the virus in vivo [4]. This stands in contrast to virtually all other viral accessory proteins, which possess defined “main” functions with major measurable phenotypes in certain cell culture systems, despite their multifunctionality. For instance, Vif counteracts APOBEC-family proteins [9], Vpu antagonizes Tetherin [10], and Vpx degrades SAMHD1 [11] and hijacks the Human Silencing Hub (HUSH) complex [12], whilst Nef interferes with CD4, MHC-I, and SERINC5 [13].
Vpr is instead implicated in disrupting multiple cellular processes in a rather nuanced manner, with various unclear biological phenotypes resulting from a large-scale remodeling of the host cell’s protein landscape [5,14]. Although the following sections will discuss the precise means through which Vpr is able to induce the dysregulation of numerous host cell factors, it is essential to clarify the established fact that this phenomenon entails through the redirection of the proteasomal degradation pathway. Under normal circumstances, DDB1, CRL4, and DCAF1 assemble into a E3 ubiquitin ligase complex capable of marking a bevy of proteins for degradation. The substrate specificity of this complex is dictated by the adaptor protein DCAF1, a multifunctional cellular factor with an additional serine/threonine kinase activity canonically involved in the regulation of various processes, such as cell cycle progression, histone modification [15], and telomerase regulation [16]. In the context of HIV-1 infection, however, Vpr has been shown to directly bind DCAF1, earning it the name “Vpr-binding protein” (VprBP). This tripartite interaction is based on the crosstalk between various regions of DCAF1’s β-propeller and Vpr’s α1- and α3-helices, with its unstructured domain further locking this contact in place. By virtue of this, Vpr can significantly influence the manner in which DCAF1 acquires its targets, thereby contributing towards reshaping the intracellular environment in favor of HIV-1 [17].
As such, this interaction is also implied to play a key role in one of Vpr’s most accepted phenotypic functions, that is, to arrest cells in G2. Nevertheless, for G2 arrest itself, neither the functional relevance nor the exact mechanism is clearly established, and compelling evidence for this happening in infected primary CD4+ T cells is not widely available; details of the latter will be discussed in the following sections. Indeed, a great deal of confusion in the Vpr field may originate from the fact that a large proportion of data stem from studies employing artificial cell lines, non-relevant models, or in vitro cell-free systems, much thereof not having yet been confirmed in a cellular context. Hence, the main objective of the present review is to summarize Vpr’s known multifaceted activities along various steps of the viral life cycle, emphasizing the evidence supporting these functions in the context of primary CD4+ T cells. Importantly, we will explicitly not delve deep into the relevance of Vpr in macrophages, which are the second most relevant type of primary target cell for HIV-1 in vivo.

2. Reverse Transcription and Nuclear Delivery of the Pre-Integration Complex

Upon host cell entry, reverse transcription of the viral genome is one crucial event to take place in the replication cycle of HIV. In order to generate a dsDNA proviral genome, the virion-packaged reverse transcriptase (RT) synthesizes a continuous (-)ssDNA strand using the (+)ssRNA viral genome as a template. However, assembling a positive-sensed strand demands the presence of RNA primers and consequently delivers a discontinuous DNA molecule featuring gaps, flaps, and tertiary structures reminiscent of the lagging strand generated during chromosomal DNA replication [18]. Already in this step, Vpr plays an important role, as a 1998 in vitro study (by the term in vitro we refer herein to cell-free systems) demonstrated its capability to sequester Lysyl RNA synthetase (LysRS) [19], thus preventing the aminoacylation of tRNALys3 that is later packaged into viral particles and serves as a primer for RT to initiate highly processive reverse transcription upon host cell delivery [20].
Patching the plus-strand into a continuous molecule is something the error-prone RT is unable to perform by itself, and Vpr has also been proven to hijack a variety of DNA repair enzymes to compensate for this. A clear example is the Vpr-mediated repurposing of UNG2 and the RPA2 subunit of the DNA damage response (DDR)-associated replication protein A complex, which together lead to the repair of abasic or APOBEC3G-deaminated sites in the said strand, reducing the overall mutation rate of reverse transcription [7,21,22]. Notably, UNG2 is a known interaction partner of the aforementioned DCAF1 protein [7,23,24,25], and to further assist this process, Vpr is thought to simultaneously induce the proteasomal degradation of various host cell factors by virtue of its ability to hijack the CRL4-DDB1-DCAF1 E3 ubiquitin ligase complex [7]; such is the case with Exonuclease 1 (Exo1) and the HLTF helicase, which would otherwise excise newly synthesized proviral DNA and catalyze the reversal and dissociation of the replication fork, respectively [26,27,28].
Concomitantly with this step, it is nowadays accepted that the intact viral capsid containing the PIC ferries its dsDNA payload along the host’s cytoskeleton into the nucleus [29]. Indeed, the protein shell of the pre-integration complex (PIC) was initially thought to act as a protective measure against restriction factors like TRIM5α [30] or MX2 [31]. Other studies, in line with growing evidence of capsid uncoating happening in the nucleus, underscore the fact that the HIV-1 capsid protein lattice prevents the recognition of viral nucleic acids by cytosolic pattern recognition receptors [31]. Vpr itself may also harbor a docking function by interacting with the nuclear pore complex, thereby supporting nuclear capsid translocation [32]. During nuclear entry, architectonical remodeling of the PIC may occur [33], where this complex eventually undergoes uncoating and so allows the preferential integration of the provirus into H3K36me3-rich, actively transcribed chromatin [34]. Aiding the integration process is one of Vpr’s first described functions, and its presence in the nucleus seems crucial for the establishment of a productive infection in CD4+ T cells [35]. Despite the fact that it does not possess a classical nuclear localization signal (NLS) in its structure, Vpr nonetheless exhibits a highly karyophilic character, a characteristic attributed to the LxxLL motifs found in α-helices 1 and 3 [36]. Vpr has been demonstrated to bind karyopherin-α in vitro and thereby increase its affinity for basic NLSs such as the ones found in the C-termini of the integrase and matrix proteins of the PIC [37]. By doing so, it would facilitate the FEZ1-mediated acceleration of the PIC’s nuclear import whilst diminishing the likelihood of RFs damaging the PIC [38], although these findings may necessitate refinement in light of the fact that intact viral capsids can enter the nucleus [29].
Given that Vpr is specifically incorporated into assembled viral particles and thus exerts a biological role upon de novo infection, it is clear that the presence of Vpr plays a pivotal role early in the HIV life cycle; not only may it support viral genome integrity before, during, and after reverse transcription, but it may also simultaneously facilitate its journey into the host’s nucleus. Importantly, these findings are conceptually compatible with the current paradigm change, which states that uncoating does not take place directly after viral fusion with the cellular plasma membrane, but rather during or after nuclear import [29,39]. In line with this, Vpr was shown to rapidly traverse to the nucleus after viral entry, although this has thus far only been established in cell lines [40].
What has been proven in primary CD4+ T cells is that Vpr degrades the aforementioned host cell factors, i.e., Exo1 and HLTF in a DCAF1-dependent manner [5]; whether Vpr selectively depletes the tRNA pool in infected primary lymphocytes remains to be confirmed, just inasmuch as it does indeed aid the processes of NPC docking of viral capsids and PIC nuclear import. Furthermore, even though Exo1 and HLTF have been shown to be degraded by Vpr in primary T cells, evidence for their functional involvement in G2 cell cycle arrest stems from immortalized cell lines. Factually, this is the most frequent level of experimental evidence available when Vpr functions are biologically connected to their supposed role in inducing G2 arrest, which is discussed in detail in the following section.

3. Viral Genome Transcription and G2 Cell Cycle Arrest

The work of Soto and colleagues highlighted how determinant the activation status of CD4+ T cells can be towards the outcome of infection; lightly activated subsets are poised to serve as reservoirs for HIV to establish latency upon returning to an inactive phenotype, whereas actively proliferating lymphocytes exhibit a fruitful but short-lived burst in viral production, culminating in apoptosis [41]. Contrastingly, unstimulated lymphocytes were long regarded as inconsequential regarding viral persistence, as HIV disfavors the establishment of infection therein [41,42], although novel data have shed light on the importance this subpopulation exhibits towards latency development. Indeed, naïve CD4+ T cells were observed not only to possess unintegrated proviruses in a variety of extrachromosomal configurations [43,44], but the extent of their infection was found to strongly predict reservoir size and diversity, given that this subset can repopulate immune compartments upon depletion of central memory CD4+ T cells [45,46]. Unintegrated proviruses could gradually decay and not ever give rise to a productive infection, but state-of-the-art diagnostics underscore the fact that infected naïve lymphocytes do harbor a mixed cohort of defective as well as intact unintegrated proviruses [47], therefore entertaining the feasibility of viremic rebound stemming from this cell population under favorable in vivo conditions [48]. Supporting this hypothesis, Dupont and collaborators recently demonstrated in primary T lymphocytes that Vpr can target the DDR protein SLF2 for degradation, which results in the SMC5/6 complex being unable to compact and repress unintegrated viral DNA intermediates. This would eventually increase the probability of unintegrated viral genomes gaining access to adequate integration sites in the host’s chromatin in a Vpr-dependent manner, eventually allowing for viral transcription and the various physiological effects resulting therefrom [49].
In T lymphocytes with successfully integrated proviruses, engendering a favorable setting for HIV-1 to exhibit optimal transcription is another phenomenon where Vpr might play an important role. Long terminal repeat (LTR) promoters are known for containing diverse cis-acting regulatory elements upstream of the proviral CATA box [50], including Sp-, NF-κB-, NFAT-, and AP1-binding sites, along with a downstream Tat-responsive element [51]. Importantly, while these transcriptional factors (TFs) are accepted as the most canonical regulators of the LTR, a recent study in CD4+ T cells from patients with low-level viremia identified a plethora of other TFs which directly or indirectly regulate LTR transcriptional activity [52]. About two decades ago, a study in Jurkat T cells postulated a correlation between Vpr’s abilities to transactivate the LTR and cause G2 arrest [53]. Since then, multiple other studies on this topic have been carried out on immortalized cell lines, including T cell-derived models, where it was observed that Vpr-dependent transactivation relies on its capacity to recruit various isoforms of the Sp transcription factor family to the LTR [54,55,56], although this effect proved highly cell type-dependent. Only in 2019 did the work of Hotter and collaborators finally demonstrate the importance of Sp1 availability towards viral expression in the context of CD4+ T cells, though the role of Vpr in this process was not assessed in their study [57]. Other work carried out in lymphocytic cell lines indicates that Vpr tethers the glucocorticoid receptor by mimicking the interaction with a steroid receptor coactivator (SRC), which induces it either to localize into the nucleus and into its binding site or to remain complexed in the cytosol in order for PARP-1 to be sequestered, thus preventing the latter from outcompeting Tat on its responsive element. Vpr was also seen to directly bind p300/CBP, itself an SRC proven to potentiate the effects of Tat [55,58]. However, again, these host cell interactions have never been functionally connected to G2 arrest in primary T cells.
As has been mentioned, it is generally hypothesized that the cell cycle plays a determinant role in retroviral physiology, as its arrest at the G2/M checkpoint seems to foster a cellular milieu optimal for HIV gene expression, which is suggested by the observation of increased LTR transcriptional activity in G2-arrested, HIV-infected Jurkat T cells [59]. Amongst other potential mechanisms, this phenomenon might be attributed to the likely presence of an internal ribosome entry site (IRES) in the 5′ leader sequence of viral RNA transcripts, since canonical cap-dependent translation is known to wane in favor of IRES-dependent initiation during the G2/M checkpoint [60]. Cell cycle checkpoint activation traditionally ensues as a consequence of DNA damage in the form of single- or double-strand breaks (SSBs and DSBs, respectively), base pair mismatches, stuck replicative machinery, or other types of lesions requiring structural amends. Nuclear sentinel proteins such as ATR or ATM are recruited to the damaged sites, where they phosphorylate various targets involved in cell cycle checkpoint control, in turn causing not only the upregulation of genes encoding DNA repair machinery but also the inactivation of cyclin-dependent kinases. This eventually results in cell cycle arrest, granting the cell time for repairs to take place or steering it towards apoptosis in case the damage is too extensive [61].
The specific involvement of Vpr in promoting G2 cell cycle arrest is one of its most studied properties, as it is one of the few Vpr phenotypes that can easily be recapitulated in transfected or infected cell lines, including immortalized CD4+ T cells [61]. Furthermore, induction of G2 arrest by Vpr can be attributed to its interaction with the CRL4-DDB1-DCAF1 E3 ubiquitin ligase complex, and this interaction is disrupted by specific mutations in Vpr (see the following section) [62]. On the other hand, induction of G2 arrest by Vpr has long been characterized as having a cytostatic effect on infected CD4+ T cell populations and leading to apoptosis [63], which is difficult to reconcile with Vpr possessing an important role during the early phase of the viral life cycle. The Bieniasz group recently demonstrated that Vpr can elicit the proteolysis of the chromosome periphery protein CCDC137 in primary lymphocytes, which could trigger an increase in γH2AX-rich nuclear foci as a consequence of DDR pathway engagement and thus lead to an interruption in cell cycle progression [64]. Another recent study in primary T cells shows that Vpr-triggered DNA damage responses also prevent the repressive TF ZBTB2 from coupling with the LTR-binding protein ZASC1, an interaction which would otherwise cause the recruitment of histone deacetylases (HDACs) and the subsequent silencing of the chromatin surrounding the proviral integration site [65]. Of note, G2 arrest in primary CD4+ T cells was not assessed in this study.
The U2OS [66], HT1080, and MIT-23 [67] cell models have been employed to demonstrate that Vpr expression is sufficient to induce DSBs and stall DNA replication, activating chromatin damage markers and leading to ATR recruitment. Independently thereof, Vpr was also able to inhibit both the homologous recombination (HR) and non-homologous end joining (NHEJ) pathways of DNA repair, contributing towards the genomic instability of the infected cell [66]. Evidence from a 2014 study suggests that the former pathway could be specifically jeopardized due to the Vpr-DCAF1-PLK1 axis hijacking the scaffolding agent SLX4com, causing the untimely activation of the EME1 and MUS81 nucleases and eventually contributing towards G2 arrest via the dismantlement of active replication forks whilst simultaneously disabling the resolution of DNA ultrafine anaphase bridges, further destabilizing the host genome [68]; again, this seemingly seminal finding still awaits confirmation in primary T cells. Meanwhile, the targeted proteasomal degradation of SIRT7, whose HDAC activity is of paramount importance in the repair of DSBs and the maintenance of DNA repair sites, was used to explain how Vpr compromises the NHEJ repair pathway in another recent study by Zhou et al. on HEK293T cells [69]. Finally, HIP1 was recently identified as yet another interaction partner of Vpr which showed a positive effect on G2 arrest irrespective of DSB induction, indicating that alternative mechanisms may be responsible for the stalling of the cell cycle and thus opening many potential research avenues in this particular field. However, this last finding was observed in the context of macrophages and immortalized cell lines, warranting confirmation in T lymphocytes as well [70].
Likewise, Vpr has been implicated in disrupting multiple other structures crucial for cell cycle progression, including the anaphase promoting complex/cyclosome, the centrosome, the replisome, as well as chromosomal centromeres and telomeres. Barbosa and colleagues recently showcased in CD4+ T cells how transmitter/founder Vpr variants coopt the DCAF1-DDB1 E3 Ligase complex to target APC1 for proteasomal degradation, potentially disabling the vital role it plays in the metaphase-to-anaphase transition [71]. Park and colleagues’ work on primary T cells evidenced that a similar mechanism is followed by Vpr in subverting centrosome activity, sequestering the kinase PLK4 and leading to centriole duplication [72]. Analogously, Wang and colleagues demonstrated that TERT, the catalytic subunit of telomerases, is also a target for Vpr-mediated proteolysis in HIV-1-infected CD4+ T lymphocytes [73]. Together, these effects destabilize the genomic integrity of the host and could perhaps be linked to a higher incidence of neoplasia in HIV patients [74]. Additionally, Vpr was shown to degrade the cell cycle-associated helicase activator MCM10, thereby stalling the DNA-unwinding machinery and cell cycle progression. Further, Vpr’s association with DCAF1 was implicated in the degradation of centromeric protein A (CENP-A), potentially contributing towards cell cycle delay and genomic instability. These findings, however, were exclusively observed in immortalized cell lines and must undergo further validation [75,76].
In summary, even though Vpr-mediated induction of G2 arrest in T cells is implied to result in a higher transcriptional state of the HIV-1 LTR, evidence for this causality happening in primary CD4+ T cells is lacking. Conversely, Vpr phenotypes associated with G2 in cell lines could also converge with regulating RT and integration in primary CD4+ T cells. Such a mechanism would explain the finding that a bevy of proteins playing a role in the regulation of the DDR are degraded by Vpr via the CRL4-DDB1-DCAF1 E3 ubiquitin ligase complex. Some recent developments in this field have seen the confirmation of various Vpr functions in the context of CD4+ T cells, including its indirect role in antagonizing ZBTB2-ZASC1 LTR silencing, as well as the targeted degradation of SLF2, CCDC137, TERT, and APC1, together with the sequestration of PLK4. Other compelling pieces of evidence, such as the directed degradation of SIRT7, HIP1, MCM10, and CENP-A, in addition to the Vpr-mediated sequestration of SLX4com to engage DNA damage response and trigger G2 arrest, still warrant validation in CD4+ T cells to attain full physiological relevance.
In total, surprisingly few studies have been able to establish a link between Vpr and G2 arrest occurring in primary CD4+ T cells. For one, the Planelles group showed that HIV-1-infected activated primary T lymphocytes exhibit a G2/M phenotype that is partially dependent on ATM/ATR and Vpr [61,63,77,78,79]. Second, a study in humanized mice described Vpr-induced G2/M arrest in regulatory T cells undergoing depletion before strongly sustaining HIV-1 replication [80]. Hence, while these studies implicate Vpr’s G2 arresting capacity in vivo, it is essential and of very high relevance to carefully delineate the detailed context and relevance of Vpr-induced G2/M arrest during primary CD4+ T cell infection. To facilitate this and recapitulate the complex and different order of events governing Vpr-mediated G2 arrest, an overview of Vpr’s involvement in this phenomenon and DDR engagement is depicted in Figure 1.

4. Systems-Level Manipulation of the Host Cell

In recent years, several studies have sought to assess the overall systems-level manipulation Vpr exerts on the host cell, hinted at by its ability to degrade a bevy of host cell factors, resulting in major changes of the host cell proteome and transcriptome. A study by Malim’s group demonstrated how HIV-1 utilizes virion-delivered Vpr to induce wide-ranging transcriptional alterations in primary CD4+ T cells as early as 4.5 h post-infection, clearly anticipating interferon-stimulated gene (ISG)-dependent expression changes by about 8 h [81]. Another study in Jurkat T cells indicated that the presence of Vpr suffices to upregulate the expression of miRNA 210-5p, itself eliciting the downmodulation of TGF-β-response repressor TGIF2. This protein is involved in proliferation and cell differentiation, meaning its depletion may indirectly also contribute towards G2 arrest [82]. In line with this finding, not only did Sato and collaborators observe enhanced cell cycle arrest in Treg cells in HIV-1 infected humanized mice, but they also demonstrated that this subset’s higher proliferative potential renders it more permissive towards productive infection of CCR5-tropic viruses, followed by large-scale apoptosis and the subsequent depletion of this lymphocyte subpopulation [80]. This phenomenon could partly explain the higher incidence of autoimmune disorders reported in HIV patients, as this regulatory T cell subset secretes TGF-β to help modulate immune responses and temper otherwise uncontrolled inflammation to commensurate proportions [83].
Tightly coupled to transcription, splicing is another phenomenon over which HIV has been demonstrated to exert a very significant influence [84,85]. The first in vitro evidence for the involvement of Vpr in this process appeared in 2005, where cellular pre-mRNA splicing was shown to be severely inhibited [86]. Shortly afterwards, a second in vitro study revealed that Vpr is able to hijack several components of the spliceosome, such as SAP145, U1-70K, U2AF65, and U2B’’, as well as U1 and U2; this led not only to the inhibition of splicing of cellular mRNAs but also enabled preferential processing of viral primary transcripts [87]. Vpr’s capability to regulate splicing of viral transcripts was later mapped to its N-terminus, as tagging this site led to oversplicing and ablated viral production in SupT1 cells [88].
In comparison to other viral pathogens, HIV-1 is notorious for its flexibility regarding ribosome recruitment strategies. As touched upon in the previous section, IRES-dependent initiation could be favored during G2 cell cycle arrest to aid viral expression [60]. However, a recent study indicated that the IRES present in the Gag-coding region showcases inefficient usage in the context of infected CD4+ T cells, with ribosome profiling identifying concomitant translation of upstream open reading frames and viral genes, significantly reducing the translation efficiency of the latter [89]. This appears to be greatly alleviated by the presence of the translation-associated RNA helicase DDX3 [90], whose pharmacological inhibition led to apoptosis in latency-reversed CD4+ T cells from combined antiretroviral therapy (cART)-treated HIV+ donors [89,91]. Further, a 2012 in vitro study correlated Vpr’s capability to induce G2 arrest with the phospho-4E-binding protein (P-4E-BP)-mediated sequestration of eIF4E, explaining the overall decline in the translation of cellular mRNAs observed upon infection; whilst fully spliced viral mRNAs continued to be detected in the vicinity of ribosomes at the 5′-cap, partially or non-spliced transcripts were instead predominantly associated with the stress-related nuclear cap-binding complex (CBC) [92,93]. This finding was recently confirmed in HIV-infected CD4+ T cells, where it became clear that the non-canonical CBC-to-CBP80 cap exchange in unspliced viral transcripts was enabled by the hypermethylation of the 5′-cap. Ultimately, this phenomenon allows HIV to escape the Vpr-elicited eIF4E-dependent translational bottleneck, permitting sustained viral protein production via an emergency cap-exchange mechanism despite global translational shutdown. However, the authors did not establish a direct link to Vpr in this study [94].
Immune function is one of the main aspects of lymphocyte physiology deeply affected by Vpr upon infection. An innovative study by Chiodi’s group performed mass cytometry on CD4+ T cells from HIV+ donors, showing a negative correlation between the timing of cART administration and the preservation of immune function; this was reflected by the abundance of surface markers associated with inflammation, cell migration, and survival [95]. Shan et al. identified that CD4+ lymphocytes undergoing an effector-to-memory transition inadvertently facilitate the establishment of HIV-1 latency, particularly of CXCR4-tropic strains. Their results indicated that this population’s initial transcriptional pattern renders them permissive to infection, but prolonged viral gene expression cannot be sustained, since the phenotypic switch to memory T cells favors quiescence [96]. This phenomenon was further explored by Jolly’s group, who further characterized the HIV-mediated immunological reprogramming of infected T cells. Their work identified that viral spread within lymphocyte populations via cell-to-cell contacts preferentially leads to infection of resting memory T cells. In this scenario, virion-delivered Vpr induces the establishment of a tissue-resident phenotype by synergizing with IL-7 to activate STAT5, ultimately inducing a virus-favorable transcriptional program [97]. Ectopic STAT5 activation has recently been observed to alter the epigenome of CD4+ T cells and push their phenotype towards polyfunctionality, although such an alteration may present yet unidentified implications in the context of infection, perhaps playing a significant role in HIV pathogenesis [98]. However, this effect does appear to synergize with the presence of other leukocytes like non-cytolytic CD8+ T cells [99] or activated macrophages [100], which have been linked with the establishment of HIV latency by pushing CD4+ T cell metabolism, signaling, and transcription towards quiescence.
Overall, and as was introduced earlier, Vpr’s capability to hijack the CRL4-DDB1-DCAF1 E3 ubiquitin ligase complex allows HIV to directly modify the host cell’s proteome upon infection, underpinning the establishment of an optimal intracellular milieu perhaps even before proviral integration and its subsequent transcriptional and translational effects. Lehner’s group recapitulated on Vpr’s capability to direct a vast array of targets towards proteasomal degradation using CEM-T4 T cells, showing how the overwhelming majority of observed alterations is mostly attributable to Vpr [14]. In parallel, they backed this up by analyzing synchronously infected primary CD4+ T cells, resulting in a detailed overview of HIV-elicited proteomic changes at 24 and 48 h post-infection. Indeed, results in primary CD4+ T cells not only mirrored the broad depletion pattern of Vpr in CEM-T4 T cells but further confirmed the targeted degradation of several proteins previously only studied in immortalized cell lines, such as HLTF, UNG2, as well as the DDR endonucleases MUS81 and EME1 [101]. Together, these groundbreaking studies present virion-delivered Vpr as the main driving force behind the proteomic remodeling elicited by HIV in the context of infection.
In summary, the evidence gathered in a physiologically relevant context indicates that Vpr induces an early uptick in ISG transcription upon infection [81] and indirectly enables cap hypermethylation, allowing viral mRNAs to overcome the host’s transcriptional blockage [92,93,94]. It has also been shown that Vpr plays a leading role in manipulating IL-7 and STAT5 immune signaling, prompting the establishment of latency with the help of the host lymphocyte’s cellular environment [97]. In addition, the CRL4-DDB1-DCAF1-mediated remodeling of the host proteome not only factually explains many of the alterations reported in this cell population but also confirms that Vpr actively targets the HR-related proteins MUS81 and EME1 for degradation, endorsing a branch of the hypothesis that attributes Vpr to the DNA damage response pathway [101]. On the other hand, and despite much evidence existing to support the notion that Vpr subverts the spliceosome to aid viral processing, its supporting facts remain to be validated in CD4+ T-cells. This also holds true for the Vpr-induced upregulation of miR-210-5p; despite intriguingly tying TGF-β and NF-κB signaling into the Vpr-induced G2 arrest phenomenon, it has thus far only been observed in immortalized cell lines and thus requires further validation in T lymphocytes.

5. Signaling Disruption and Immune Evasion

Vpr is thought to modulate various signaling pathways, many of them intrinsically involved in immune activation, inflammation, proliferation, and survival [102,103,104]. Amongst them, the NF-kB signaling cascade has long been suspected of playing a central role. However, comprehending how the activation and nuclear translocation of NF-κB is influenced by Vpr remains an undeciphered riddle, largely due to the puzzling nature of the available evidence in this field. A 1997 study on PBMCs implicated Vpr in suppressing NF-κB-mediated gene activation [105], an effect later observed in CD4+ T cells, albeit without establishing a direct connection to Vpr [100]. This link was first proven to ensue via DCAF1 in 2013 [106] and then thoroughly explained by Khan et al. in 2020, who showed how virion-delivered Vpr associates with DCAF1 in order to interact with subunit 1 of karyopherin-α (KPNA1), thus sabotaging the nuclear import of NF-κB in CD4+ T cells [107]. Another study examined the NF-κB-inhibiting capability of multiple Vpr alleles in PBMCs, the results indicating that only de novo-synthetized Vpr from simian immunodeficiency viruses SIVcol and SIVolc retains this trait; in the case of HIV-1, this function is seemingly delegated to Vpu instead [108].
Antithetically, a relatively large body of evidence supports the notion that both the canonical as well as the non-canonical pathways of NF-κB activation are directly triggered by Vpr to varying extents. Such is the case with the Vpr-induced phosphorylation of the kinase TAK1, proven to result in the nuclear translocation of NF-κB and AP-1 in PBMCs via the canonical pathway [109]. In Jurkat T cells, a similar mechanism is followed by Vpr in order to phosphorylate IKKα/β, activating both pathways in parallel for an enhanced NF-κB response [110]. Specifically, activating NF-κB via the non-canonical pathway has recently been described to aid HIV latency reversal [111], demonstrated in CD4+ T cells from healthy donors [112]. On the other hand, the “atypical” pathway of NF-κB activation is known to be triggered by genotoxic stress and may play a role in HIV pathogenesis [113]. Liang and collaborators hinted at a link existing between Vpr-mediated cell cycle arrest and NF-κB translocation [114], with a novel study performed in macrophages indicating that Vpr-triggered DNA damage activates ATM, followed by the nuclear transport of various TFs, including NF-κB, CEBPB, and JUN [115]. Vpr’s capacity to interact with the SRC p300/CBP, until now only described in immortalized cell lines, did not only aid Tat’s transactivating role but was also accompanied by a significant increase in binding affinity between NF-κB and its LTR regulatory sequence [55,58].
NF-κB activity proves relevant even when HIV-1 is pitted against antiviral defenses, as might be the case with PYHIN-mediated innate immunity. Kirchhoff’s group demonstrated in CD4+ T cells that various members of this restriction factor family [116] antagonize HIV transcriptional activity by sequestering Sp1 [57,117], with their newest results indicating that an additional NF-κB binding site in the LTR of subtype C members enables sustained transcription and enhanced reactivation potential, a notion that could explain the overrepresentation of this HIV-1 clade at the global scale [118]. Given that NF-κB regulates inflammation and the induction of ISGs but is simultaneously essential for proliferation and survival as well activating the HIV-1 LTR, its manipulation represents a double-edged sword in the context of HIV infection. As such, it warrants a highly precise timing for the virus to exploit its stimulatory boost whilst escaping subsequent deleterious effects. All in all, the reigning theory in this field describes NF-κB activation as being required during the early stages of the viral life cycle to transcriptionally engage the proviral LTR, but its translocation is inhibited at later stages via diverse mechanisms, thereby lowering the expression of genes associated with antiviral innate immunity [119]. Nevertheless, it remains an enigmatic phenomenon whether Vpr activates or suppresses NF-κB differentially across the viral life cycle, as there is a large body of conflicting data related to this topic, even in primary CD4+ T cells.
Apart from NF-κB, Vpr influences other transcription factors involved in the regulation of lymphocytic physiology and regulation of the HIV-1 LTR. The nuclear factor of activated T cells (NFAT), of paramount importance for T cell activation, proliferation, differentiation, and survival, is canonically activated by membrane receptors coupled to Ca2+ mobilization, resulting in its nuclear translocation. Using a variety of cellular models as well as primary lymphocytes, our group demonstrated that virion-delivered Vpr is able to bypass upstream stimulation to activate NFAT and thus support productive HIV infection, a phenotype that is additionally correlated with Vpr-mediated LTR transactivation and G2 arrest in T cell lines [120]. Closely associated with innate immune activation and inflammation, interferon regulatory factor 3 (IRF3) is another signaling intermediate acting downstream of TLR activation. This protein has been shown to undergo proteasomal degradation via the presence of both Vpr and Vif in Jurkat T cells and PBMCs [121]. Further, the possibility of IRF3 localizing to the nucleus is negated by Vpr’s ability to sequester and degrade KPNA1, a phenomenon demonstrated in CD4+ T cells [107].
Besides rewiring intracellular communications to seize control over the host, HIV-1 also employs a multipronged approach to circumvent most defensive mechanisms placed against it. Though all of its accessory proteins participate in this task to different extents [122], Vpr might possess a thus far unprecedented role therein, due to its immediate deployment directly upon host cell entry as well as its unique ability to induce the degradation of target proteins. For example, the interaction between Vpr and SLX4com through VprBP, implicated in eliciting G2 arrest, was also shown to induce the processing of excess HIV-1 DNA, preventing its sensing and the consequent type 1 IFN response. This, however, has only been shown in HeLa cells thus far [68]. APOBEC3G, one of the most potent antiretroviral restriction factors identified to date [123], was originally thought to be exclusively counteracted by Vif, although novel evidence from a study in the MT4 and H9 T cell lines suggests that Vpr can initially perform this task before de novo production of Vif ensues [124]. LAPTM5, a restriction factor that reroutes HIV-1 gp120 for lysosomal degradation in macrophages, has been recently observed to be a target of Vpr via DCAF1; this was additionally observed in LAPTM5-retransfected CD4+ T-cells, confirming its interaction with Vpr. Whilst intriguing, the physiological relevance of this finding to T cells remains enigmatic, as this restriction factor is normally not present in the lymphoid proteome [125].
Silencing, both at the transcriptional and epigenetic levels, is another antiviral mechanism that HIV-1 can override with the help of Vpr. SRSF1, an IFN-modulated splicing factor, was observed to hinder overall LTR activity in Jurkat and THP-1 cells [126]. In this context, a recent pioneering study involving direct RNA nanopore sequencing in CD4+ T cells identified the Vpr-mediated upregulation of ciTRAN, an HIV-modulating circular RNA that sequesters SRSF1 to prevent it from burdening viral gene expression [127]. Further, our group employed primary T lymphocytes to demonstrate that, after aiding proviral integration, PHF13 is robustly degraded by Vpr in order for HIV to escape its restrictive effects on transcription [128]. Of note, PHF13 might regulate major epigenetic DNA modifications [129,130], and its degradation by HIV-1 Vpr is thus highly reminiscent of HIV-2/SIV Vpx-mediated degradation of the HUSH complex [12]. Other mechanisms to counter epigenetic silencing by Vpr happen through the targeted proteasomal degradation of various nucleosome remodeling and deacetylase (NuRD) complex components. This was first observed in the HEK293T and HeLa cell lines, where Vpr coopted DCAF1 to degrade ZIP and sZIP, two DNA-binding adaptors known to interact with NuRD [131]. Later, a Vpr-dependent degradation of the scaffolding protein CTIP2 was reported in CD4+ T cells, thereby negating the NuRD-mediated recruitment of HDACs, lysine demethylases, and histone methyltransferases to the LTR [132]. Finally, the recently discovered putative RNA ploymerase-binding factor RPRD2 [133] is able to restrict HIV-1 by binding viral nucleic acids and recruiting the epigenetic repressors HUSH [134] and PAF1 [135] to deposit silencing marks therein. Though only so far observed in macrophages and HeLa CD4+ cells, McKnight’s group demonstrated that Vpr targets RPRD2 for degradation via CRL4-DDB1-DCAF1, protecting the provirus from silencing and boosting overall transcription rates in the affected cell [136,137].
Both the timing Vpr exhibits in re-orchestrating intracellular signaling and its swift defusal of pre-deployed antiviral defenses clearly indicate that the strategy followed by HIV is inherently aimed towards preventing the establishment of a widespread antiviral response while triggering several pathways that generate a cellular milieu favorable for viral replication. So far, the available evidence pertaining to CD4+ T cells paints Vpr as a central mediator in the induction of early NFAT signaling that is later on sustained by Nef [138,139]. Induction of ISGs and the interferon response is ablated by interfering with IRF3 nuclear translocation. NF-κB signaling is manipulated both negatively by sabotaging its nuclear translocation together with DCAF1 and KPNA1 as well as positively by triggering its non-canonical and atypical activatory pathways. Whether Vpr is also able to trigger NF-κB’s canonical activation pathway and the DNA damage response is thereby activated warrants further study in T lymphocytes, along with the characterization of the timepoint at which enhanced NF-κB signaling stops being favored by HIV. Likewise, Vpr has been confirmed to degrade LAPTM5, PHF13, and CTIP2, while upregulating ciTRAN, helping HIV evade transcriptional and epigenetic repression in primary T lymphocytes. RPRD2 and APOBEC3G degradation, albeit highly interesting, have yet to be studied in primary CD4+ T cells.

6. Inflammation and Apoptosis

Heavily intertwined with both signaling and immunity, inflammation and apoptosis are essential aspects of cellular physiology that could be affected by Vpr [140]. In the year 2000, Roux and collaborators used primary CD4+ T cells to demonstrate that Vpr elicits the expression of IL-8 and TNF-α upon infection [141], with a 2015 study further confirming not only the release of the latter but also mechanistically implicating the previously mentioned DDB1 and TAK1 proteins therein [142]. Notably, the aforementioned proinflammatory transcriptional program induced by Vpr at the systems level [81] seems to be at least partly dependent on Vpr enhancing cGAS-mediated viral DNA sensing, boosting innate immunity and inflammation [143]. While this may initially appear counterintuitive, the same study demonstrated that Vpu negatively regulates this effect at later stages of infection. Therefore, a certain level of proinflammatory activity by Vpr could prove beneficial for viral replication, or alternatively arise as a “side effect” of its biological activity. Further, a recent genome-wide association study performed on PBMCs from a broad cohort of HIV+ patients unraveled that infected cell populations exhibit an epigenetic dysregulation in distinct gene clusters; ISG promoters of patients with higher viral loads presented a hypomethylation pattern, which translated into an enhanced proinflammatory response. In parallel, uncontrolled infection was also found to correlate with the hypermethylation of cell markers associated with the commitment to a T follicular helper cell fate, a lymphocyte subset essential for the sustainment of long-term immunity and the development of specialized humoral responses [144]. Though not assessed by this study, it is plausible that Vpr could play a significant role in the early establishment of an epigenetic landscape favoring viral proliferation, as evidenced by its capacity to selectively manipulate mediators of chromatin remodeling, such as PHF13 and CTIP2 [128,132].
Programmed cell death is frequently the endgame of inflammatory responses, as well as one of the leading mechanisms behind lymphocyte depletion in HIV+ patients, a hallmark of retroviral pathophysiology. Vpr’s inherence therein has long been observed, but the precise mechanism whereby it leads to cell death remains a highly controversial topic. Virion-associated Vpr has been demonstrated to induce cytotoxicity by itself whilst also potentiating the effect of Fas-mediated apoptosis in activated PBMCs [145]. Conti and collaborators observed that, at early timepoints post-infection, Vpr confers on infected PBMCs protection from apoptosis despite the presence of TNF-α signaling, although a completely opposite outcome presented itself at later timepoints [146]. In 2002, a study in PBMCs and primary T lymphocytes demonstrated that Vpr permeabilizes the mitochondrial membrane and induces the release of proapoptotic proteins towards the cytosol, triggering Caspase 9-mediated cell death [147]. This effect was later attributed to the Vpr-mediated activation of the apoptosis regulator Bax and to the DCAF1-mediated degradation of mitofusin 2 (Mfn2), both in T cell-derived models as well as in primary T lymphocytes [77,148]. In light of the fact that Bax is a known effector of inflammation and apoptosis [149] acting downstream of the DNA damage response [150], a mechanistic correlation between the Vpr-triggered induction of cell death and G2 checkpoint arrest became increasingly debated, with Sakai and colleagues proposing such a link existing not only for Vpr but independently also for Vif [151].
However, the initiation of programmed cell death in CD4+ T cells is also dependent on multiple other factors, including the microenvironment a cell may encounter itself in. Trinité et al. demonstrated in primary T lymphocytes how the proapoptotic effect of virion-delivered Vpr can be counteracted by the activity of common γ-chain cytokines like IL-7 and IL-4, frequently found in the sites of active HIV-1 replication [152]. Detected in the serum of infected patients [153,154], virion-free Vpr has also been suggested to permeate into T cells and thus elicit a diverse array of effects depending on the identity of the target. In lymph nodes, such an effect may trigger the death of uninfected bystander T cells [155] through both the intrinsic and extrinsic pathways; indeed, Cohen’s group demonstrated that soluble recombinant Vpr entering bystander T-cells can directly rouse the G2 arrest-inducing DNA damage response whilst simultaneously upregulating the NKG2D ligands ULBP-1 and ULBP-2 [156,157]. Analogously, Vassena et al. observed how this Vpr-triggered DDR also leads to the upregulation of poliovirus receptor (PVR) on the surface of T lymphocytes, increasing the likelihood to trigger NK-mediated cytotoxicity [158]. Other studies showed that both IL-6 and soluble Vpr stimulate viral reactivation in latently infected cell lines through the ROS-induced TLR4-MyD88 signaling axis, leading to further IL-6 release and the subsequent expansion of the reactivating stimulus [154]. Tying into this hypothesis, a recent study in macrophages unveiled that Vpr can selectively deplete the methylcytosine dioxygenase TET2 via CRL4-DDB1-DCAF1, negatively impacting 5mC levels in the IL-6 promoter and likely leading to sustained IL-6 production and release, greatly fueling the reactivation of latent HIV-1 in adjacent T cells [159].
In a nutshell, HIV appears to ensure its spread across critical leukocytic populations by promoting constant inflammation, thereby sensitizing cells for productive infection, whilst remaining unnoticed by immune surveillance; in cells harboring latent provirus, such an environment can trigger viral reemergence and production. It has so far been shown in CD4+ T cells that Vpr degrades TAK1 to boost TNF-α and IL-8 signaling, while simultaneously activating mitochondrial Bax and degrading Mfn2 to initiate apoptosis. Recombinant Vpr has also been shown to permeate bystander lymphocytes, upregulating markers that increase their likelihood of undergoing cytotoxic cell death. Future research has yet to unveil if the targeted degradation of TET2 observed in macrophages holds true in CD4+ T cells, as well as if this protein provides infected cells protection from apoptosis under specific conditions, if at all. Additionally, it remains an open research avenue to identify whether Vpr represents a nexus between G2 arrest induction, DNA damage responses, and cell death.

7. Variation in Vpr’s Primary Structure

Multiple studies have sought to map the functional properties of Vpr down to the amino acid level by employing various naturally occurring and culture-derived variants, together with site-directed mutagenesis. These efforts have resulted in the identification of several amino acid positions that play a vital role in defining the functional properties of Vpr’s structural elements, eventually granting it the ability to interact with diverse partner proteins as well as to effectively localize to certain subcellular compartments in order to exert many of its hallmark effects. Of central importance to Vpr’s pathogenicity and functionality, glutamine 65 is located in a short leucine-rich motif within the α3 helix, a domain responsible for several Vpr attributes, including nuclear localization, cell cycle arrest, and multimerization. The work of Wu et al. demonstrated that Q65 is critical for stabilizing the Vpr–DCAF1 interaction, as this residue partakes in the formation of a hydrophobic pocket where Trp1156 of DCAF1 docks whilst concomitantly engaging in hydrogen bonding with Ser1136 of the latter’s WD40 β-propeller domain [7]. The Q65R substitution, detected in long-term non-progressor HIV patients, renders Vpr unable to interact with DCAF1 [160], as experimentally confirmed by Skowronski’s group in ex vivo-infected CD4+ T cells [26]. Given how crucial the takeover of the said E3 ligase adaptor is to Vpr’s broad-acting effects in the host cell, it comes as no surprise that this mutant is associated with a milder course of infection, in addition to the fact that it fails to induce cell cycle arrest, cannot localize to the nuclear envelope, and does not present any alteration in cytopathic potential [160].
Cytopathicity, in turn, is another extensively studied Vpr function. Soares and collaborators first identified a R77Q mutation, together with additional Q3R substitution, in a long-term non-progressor pediatric patient who showed no symptoms other than recurrent ear infections [161]. Ever since, multiple studies have collaboratively mapped cytopathic potential to a Vpr moiety known as the mitochondrial membrane permeabilization (MMP)-inducing sequence, which contains two H(F/S)RIG motifs along the amino acid stretch 72–80. These motifs, observed to bring about gross cell enlargement and replication arrest [5,162,163], were more closely studied in ex vivo-cultured human lymphoid tissue. This work unraveled the physiological importance of residues R77 and R80, since point mutations R77A, R77Q, and R80A all led to reduced viral replication, whilst the latter two mutants also showed an impairment in the induction of T cell depletion [164]. Markedly, the R80A mutation possesses a dominant-negative phenotype in the context of co-expression with wt Vpr, disabling its capability to trigger G2 arrest. This phenotype is reverted when Vpr presents both R80A and Q65R mutations, further underscoring the link between Vpr’s G2 arrest-inducing and DCAF1-recruiting abilities [61,62,160] A recent structural study by Byeon et al. demonstrated that helix α3 of Vpr represents an important contact interface both for DCAF1 as well as for the DNA repair protein hHR23A; this finding links proteasomal degradation and ubiquitin recognition with the nucleotide excision repair pathway and G2 arrest, potentially explaining the phenotypes engendered by mutations on residues R77 and R80 [165].
Other physiologically relevant sites are Vpr’s hydrophobic amino acids, not only responsible for stabilizing its inner core [5] but also having been observed to affect the spatial configuration of its α-helices, influencing Vpr’s capability to localize to the nucleus as well as to oligomerize at the nuclear envelope; such is the case with positions I63, L67, I70, and I74 [166]. In addition, hydrophobic positions in helix α1, such as W18 and L22, have been described as crucial to oligomerization, whilst potentially also contributing towards G2 arrest and cytopathicity [167]. Further, S79 has been identified to be a target for phosphorylation [168], this step seemingly being necessary for the cell cycle arresting activity of Vpr [169]. Thus far, the only mutations associated with rapid disease progression are the relatively uncommon R36W, L68M, and R85Y, the first also being characterized as eliciting increased viral replication by granting Vpr a higher oligomerization potential [170]; this evidence notwithstanding, the mechanisms whereby such an effect takes place remain unelucidated. All these mutations are part of a long list of single sequence polymorphisms experimentally associated with a variety of phenotypes, which, despite greatly contributing to a better understanding of Vpr’s functional properties, have so far only been studied in immortalized cell lines and/or are not necessarily naturally occurring. As such, their study in physiologically relevant models as well as their further characterization at the molecular level represent an open research avenue. A complete summary of these findings can be found in Table 1.

8. Concluding Remarks

The present work has evidenced that great advances have been achieved in elucidating the role that Vpr plays during HIV-1 infection. It has become clear that this accessory protein is a virulence determinant for HIV-1 to achieve full in vivo pathogenicity, transmission, spread, and cellular persistence. Considering that it is abundantly present within viral particles, its direct intracellular delivery is one of the first events to take place during the HIV life cycle, representing an optimal timepoint for its multifaceted effects to execute the furtive domination of infected lymphocytes.
Given the compelling evidence that has been gathered regarding Vpr’s functionality in the context of primary T cells and presented within this review (summarized in Figure 2), we postulate that Vpr’s proinflammatory and activating nature sensitizes newly infected as well as bystander T cells for productive infection. Concomitantly, it exploits a repertoire of mechanisms to manipulate host cell factors involved in organizing the cell cycle and DNA repair, further aiding reverse transcription and viral integration. As these processes are intertwined with the activation of antiviral defenses, including innate immunity and the interferon response, Vpr has likely also evolved mechanisms to counteract the detrimental effects induced by its own proviral activities. Of course, this model warrants experimental validation in addition to the evidence detailed herein. Additionally, various existing pieces of the complex Vpr puzzle await their final validation in CD4+ T lymphocytes (see Figure 1 and Figure 2). As such, the detailed spatiotemporal characterization of the mechanisms encompassing Vpr-mediated G2 arrest, as well as their functional interconnection with essential steps of viral replication, will undoubtedly serve as the necessary precedent to attain a more exhaustive understanding of Vpr’s role in mediating viral pathogenicity, taking special consideration of the particularities various CD4+ T cell subsets exhibit during the course of HIV-1 infection.

Author Contributions

C.A.V.-T. wrote the initial manuscript draft. C.A.V.-T. and M.S. jointly developed the manuscript to its final form. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by the German Research Foundation (DFG), grant number SCHI1073/7-2 (project number 318213433), in the context of the priority program SPP1923, awarded to MS. Support for the APC was given by the Open Access Publication Fund of the University of Tübingen.

Acknowledgments

We would like to thank Kristina Hopfensperger for critical reading of the manuscript and Maximilian Bunz, Rishikesh Lotke, and Daniel Sauter for helpful comments and discussion.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Anderson, J.L.; Hope, T.J. HIV Accessory Proteins and Surviving the Host Cell. Curr. HIV/AIDS Rep. 2004, 1, 47–53. [Google Scholar] [CrossRef] [PubMed]
  2. Woodburn, B.; Emery, A.; Swanstrom, R. Human Immunodeficiency Virus (Retroviridae). In Encyclopedia of Virology, 4th ed.; Elsevier: Amsterdam, The Netherlands, 2020; Volume 1–5, pp. 460–474. ISBN 9780128145166. [Google Scholar]
  3. Fabryova, H.; Strebel, K. Vpr and Its Cellular Interaction Partners: R We There Yet? Cells 2019, 8, 1310. [Google Scholar] [CrossRef] [PubMed]
  4. Laguette, N.; Benkirane, M. Shaping of the Host Cell by Viral Accessory Proteins. Front. Microbiol. 2015, 6, 142. [Google Scholar] [CrossRef] [PubMed]
  5. Gonzalez, M.E. The HIV-1 Vpr Protein: A Multifaceted Target for Therapeutic Intervention. Int. J. Mol. Sci. 2017, 18, 126. [Google Scholar] [CrossRef] [PubMed]
  6. Bredbenner, K.; Simon, S.M. Vpr Co-Assembles with Gag during HIV-1 Assembly. bioRxiv 2020. [Google Scholar] [CrossRef]
  7. Wu, Y.; Zhou, X.; Barnes, C.O.; DeLucia, M.; Cohen, A.E.; Gronenborn, A.M.; Ahn, J.; Calero, G. The DDB1-DCAF1-Vpr-UNG2 Crystal Structure Reveals How HIV-1 Vpr Steers Human UNG2 toward Destruction. Nat. Struct. Mol. Biol. 2016, 23, 933–940. [Google Scholar] [CrossRef] [PubMed]
  8. Morellet, N.; Bouaziz, S.; Petitjean, P.; Roques, B.P. NMR Structure of the HIV-1 Regulatory Protein VPR. J. Mol. Biol. 2003, 327, 215–227. [Google Scholar] [CrossRef] [PubMed]
  9. Li, Y.L.; Langley, C.A.; Azumaya, C.M.; Echeverria, I.; Chesarino, N.M.; Emerman, M.; Cheng, Y.; Gross, J.D. The Structural Basis for HIV-1 Vif Antagonism of Human APOBEC3G. Nature 2023, 615, 728–733. [Google Scholar] [CrossRef]
  10. Kmiec, D.; Iyer, S.S.; Stürzel, C.M.; Sauter, D.; Hahn, B.H.; Kirchhoff, F. Vpu-Mediated Counteraction of Tetherin Is a Major Determinant of HIV-1 Interferon Resistance. mBio 2016, 7, e00934-16. [Google Scholar] [CrossRef]
  11. Hofmann, H.; Logue, E.C.; Bloch, N.; Daddacha, W.; Polsky, S.B.; Schultz, M.L.; Kim, B.; Landau, N.R. The Vpx Lentiviral Accessory Protein Targets SAMHD1 for Degradation in the Nucleus. J. Virol. 2012, 86, 12552–12560. [Google Scholar] [CrossRef]
  12. Chougui, G.; Munir-Matloob, S.; Matkovic, R.; Martin, M.M.; Morel, M.; Lahouassa, H.; Leduc, M.; Ramirez, B.C.; Etienne, L.; Margottin-Goguet, F. HIV-2/SIV Viral Protein X Counteracts HUSH Repressor Complex. Nat. Microbiol. 2018, 3, 891–897. [Google Scholar] [CrossRef]
  13. Staudt, R.P.; Alvarado, J.J.; Emert-Sedlak, L.A.; Shi, H.; Shu, S.T.; Wales, T.E.; Engen, J.R.; Smithgall, T.E. Structure, Function, and Inhibitor Targeting of HIV-1 Nef-Effector Kinase Complexes. J. Biol. Chem. 2020, 295, 15158–15171. [Google Scholar] [CrossRef]
  14. Greenwood, E.J.D.; Williamson, J.C.; Sienkiewicz, A.; Naamati, A.; Matheson, N.J.; Lehner, P.J. Promiscuous Targeting of Cellular Proteins by Vpr Drives Systems-Level Proteomic Remodeling in HIV-1 Infection. Cell Rep. 2019, 27, 1579–1596.e7. [Google Scholar] [CrossRef]
  15. Schabla, N.M.; Mondal, K.; Swanson, P.C. DCAF1 (VprBP): Emerging Physiological Roles for a Unique Dual-Service E3 Ubiquitin Ligase Substrate Receptor. J. Mol. Cell Biol. 2019, 11, 725–735. [Google Scholar] [CrossRef]
  16. Nakagawa, T.; Mondal, K.; Swanson, P.C. VprBP (DCAF1): A Promiscuous Substrate Recognition Subunit That Incorporates into Both RING-Family CRL4 and HECT-Family EDD/UBR5 E3 Ubiquitin Ligases. BMC Mol. Biol. 2013, 14, 22. [Google Scholar] [CrossRef] [PubMed]
  17. Romani, B.; Cohen, É.A. Lentivirus Vpr and Vpx Accessory Proteins Usurp the Cullin4-DDB1 (DCAF1) E3 Ubiquitin Ligase. Curr. Opin. Virol. 2012, 2, 755–763. [Google Scholar] [CrossRef]
  18. Maertens, G.N.; Engelman, A.N.; Cherepanov, P. Structure and Function of Retroviral Integrase. Nat. Rev. Microbiol. 2022, 20, 20–34. [Google Scholar] [CrossRef] [PubMed]
  19. Stark, L.A.; Hay, R.T. Human Immunodeficiency Virus Type 1 (HIV-1) Viral Protein R (Vpr) Interacts with Lys-TRNA Synthetase: Implications for Priming of HIV-1 Reverse Transcription. J. Virol. 1998, 72, 3037–3044. [Google Scholar] [CrossRef] [PubMed]
  20. Kleiman, L. TRNA Lys3: The Primer TRNA for Reverse Transcription in HIV-1. IUBMB Life 2002, 53, 107–114. [Google Scholar] [CrossRef]
  21. Guenzel, C.A.; Hérate, C.; Le Rouzic, E.; Maidou-Peindara, P.; Sadler, H.A.; Rouyez, M.-C.; Mansky, L.M.; Benichou, S. Recruitment of the Nuclear Form of Uracil DNA Glycosylase into Virus Particles Participates in the Full Infectivity of HIV-1. J. Virol. 2012, 86, 2533–2544. [Google Scholar] [CrossRef]
  22. Eldin, P.; Chazal, N.; Fenard, D.; Bernard, E.; Guichou, J.F.; Briant, L. Vpr Expression Abolishes the Capacity of HIV-1 Infected Cells to Repair Uracilated DNA. Nucleic Acids Res. 2014, 42, 1698–1710. [Google Scholar] [CrossRef] [PubMed]
  23. Wen, X.; Klockow, L.; Nekorchuk, M.; Sharifi, H.J.; de Noronha, C.M.C. The HIV1 Protein Vpr Acts to Enhance Constitutive DCAF1-Dependent UNG2 Turnover. PLoS ONE 2012, 7, e30939. [Google Scholar] [CrossRef] [PubMed]
  24. Langevin, C.; Maidou-Peindara, P.; Aas, P.A.; Jacquot, G.; Otterlei, M.; Slupphaug, G.; Benichou, S. Human Immunodeficiency Virus Type 1 Vpr Modulates Cellular Expression of UNG2 via a Negative Transcriptional Effect. J. Virol. 2009, 83, 10256–10263. [Google Scholar] [CrossRef] [PubMed]
  25. Hrecka, K.; Hao, C.; Shun, M.C.; Kaur, S.; Swanson, S.K.; Florens, L.; Washburn, M.P.; Skowronski, J. HIV-1 and HIV-2 Exhibit Divergent Interactions with HLTF and UNG2 DNA Repair Proteins. Proc. Natl. Acad. Sci. USA 2016, 113, E3921–E3930. [Google Scholar] [CrossRef] [PubMed]
  26. Yan, J.; Shun, M.C.; Zhang, Y.; Hao, C.; Skowronski, J. HIV-1 Vpr Counteracts HLTF-Mediated Restriction of HIV-1 Infection in T Cells. Proc. Natl. Acad. Sci. USA 2019, 116, 9568–9577. [Google Scholar] [CrossRef] [PubMed]
  27. Yan, J.; Shun, M.-C.; Hao, C.; Zhang, Y.; Qian, J.; Hrecka, K.; Delucia, M.; Monnie, C.; Ahn, J.; Skowronski, J. HIV-1 Vpr Reprograms CLR4 DCAF1 E3 Ubiquitin Ligase to Antagonize Exonuclease 1-Mediated Restriction of HIV-1 Infection. mBio 2018, 9, e01732-18. [Google Scholar] [CrossRef] [PubMed]
  28. Lahouassa, H.; Blondot, M.L.; Chauveau, L.; Chougui, G.; Morel, M.; Leduc, M.; Guillonneau, F.; Ramirez, B.C.; Schwartz, O.; Margottin-Goguet, F. HIV-1 Vpr Degrades the HLTF DNA Translocase in T Cells and Macrophages. Proc. Natl. Acad. Sci. USA 2016, 113, 5311–5316. [Google Scholar] [CrossRef] [PubMed]
  29. Müller, T.G.; Zila, V.; Müller, B.; Kräusslich, H.-G. Nuclear Capsid Uncoating and Reverse Transcription of HIV-1. Annu. Rev. Virol. 2022, 9, 261–284. [Google Scholar] [CrossRef]
  30. Selyutina, A.; Simons, L.M.; Kirby, K.A.; Bulnes-Ramos, A.; Hu, P.; Sarafianos, S.G.; Hultquist, J.F.; Diaz-Griffero, F. TRIM5α Restriction of HIV-1-N74D Viruses in Lymphocytes Is Caused by a Loss of Cyclophilin A Protection. Viruses 2022, 14, 363. [Google Scholar] [CrossRef]
  31. Chintala, K.; Mohareer, K.; Banerjee, S. Dodging the Host Interferon-Stimulated Gene Mediated Innate Immunity by HIV-1: A Brief Update on Intrinsic Mechanisms and Counter-Mechanisms. Front. Immunol. 2021, 12, 716927. [Google Scholar] [CrossRef]
  32. Fouchier, R.A.M.; Meyer, B.E.; Simon, J.H.M.; Fischer, U.; Albright, A.V.; Gonzalez-Scarano, F.; Malim, M.H. Interaction of the Human Immunodeficiency Virus Type 1 Vpr Protein with the Nuclear Pore Complex. J. Virol. 1998, 72, 6004–6013. [Google Scholar] [CrossRef] [PubMed]
  33. Blanco-Rodriguez, G.; Gazi, A.; Monel, B.; Frabetti, S.; Scoca, V.; Mueller, F.; Schwartz, O.; Krijnse-Locker, J.; Charneau, P.; Di Nunzio, F. Remodeling of the Core Leads HIV-1 Preintegration Complex into the Nucleus of Human Lymphocytes. J. Virol. 2020, 94, e00135-20. [Google Scholar] [CrossRef] [PubMed]
  34. Sapp, N.; Burge, N.; Cox, K.; Prakash, P.; Balasubramaniam, M.; Thapa, S.; Christensen, D.; Li, M.; Linderberger, J.; Kvaratskhelia, M.; et al. HIV-1 Preintegration Complex Preferentially Integrates the Viral DNA into Nucleosomes Containing Trimethylated Histone 3-Lysine 36 Modification and Flanking Linker DNA. J. Virol. 2022, 96, e101122. [Google Scholar] [CrossRef]
  35. Iijima, S.; Nitahara-Kasahara, Y.; Kimata, K.; Zhong Zhuang, W.; Kamata, M.; Isogai, M.; Miwa, M.; Tsunetsugu-Yokota, Y.; Aida, Y. Nuclear Localization of Vpr Is Crucial for the Efficient Replication of HIV-1 in Primary CD4 + T Cells. Virology 2004, 327, 249–261. [Google Scholar] [CrossRef] [PubMed]
  36. Wallet, C.; Rohr, O.; Schwartz, C. Evolution of a Concept: From Accessory Protein to Key Virulence Factor, the Case of HIV-1 Vpr. Biochem. Pharmacol. 2020, 180, 114128. [Google Scholar] [CrossRef]
  37. Popov, S.; Rexach, M.; Zybarth, G.; Reiling, N.; Lee, M.-A.; Ratner, L.; Lane, C.; Shannon Moore, M.; Blobel, G.; Bukrinsky, M. Viral Protein R Regulates Nuclear Import of the HIV-1 Pre-Integration Complex. EMBO J. 1998, 17, 909–917. [Google Scholar] [CrossRef]
  38. Hadpech, S.; Moonmuang, S.; Chupradit, K.; Yasamut, U.; Tayapiwatana, C. Updating on Roles of HIV Intrinsic Factors: A Review of Their Antiviral Mechanisms and Emerging Functions. Intervirology 2022, 65, 67–79. [Google Scholar] [CrossRef] [PubMed]
  39. Gifford, L.B.; Melikyan, G.B. HIV-1 Capsid Uncoating Is a Multistep Process That Proceeds through Defect Formation Followed by Disassembly of the Capsid Lattice. ACS Nano 2023, 18, 2928–2947. [Google Scholar] [CrossRef]
  40. Desai, T.M.; Marin, M.; Sood, C.; Shi, J.; Nawaz, F.; Aiken, C.; Melikyan, G.B. Fluorescent Protein-Tagged Vpr Dissociates from HIV-1 Core after Viral Fusion and Rapidly Enters the Cell Nucleus. Retrovirology 2015, 12, 99. [Google Scholar] [CrossRef]
  41. Soto, P.C.; Terry, V.H.; Lewinski, M.K.; Deshmukh, S.; Beliakova-Bethell, N.; Spina, C.A. HIV-1 Latency Is Established Preferentially in Minimally Activated and Non-Dividing Cells during Productive Infection of Primary CD4 T Cells. PLoS ONE 2022, 17, e271674. [Google Scholar] [CrossRef]
  42. Brenchley, J.M.; Hill, B.J.; Ambrozak, D.R.; Price, D.A.; Guenaga, F.J.; Casazza, J.P.; Kuruppu, J.; Yazdani, J.; Migueles, S.A.; Connors, M.; et al. T-Cell Subsets That Harbor Human Immunodeficiency Virus (HIV) In Vivo: Implications for HIV Pathogenesis. J. Virol. 2004, 78, 1160–1168. [Google Scholar] [CrossRef] [PubMed]
  43. Koelsch, K.K.; Liu, L.; Haubrich, R.; May, S.; Havlir, D.; Günthard, H.F.; Ignacio, C.C.; Campos-Soto, P.; Little, S.J.; Shafer, R.; et al. Dynamics of Total, Linear Nonintegrated, and Integrated HIV-1 DNA in Vivo and in Vitro. J. Infect. Dis. 2008, 197, 411–419. [Google Scholar] [CrossRef] [PubMed]
  44. Martinez-Picado, J.; Zurakowski, R.; Buzón, M.J.; Stevenson, M. Episomal HIV-1 DNA and Its Relationship to Other Markers of HIV-1 Persistence. Retrovirology 2018, 15, 15. [Google Scholar] [CrossRef] [PubMed]
  45. Pinzone, M.R.; Weissman, S.; Pasternak, A.O.; Zurakowski, R.; Migueles, S.; O’Doherty, U. Naive Infection Predicts Reservoir Diversity and Is a Formidable Hurdle to HIV Eradication. JCI Insight 2021, 6, e150794. [Google Scholar] [CrossRef] [PubMed]
  46. Rullo, E.V.; Pinzone, M.R.; Cannon, L.M.; Weissman, S.; Ceccarelli, M.; Zurakowski, R.; Nunnari, G.; O’Doherty, U. Persistence of an Intact HIV Reservoir in Phenotypically Naive T Cells. JCI Insight 2020, 5, e133157. [Google Scholar] [CrossRef] [PubMed]
  47. Poon, B.; Chen, I.S.Y. Human Immunodeficiency Virus Type 1 (HIV-1) Vpr Enhances Expression from Unintegrated HIV-1 DNA. J. Virol. 2003, 77, 3962–3972. [Google Scholar] [CrossRef]
  48. Bruner, K.M.; Wang, Z.; Simonetti, F.R.; Bender, A.M.; Kwon, K.J.; Sengupta, S.; Fray, E.J.; Beg, S.A.; Antar, A.A.R.; Jenike, K.M.; et al. A Quantitative Approach for Measuring the Reservoir of Latent HIV-1 Proviruses. Nature 2019, 566, 120–125. [Google Scholar] [CrossRef]
  49. Dupont, L.; Bloor, S.; Williamson, J.C.; Cuesta, S.M.; Shah, R.; Teixeira-Silva, A.; Naamati, A.; Greenwood, E.J.D.; Sarafianos, S.G.; Matheson, N.J.; et al. The SMC5/6 Complex Compacts and Silences Unintegrated HIV-1 DNA and Is Antagonized by Vpr. Cell Host Microbe 2021, 29, 792–805.e6. [Google Scholar] [CrossRef]
  50. Nikolaitchik, O.A.; Islam, S.; Kitzrow, J.P.; Duchon, A.; Cheng, Z.; Liu, Y.; Rawson, J.M.O.; Shao, W.; Nikolaitchik, M.; Kearney, M.F.; et al. HIV-1 Usurps Transcription Start Site Heterogeneity of Host RNA Polymerase II to Maximize Replication Fitness. Proc. Natl. Acad. Sci. USA 2023, 120, e2305103120. [Google Scholar] [CrossRef]
  51. Dutilleul, A.; Rodari, A.; Van Lint, C. Depicting HIV-1 Transcriptional Mechanisms: A Summary of What We Know. Viruses 2020, 12, 1385. [Google Scholar] [CrossRef]
  52. Chen, J.; He, Y.; Zhong, H.; Hu, F.; Li, Y.; Zhang, Y.; Zhang, X.; Lin, W.; Li, Q.; Xu, F.; et al. Transcriptome Analysis of CD4+ T Cells from HIV-Infected Individuals Receiving ART with LLV Revealed Novel Transcription Factors Regulating HIV-1 Promoter Activity. Virol. Sin. 2023, 38, 398–408. [Google Scholar] [CrossRef]
  53. Forget, J.; Yao, X.-J.; Mercier, J.; Cohen, E. Human Immunodeficiency Virus Type 1 Vpr Protein Transactivation Function: Mechanism and Identification of Domains Involved. J. Mol. Biol. 1998, 284, 915–923. [Google Scholar] [CrossRef] [PubMed]
  54. McAllister, J.J.; Dahiya, S.; Berman, R.; Collins, M.; Nonnemacher, M.R.; Burdo, T.H.; Wigdahl, B. Altered Recruitment of Sp Isoforms to HIV-1 Long Terminal Repeat between Differentiated Monoblastic Cell Lines and Primary Monocyte-Derived Macrophages. Front. Virol. 2022, 2, 971293. [Google Scholar] [CrossRef]
  55. Thieu, K.P.; Morrow, M.P.; Shedlock, D.J.; Schoenly, K.A.; Mallilankaraman, K.; Choo, A.Y.; Fagone, P.; Weiner, D.B.; Muthumani, K. HIV-1 Vpr: Regulator of Viral Survival. Curr. HIV Res. 2009, 7, 153–162. [Google Scholar] [CrossRef] [PubMed]
  56. McAllister, J.J.; Phillips, D.; Millhouse, S.; Conner, J.; Hogan, T.; Ross, H.L.; Wigdahl, B. Analysis of the HIV-1 LTR NF-ΚB-Proximal Sp Site III: Evidence for Cell Type-Specific Gene Regulation and Viral Replication. Virology 2000, 274, 262–277. [Google Scholar] [CrossRef] [PubMed]
  57. Hotter, D.; Bosso, M.; Jønsson, K.L.; Krapp, C.; Stürzel, C.M.; Das, A.; Littwitz-Salomon, E.; Berkhout, B.; Russ, A.; Wittmann, S.; et al. IFI16 Targets the Transcription Factor Sp1 to Suppress HIV-1 Transcription and Latency Reactivation. Cell Host Microbe 2019, 25, 858–872.e13. [Google Scholar] [CrossRef] [PubMed]
  58. Felzien, L.K.; Woffendin, C.; Hottiger, M.O.; Subbramanian, R.A.; Cohen, E.A.; Nabel, G.J. HIV Transcriptional Activation by the Accessory Protein, VPR, Is Mediated by the P300 Co-Activator. Med. Sci. 1998, 95, 5281–5286. [Google Scholar] [CrossRef] [PubMed]
  59. Goh, W.C.; Rogel, M.E.; Kinsey, C.M.; Michael, S.F.; Fultz, P.N.; Nowak, M.A.; Hahn, B.H.; Emerman, M. HIV-1 Vpr Increases Viral Expression by Manipulation of the Cell Cycle: A Mechanism for Selection of Vpr in Vivo. Nat. Med. 1998, 4, 65–71. [Google Scholar] [CrossRef]
  60. Brasey, A.; Lopez-Lastra, M.; Ohlmann, T.; Beerens, N.; Berkhout, B.; Darlix, J.-L.; Sonenberg, N. The Leader of Human Immunodeficiency Virus Type 1 Genomic RNA Harbors an Internal Ribosome Entry Segment That Is Active during the G2/M Phase of the Cell Cycle. J. Virol. 2003, 77, 3939–3949. [Google Scholar] [CrossRef]
  61. Andersen, J.L.; Le Rouzic, E.; Planelles, V. HIV-1 Vpr: Mechanisms of G2 Arrest and Apoptosis. Exp. Mol. Pathol. 2008, 85, 2–10. [Google Scholar] [CrossRef]
  62. DeHart, J.L.; Zimmerman, E.S.; Ardon, O.; Monteiro-Filho, C.M.R.; Argañaraz, E.R.; Planelles, V. HIV-1 Vpr Activates the G2 Checkpoint through Manipulation of the Ubiquitin Proteasome System. Virol. J. 2007, 4, 57. [Google Scholar] [CrossRef] [PubMed]
  63. Jowett, J.B.M.; Planelles, V.; Poon, B.; Shah, N.P.; Chen, M.-L.; Chen, I.S.Y. The Human Immunodeficiency Virus Type 1 Vpr Gene Arrests Infected T Cells in the G2/M Phase of the Cell Cycle. J. Virol. 1995, 69, 6304–6313. [Google Scholar] [CrossRef] [PubMed]
  64. Zhang, F.; Bieniasz, P.D. HIV-1 Vpr Induces Cell Cycle Arrest and Enhances Viral Gene Expression by Depleting CCDC137. Elife 2020, 9, e55806-22. [Google Scholar] [CrossRef] [PubMed]
  65. Bruce, J.W.; Bracken, M.; Evans, E.; Sherer, N.; Ahlquist, P. ZBTB2 Represses HIV-1 Transcription and Is Regulated by HIV-1 Vpr and Cellular DNA Damage Responses. PLoS Pathog. 2021, 17, e1009364. [Google Scholar] [CrossRef] [PubMed]
  66. Li, D.; Lopez, A.; Sandoval, C.; Doyle, R.N.; Fregoso, O.I. HIV Vpr Modulates the Host DNA Damage Response at Two Independent Steps to Damage DNA and Repress Double-Strand DNA Break Repair. mBio 2020, 11, e00940-20. [Google Scholar] [CrossRef] [PubMed]
  67. Nakai-Murakami, C.; Shimura, M.; Kinomoto, M.; Takizawa, Y.; Tokunaga, K.; Taguchi, T.; Hoshino, S.; Miyagawa, K.; Sata, T.; Kurumizaka, H.; et al. HIV-1 Vpr Induces ATM-Dependent Cellular Signal with Enhanced Homologous Recombination. Oncogene 2007, 26, 477–486. [Google Scholar] [CrossRef]
  68. Laguette, N.; Brégnard, C.; Hue, P.; Basbous, J.; Yatim, A.; Larroque, M.; Kirchhoff, F.; Constantinou, A.; Sobhian, B.; Benkirane, M. Premature Activation of the SLX4 Complex by Vpr Promotes G2/M Arrest and Escape from Innate Immune Sensing. Cell 2014, 156, 134–145. [Google Scholar] [CrossRef]
  69. Zhou, X.; Monnie, C.; DeLucia, M.; Ahn, J. HIV-1 Vpr Activates Host CRL4-DCAF1 E3 Ligase to Degrade Histone Deacetylase SIRT7. Virol. J. 2021, 18, 48. [Google Scholar] [CrossRef]
  70. Murakami, T.; Matsuura, R.; Chutiwitoonchai, N.; Takei, M.; Aida, Y. Huntingtin-Interacting Protein 1 Promotes Vpr-Induced G2 Arrest and Hiv-1 Infection in Macrophages. Viruses 2021, 13, 2308. [Google Scholar] [CrossRef]
  71. Barbosa, J.A.F.; Sparapani, S.; Boulais, J.; Lodge, R.; Cohen, É.A. Human Immunodeficiency Virus Type 1 Vpr Mediates Degradation of APC1, a Scaffolding Component of the Anaphase-Promoting Complex/Cyclosome. J. Virol. 2021, 95, e0097120. [Google Scholar] [CrossRef]
  72. Park, J.-E.; Kim, T.-S.; Zeng, Y.; Monnie, C.M.; Alam, M.S.; Zhou, M.; Mikolaj, M.; Maldarelli, F.; Narayan, K.; Ahn, J.; et al. Centrosome Amplification and Aneuploidy Driven by the HIV-1-Induced Vpr•VprBP•Plk4 Complex in CD4+ T Cells. ResearchSquare, 2023; Preprint. [Google Scholar] [CrossRef]
  73. Wang, X.; Singh, S.; Jung, H.Y.; Yang, G.; Jun, S.; Sastry, K.J.; Park, J. Il HIV-1 Vpr Protein Inhibits Telomerase Activity via the EDD-DDB1-VPRBP E3 Ligase Complex. J. Biol. Chem. 2013, 288, 15474–15480. [Google Scholar] [CrossRef]
  74. Morales, D.R.; Moreno-Martos, D.; Matin, N.; McGettigan, P. Health Conditions in Adults with HIV Compared with the General Population: A Population-Based Cross-Sectional Analysis. EClinicalMedicine 2022, 47, 101392. [Google Scholar] [CrossRef]
  75. Chang, H.; Siarot, L.; Matsuura, R.; Lo, C.W.; Sato, H.; Otsuki, H.; Aida, Y. Distinct MCM10 Proteasomal Degradation Profiles by Primate Lentiviruses Vpr Proteins. Viruses 2020, 12, 98. [Google Scholar] [CrossRef]
  76. Zeitlin, S.G.; Chapados, B.R.; Baker, N.M.; Tai, C.; Slupphaug, G.; Wang, J.Y.J. Uracil DNA N-Glycosylase Promotes Assembly of Human Centromere Protein A. PLoS ONE 2011, 6, e17151. [Google Scholar] [CrossRef]
  77. Andersen, J.L.; DeHart, J.L.; Zimmerman, E.S.; Ardon, O.; Kim, B.; Jacquot, G.; Benichou, S.; Planelles, V. HIV-1 Vpr-Induced Apoptosis Is Cell Cycle Dependent and Requires Bax but Not ANT. PLoS Pathog. 2006, 2, 1106–1119. [Google Scholar] [CrossRef]
  78. Roshal, M.; Kim, B.; Zhu, Y.; Nghiem, P.; Planelles, V. Activation of the ATR-Mediated DNA Damage Response by the HIV-1 Viral Protein R. J. Biol. Chem. 2003, 278, 25879–25886. [Google Scholar] [CrossRef] [PubMed]
  79. Lai, M.; Zimmerman, E.S.; Planelles, V.; Chen, J. Activation of the ATR Pathway by Human Immunodeficiency Virus Type 1 Vpr Involves Its Direct Binding to Chromatin In Vivo. J. Virol. 2005, 79, 15443–15451. [Google Scholar] [CrossRef] [PubMed]
  80. Sato, K.; Misawa, N.; Iwami, S.; Satou, Y.; Matsuoka, M.; Ishizaka, Y.; Ito, M.; Aihara, K.; An, D.S.; Koyanagi, Y. HIV-1 Vpr Accelerates Viral Replication during Acute Infection by Exploitation of Proliferating CD4+ T Cells In Vivo. PLoS Pathog. 2013, 9, e1003812. [Google Scholar] [CrossRef] [PubMed]
  81. Bauby, H.; Ward, C.C.; Hugh-White, R.; Swanson, C.M.; Schulz, R.; Goujon, C.; Malim, M.H. HIV-1 Vpr Induces Widespread Transcriptomic Changes in CD4+ T Cells Early Postinfection. mBio 2021, 12, e0136921. [Google Scholar] [CrossRef] [PubMed]
  82. Qiao, J.; Peng, Q.; Qian, F.; You, Q.; Feng, L.; Hu, S.; Liu, W.; Huang, L.; Shu, X.; Sun, B. HIV-1 Vpr Protein Upregulates MicroRNA-210-5p Expression to Induce G2 Arrest by Targeting TGIF2. PLoS ONE 2021, 16, 0261971. [Google Scholar] [CrossRef] [PubMed]
  83. Damba, J.J.; Laskine, M.; Jin, Y.; Sinyavskaya, L.; Durand, M. Incidence of Autoimmune Diseases in People Living with HIV Compared to a Matched Population: A Cohort Study. Clin. Rheumatol. 2021, 40, 2439–2445. [Google Scholar] [CrossRef] [PubMed]
  84. Emery, A.; Swanstrom, R. HIV-1: To Splice or Not to Splice, That Is the Question. Viruses 2021, 13, 181. [Google Scholar] [CrossRef]
  85. Nguyen Quang, N.; Goudey, S.; Ségéral, E.; Mohammad, A.; Lemoine, S.; Blugeon, C.; Versapuech, M.; Paillart, J.C.; Berlioz-Torrent, C.; Emiliani, S.; et al. Dynamic Nanopore Long-Read Sequencing Analysis of HIV-1 Splicing Events during the Early Steps of Infection. Retrovirology 2020, 17, 25. [Google Scholar] [CrossRef] [PubMed]
  86. Kuramitsu, M.; Hashizume, C.; Yamamoto, N.; Azuma, A.; Kamata, M.; Yamamoto, N.; Tanaka, Y.; Aida, Y. A Novel Role for Vpr of Human Immunodeficiency Virus Type 1 as a Regulator of the Splicing of Cellular Pre-MRNA. Microbes Infect. 2005, 7, 1150–1160. [Google Scholar] [CrossRef]
  87. Zhang, X.; Aida, Y. HIV-1 Vpr: A Novel Role in Regulating RNA Splicing. Curr. HIV Res. 2009, 7, 163–168. [Google Scholar] [CrossRef]
  88. Baeyens, A.; Naessens, E.; Van Nuffel, A.; Weening, K.E.; Reilly, A.M.; Claeys, E.; Trypsteen, W.; Vandekerckhove, L.; Eyckerman, S.; Gevaert, K.; et al. HIV-1 Vpr N-Terminal Tagging Affects Alternative Splicing of the Viral Genome. Sci. Rep. 2016, 6, 34573. [Google Scholar] [CrossRef]
  89. Labaronne, E.; Décimo, D.; Bertrand, L.; Guiguettaz, L.; Sohier, T.J.; Cluet, D.; Vivet-Boubou, V.; Dahoui, C.; François, P.; Hatin, I.; et al. Extensive UORF Translation from HIV-1 Transcripts Conditions DDX3 Dependency for Expression of Main ORFs and Elicits Specific T Cell Immune Responses in Infected Individuals. bioRxiv 2022. [Google Scholar] [CrossRef]
  90. Park, J.T.; Oh, S. The Translational Landscape as Regulated by the RNA Helicase DDX3. BMB Rep. 2022, 55, 125–135. [Google Scholar] [CrossRef]
  91. Rao, S.; Lungu, C.; Crespo, R.; Steijaert, T.H.; Gorska, A.; Palstra, R.J.; Prins, H.A.B.; van Ijcken, W.; Mueller, Y.M.; van Kampen, J.J.A.; et al. Selective Cell Death in HIV-1-Infected Cells by DDX3 Inhibitors Leads to Depletion of the Inducible Reservoir. Nat. Commun. 2021, 12, 2475. [Google Scholar] [CrossRef]
  92. Sharma, A.; Yilmaz, A.; Marsh, K.; Cochrane, A.; Boris-Lawrie, K. Thriving under Stress: Selective Translation of HIV-1 Structural Protein MRNA during Vpr-Mediated Impairment of EiF4E Translation Activity. PLoS Pathog. 2012, 8, e1002612. [Google Scholar] [CrossRef]
  93. de Breyne, S.; Ohlmann, T. Focus on Translation Initiation of the HIV-1 MRNAs. Int. J. Mol. Sci. 2019, 20, 101. [Google Scholar] [CrossRef] [PubMed]
  94. Singh, G.; Seufzer, B.; Song, Z.; Zucko, D.; Heng, X.; Boris-Lawrie, K.; Goff, S. HIV-1 Hypermethylated Guanosine Cap Licenses Specialized Translation Unaffected by MTOR. Proc. Natl. Acad. Sci. USA 2021, 119, e2105153118. [Google Scholar] [CrossRef]
  95. Bekele, Y.; Lakshmikanth, T.; Chen, Y.; Mikes, J.; Nasi, A.; Petkov, S.; Hejdeman, B.; Brodin, P.; Chiodi, F. Mass Cytometry Identifies Distinct CD4+ T Cell Clusters Distinguishing HIV-1–Infected Patients According to Antiretroviral Therapy Initiation. JCI Insight 2019, 4, e125442. [Google Scholar] [CrossRef] [PubMed]
  96. Shan, L.; Deng, K.; Gao, H.; Xing, S.; Capoferri, A.A.; Durand, C.M.; Rabi, S.A.; Laird, G.M.; Kim, M.; Hosmane, N.N.; et al. Transcriptional Reprogramming during Effector-to-Memory Transition Renders CD4+ T Cells Permissive for Latent HIV-1 Infection. Immunity 2017, 47, 766–775.e3. [Google Scholar] [CrossRef]
  97. Reuschl, A.K.; Mesner, D.; Shivkumar, M.; Whelan, M.V.X.; Pallett, L.J.; Guerra-Assunção, J.A.; Madansein, R.; Dullabh, K.J.; Sigal, A.; Thornhill, J.P.; et al. HIV-1 Vpr Drives a Tissue Residency-like Phenotype during Selective Infection of Resting Memory T Cells. Cell Rep. 2022, 39, 110650. [Google Scholar] [CrossRef]
  98. Ding, Z.C.; Shi, H.; Aboelella, N.S.; Fesenkova, K.; Park, E.J.; Liu, Z.; Pei, L.; Li, J.; McIndoe, R.A.; Xu, H.; et al. Persistent STAT5 Activation Reprograms the Epigenetic Landscape in CD4+ T Cells to Drive Polyfunctionality and Antitumor Immunity. Sci. Immunol. 2020, 5, eaba5962. [Google Scholar] [CrossRef] [PubMed]
  99. Mutascio, S.; Mota, T.; Franchitti, L.; Sharma, A.A.; Willemse, A.; Bergstresser, S.N.; Wang, H.; Statzu, M.; Tharp, G.K.; Weiler, J.; et al. CD8+ T Cells Promote HIV Latency by Remodeling CD4+ T Cell Metabolism to Enhance Their Survival, Quiescence, and Stemness. Immunity 2023, 56, 1132–1147.e6. [Google Scholar] [CrossRef]
  100. Kumar, N.A.; Cheong, K.; Powell, D.R.; da Fonseca Pereira, C.; Anderson, J.; Evans, V.A.; Lewin, S.R.; Cameron, P.U. The Role of Antigen Presenting Cells in the Induction of HIV-1 Latency in Resting CD4+ T-Cells. Retrovirology 2015, 12, 76. [Google Scholar] [CrossRef]
  101. Naamati, A.; Williamson, J.C.; Greenwood, E.J.; Marelli, S.; Lehner, P.J.; Matheson, N.J. Functional Proteomic Atlas of HIV Infection in Primary Human CD4+ T Cells. Elife 2019, 8, e41431. [Google Scholar] [CrossRef]
  102. Varin, A.; Decrion, A.Z.; Sabbah, E.; Quivy, V.; Sire, J.; Van Lint, C.; Roques, B.P.; Aggarwal, B.B.; Herbein, G. Synthetic Vpr Protein Activates Activator Protein-1, c-Jun N-Terminal Kinase, and NF-ΚB and Stimulates HIV-1 Transcription in Promonocytic Cells and Primary Macrophages. J. Biol. Chem. 2005, 280, 42557–42567. [Google Scholar] [CrossRef]
  103. Muthumani, K.; Choo, A.Y.; Hwang, D.S.; Dayes, N.S.; Chattergoon, M.; Mayilvahanan, S.; Thieu, K.P.; Buckley, P.T.; Emmanuel, J.; Premkumar, A.; et al. HIV-1 Viral Protein-R (VPR) Protects against Lethal Superantigen Challenge While Maintaining Homeostatic T Cell Levels in Vivo. Mol. Ther. 2005, 12, 910–921. [Google Scholar] [CrossRef] [PubMed]
  104. Vogel, O.A.; Forwood, J.K.; Leung, D.W.; Amarasinghe, G.K.; Basler, C.F. Viral Targeting of Importin Alpha-Mediated Nuclear Import to Block Innate Immunity. Cells 2023, 13, 71. [Google Scholar] [CrossRef]
  105. Ayyavoo, V.; Mahboubi, A.; Mahalingam, S.; Ramalingam, R.; Kudchodkar, S.; Williams, W.V.; Green, D.R.; Weiner, D.B. HIV-1 Vpr Suppresses Immune Activation and Apoptosis through Regulation of Nuclear Factor KB. Nat. Med. 1997, 2, 1117–1123. [Google Scholar] [CrossRef] [PubMed]
  106. Kogan, M.; Deshmane, S.; Sawaya, B.E.; Gracely, E.J.; Khalili, K.; Rappaport, J. Inhibition of NF-ΚB Activity by HIV-1 Vpr Is Dependent on Vpr Binding Protein. J. Cell Physiol. 2013, 228, 781–790. [Google Scholar] [CrossRef]
  107. Khan, H.; Sumner, R.P.; Rasaiyaah, J.; Tan, C.P.; Rodriguez-Plata, M.T.; Van Tulleken, C.; Fink, D.; Zuliani-Alvarez, L.; Thorne, L.; Stirling, D.; et al. HIV-1 Vpr Antagonizes Innate Immune Activation by Targeting Karyopherin-Mediated NF-ΚB/IRF3 Nuclear Transport. Elife 2020, 9, e60821. [Google Scholar] [CrossRef] [PubMed]
  108. Hotter, D.; Krabbe, T.; Reith, E.; Gawanbacht, A.; Rahm, N.; Ayouba, A.; Van Driessche, B.; Van Lint, C.; Peeters, M.; Kirchhoff, F.; et al. Primate Lentiviruses Use at Least Three Alternative Strategies to Suppress NF-ΚB-Mediated Immune Activation. PLoS Pathog. 2017, 13, e1006598. [Google Scholar] [CrossRef]
  109. Liu, R.; Lin, Y.; Jia, R.; Geng, Y.; Liang, C.; Tan, J.; Qiao, W. HIV-1 Vpr Stimulates NF-ΚB and AP-1 Signaling by Activating TAK1. Retrovirology 2014, 11, 45. [Google Scholar] [CrossRef]
  110. Liu, R.; Tan, J.; Lin, Y.; Jia, R.; Yang, W.; Liang, C.; Geng, Y.; Qiao, W. HIV-1 Vpr Activates Both Canonical and Noncanonical NF-ΚB Pathway by Enhancing the Phosphorylation of IKKα/Β. Virology 2013, 439, 47–56. [Google Scholar] [CrossRef]
  111. Wong, L.M.; Jiang, G. NF-ΚB Sub-Pathways and HIV Cure: A Revisit. EBioMedicine 2021, 63, 103159. [Google Scholar] [CrossRef]
  112. Campbell, G.R.; Bruckman, R.S.; Chu, Y.L.; Trout, R.N.; Spector, S.A. SMAC Mimetics Induce Autophagy-Dependent Apoptosis of HIV-1-Infected Resting Memory CD4+ T Cells. Cell Host Microbe 2018, 24, 689–702.e7. [Google Scholar] [CrossRef]
  113. Konrath, F.; Willenbrock, M.; Busse, D.; Scheidereit, C.; Wolf, J. A Computational Model of the DNA Damage-Induced IKK/ NF-ΚB Pathway Reveals a Critical Dependence on Irradiation Dose and PARP-1. iScience 2023, 26, 107917. [Google Scholar] [CrossRef]
  114. Liang, Z.; Liu, R.; Lin, Y.; Liang, C.; Tan, J.; Qiao, W. HIV-1 Vpr Protein Activates the NF-ΚB Pathway to Promote G2/M Cell Cycle Arrest. Virol. Sin. 2015, 30, 441–448. [Google Scholar] [CrossRef]
  115. Sandoval, C.; Fregoso, O.I. HIV-1 Vpr-Induced DNA Damage Activates NF-ΚB Independent of Cell Cycle Arrest and CRL4A DCAF1 Engagement. bioRxiv 2023. [Google Scholar] [CrossRef]
  116. Bosso, M.; Kirchhoff, F. Emerging Role of PYHIN Proteins as Antiviral Restriction Factors. Viruses 2020, 12, 1464. [Google Scholar] [CrossRef] [PubMed]
  117. Bosso, M.; Bozzo, C.P.; Hotter, D.; Volcic, M.; Stürzel, C.M.; Rammelt, A.; Ni, Y.; Urban, S.; Becker, M.; Schelhaas, M.; et al. Nuclear PYHIN Proteins Target the Host Transcription Factor Sp1 Thereby Restricting HIV-1 in Human Macrophages and CD4+ T Cells. PLoS Pathog. 2020, 16, e1008752. [Google Scholar] [CrossRef] [PubMed]
  118. Bosso, M.; Stürzel, C.M.; Kmiec, D.; Badarinarayan, S.S.; Braun, E.; Ito, J.; Sato, K.; Hahn, B.H.; Sparrer, K.M.J.; Sauter, D.; et al. An Additional NF-ΚB Site Allows HIV-1 Subtype C to Evade Restriction by Nuclear PYHIN Proteins. Cell Rep. 2021, 36, 109735. [Google Scholar] [CrossRef] [PubMed]
  119. Heusinger, E.; Kirchhoff, F. Primate Lentiviruses Modulate NF-HIV, SIV, NF-ΚB, Nef, Vpu, Tat, LTRB Activity by Multiple Mechanisms to Fine-Tune Viral and Cellular Gene Expression. Front. Microbiol. 2017, 8, 198. [Google Scholar] [CrossRef] [PubMed]
  120. Höhne, K.; Businger, R.; Van Nuffel, A.; Bolduan, S.; Koppensteiner, H.; Baeyens, A.; Vermeire, J.; Malatinkova, E.; Verhasselt, B.; Schindler, M. Virion Encapsidated HIV-1 Vpr Induces NFAT to Prime Non-Activated T Cells for Productive Infection. Open Biol. 2016, 6, 160046. [Google Scholar] [CrossRef]
  121. Okumura, A.; Alce, T.; Lubyova, B.; Ezelle, H.; Strebel, K.; Pitha, P.M. HIV-1 Accessory Proteins VPR and Vif Modulate Antiviral Response by Targeting IRF-3 for Degradation. Virology 2008, 373, 85–97. [Google Scholar] [CrossRef] [PubMed]
  122. Boso, G.; Kozak, C.A. Retroviral Restriction Factors and Their Viral Targets: Restriction Strategies and Evolutionary Adaptations. Microorganisms 2020, 8, 1965. [Google Scholar] [CrossRef]
  123. Gaba, A.; Flath, B.; Chelico, L. Examination of the Apobec3 Barrier to Cross Species Transmission of Primate Lentiviruses. Viruses 2021, 13, 1084. [Google Scholar] [CrossRef]
  124. Zhou, D.; Wang, Y.; Tokunaga, K.; Huang, F.; Sun, B.; Yang, R. The HIV-1 Accessory Protein Vpr Induces the Degradation of the Anti-HIV-1 Agent APOBEC3G through a VprBP-Mediated Proteasomal Pathway. Virus Res. 2015, 195, 25–34. [Google Scholar] [CrossRef]
  125. Zhao, L.; Wang, S.; Xu, M.; He, Y.; Zhang, X.; Xiong, Y.; Sun, H.; Ding, H.; Geng, W.; Shang, H.; et al. Vpr Counteracts the Restriction of LAPTM5 to Promote HIV-1 Infection in Macrophages. Nat. Commun. 2021, 12, 3691. [Google Scholar] [CrossRef] [PubMed]
  126. Sertznig, H.; Roesmann, F.; Wilhelm, A.; Heininger, D.; Bleekmann, B.; Elsner, C.; Santiago, M.; Schuhenn, J.; Karakoese, Z.; Benatzy, Y.; et al. SRSF1 Acts as an IFN-I-Regulated Cellular Dependency Factor Decisively Affecting HIV-1 Post-Integration Steps. Front. Immunol. 2022, 13, 935800. [Google Scholar] [CrossRef]
  127. Bhardwaj, V.; Singh, A.; Choudhary, A.; Dalavi, R.; Ralte, L.; Chawngthu, R.L.; Kumar, N.S.; Vijay, N.; Chande, A. HIV-1 Vpr Induces CiTRAN to Prevent Transcriptional Repression of the Provirus. Sci. Adv. 2023, 9, eadh9170. [Google Scholar] [CrossRef] [PubMed]
  128. Hofmann, S.; Dehn, S.; Businger, R.; Bolduan, S.; Schneider, M.; Debyser, Z.; Brack-Werner, R.; Schindler, M. Dual Role of the Chromatin-Binding Factor PHF13 in the Pre- and Post-Integration Phases of HIV-1 Replication. Open Biol. 2017, 7. [Google Scholar] [CrossRef] [PubMed]
  129. Chung, H.-R.; Xu, C.; Fuchs, A.; Mund, A.; Staege, H.; Schubert, T.; Lange, M.; Bian, C.; Dunkel, I.; Eberharter, A.; et al. PHF13 Is a Molecular Reader and transcriptional Co-Regulator of H3K4me2/3. Elife 2016, 5. [Google Scholar] [CrossRef] [PubMed]
  130. Rossi, F.; Buschow, R.; Glaser, L.V.; Schubert, T.; Staege, H.; Grimme, A.; Will, H.; Milke, T.; Vingron, M.; Chiariello, A.M.; et al. Connecting the Dots: PHF13 and Cohesin Promote Polymer-Polymer Phase Separation of Chromatin into Chromosomes. bioRxiv 2022. [Google Scholar] [CrossRef]
  131. Maudet, C.; Sourisce, A.; Dragin, L.; Lahouassa, H.; Rain, J.C.; Bouaziz, S.; Ramirez, B.C.; Margottin-Goguet, F. HIV-1 Vpr Induces the Degradation of ZIP and SZIP, Adaptors of the NuRD Chromatin Remodeling Complex, by Hijacking DCAF1/VprBP. PLoS ONE 2013, 8, e77320. [Google Scholar] [CrossRef]
  132. Forouzanfar, F.; Ali, S.; Wallet, C.; De Rovere, M.; Ducloy, C.; El Mekdad, H.; El Maassarani, M.; Aït-Ammar, A.; Van Assche, J.; Boutant, E.; et al. HIV-1 Vpr Mediates the Depletion of the Cellular Repressor CTIP2 to Counteract Viral Gene Silencing. Sci. Rep. 2019, 9, 13154. [Google Scholar] [CrossRef]
  133. Marno, K.M.; Ogunkolade, B.W.; Pade, C.; Oliveira, N.M.; Sullivan, O.; McKnight, Á. Novel Restriction Factor RNA-Associated Early-Stage Anti-Viral Factor (REAF) Inhibits Human and Simian Immunodeficiency Viruses. Retrovirology 2014, 11, 3. [Google Scholar] [CrossRef]
  134. Seczynska, M.; Lehner, P.J. The Sound of Silence: Mechanisms and Implications of HUSH Complex Function. Trends Genet. 2023, 39, 251–267. [Google Scholar] [CrossRef]
  135. Wang, Z.; Song, A.; Xu, H.; Hu, S.; Tao, B.; Peng, L.; Wang, J.; Li, J.; Yu, J.; Wang, L.; et al. Coordinated Regulation of RNA Polymerase II Pausing and Elongation Progression by PAF1. Sci. Adv. 2022, 8, 5504. [Google Scholar] [CrossRef] [PubMed]
  136. Gibbons, J.M.; Marno, K.M.; Pike, R.; Lee, W.J.; Jones, C.E.; Ogunkolade, B.W.; Pardieu, C.; Bryan, A.; Fu, R.M.; Warnes, G.; et al. HIV-1 Accessory Protein Vpr Interacts with REAF/RPRD2 To Mitigate Its Antiviral Activity. J. Virol. 2020, 94, e01591-19. [Google Scholar] [CrossRef] [PubMed]
  137. Jackson-Jones, K.A.; McKnight, Á.; Sloan, R.D. The Innate Immune Factor RPRD2/REAF and Its Role in the Lv2 Restriction of HIV. mBio 2023. [Google Scholar] [CrossRef]
  138. Markle, T.J.; Mwimanzi, P.; Brockman, M.A. HIV-1 Nef and T-Cell Activation: A History of Contradictions. Future Virol. 2013, 8, 391–404. [Google Scholar] [CrossRef] [PubMed]
  139. Schindler, M.; Münch, J.; Kutsch, O.; Li, H.; Santiago, M.L.; Bibollet-Ruche, F.; Müller-Trutwin, M.C.; Novembre, F.J.; Peeters, M.; Courgnaud, V.; et al. Nef-Mediated Suppression of T Cell Activation Was Lost in a Lentiviral Lineage That Gave Rise to HIV-1. Cell 2006, 125, 1055–1067. [Google Scholar] [CrossRef]
  140. Lv, T.; Cao, W.; Li, T. HIV-Related Immune Activation and Inflammation: Current Understanding and Strategies. J. Immunol. Res. 2021, 2021, 7316456. [Google Scholar] [CrossRef]
  141. Roux, P.; Alfieri, C.; Hrimech, M.; Cohen, E.A.; Tanner, J.E. Activation of Transcription Factors NF-B and NF-IL-6 by Human Immunodeficiency Virus Type 1 Protein R (Vpr) Induces Interleukin-8 Expression. J. Virol. 2000, 74, 4658–4665. [Google Scholar] [CrossRef]
  142. Roesch, F.; Richard, L.; Rua, R.; Porrot, F.; Casartelli, N.; Schwartz, O. Vpr Enhances Tumor Necrosis Factor Production by HIV-1-Infected T Cells. J. Virol. 2015, 89, 12118–12130. [Google Scholar] [CrossRef] [PubMed]
  143. Vermeire, J.; Roesch, F.; Sauter, D.; Rua, R.; Hotter, D.; Van Nuffel, A.; Vanderstraeten, H.; Naessens, E.; Iannucci, V.; Landi, A.; et al. HIV Triggers a CGAS-Dependent, Vpu- and Vpr-Regulated Type I Interferon Response in CD4+ T Cells. Cell Rep. 2016, 17, 413–424. [Google Scholar] [CrossRef] [PubMed]
  144. Oriol-Tordera, B.; Berdasco, M.; Llano, A.; Mothe, B.; Gálvez, C.; Martinez-Picado, J.; Carrillo, J.; Blanco, J.; Duran-Castells, C.; Ganoza, C.; et al. Methylation Regulation of Antiviral Host Factors, Interferon Stimulated Genes (ISGs) and T-Cell Responses Associated with Natural HIV Control. PLoS Pathog. 2020, 16, e1008678. [Google Scholar] [CrossRef] [PubMed]
  145. Arokium, H.; Kamata, M.; Chen, I. Virion-Associated Vpr of Human Immunodeficiency Virus Type 1 Triggers Activation of Apoptotic Events and Enhances Fas-Induced Apoptosis in Human T Cells. J. Virol. 2009, 83, 11283–11297. [Google Scholar] [CrossRef] [PubMed]
  146. Conti, L.; Matarrese, P.; Varano, B.; Gauzzi, M.C.; Sato, A.; Malorni, W.; Belardelli, F.; Gessani, S. Dual Role of the HIV-1 Vpr Protein in the Modulation of the Apoptotic Response of T Cells. J. Immunol. 2000, 165, 3293–3300. [Google Scholar] [CrossRef]
  147. Muthumani, K.; Hwang, D.S.; Desai, B.M.; Zhang, D.; Dayes, N.; Green, D.R.; Weiner, D.B. HIV-1 Vpr Induces Apoptosis through Caspase 9 in T Cells and Peripheral Blood Mononuclear Cells. J. Biol. Chem. 2002, 277, 37820–37831. [Google Scholar] [CrossRef]
  148. Huang, C.Y.; Chiang, S.F.; Lin, T.Y.; Chiou, S.H.; Chow, K.C. HIV-1 Vpr Triggers Mitochondrial Destruction by Impairing Mfn2-Mediated ER-Mitochondria Interaction. PLoS ONE 2012, 7, e33657. [Google Scholar] [CrossRef]
  149. Cosentino, K.; Hertlein, V.; Jenner, A.; Dellmann, T.; Gojkovic, M.; Peña-Blanco, A.; Dadsena, S.; Wajngarten, N.; Danial, J.S.H.; Thevathasan, J.V.; et al. The Interplay between BAX and BAK Tunes Apoptotic Pore Growth to Control Mitochondrial-DNA-Mediated Inflammation. Mol. Cell 2022, 82, 933–949.e9. [Google Scholar] [CrossRef]
  150. Wang, Y.C.; Wang, L.T.; Hung, T.I.; Hong, Y.R.; Chen, C.H.; Ho, C.J.; Wang, C. Severe Cellular Stress Drives Apoptosis through a Dual Control Mechanism Independently of P53. Cell Death Discov. 2022, 8, 282. [Google Scholar] [CrossRef] [PubMed]
  151. Sakai, K.; Dimas, J.; Lenardo, M.J. The Vif and Vpr Accessory Proteins Independently Cause HIV-1-Induced T Cell Cytopathicity and Cell Cycle Arrest. Proc. Natl. Acad. Sci. USA 2006, 103, 3369–3374. [Google Scholar] [CrossRef] [PubMed]
  152. Trinité, B.; Chan, C.N.; Lee, C.S.; Levy, D.N. HIV-1 Vpr- and Reverse Transcription-Induced Apoptosis in Resting Peripheral Blood CD4+ T Cells and Protection by Common Gamma-Chain Cytokines. J. Virol. 2016, 90, 904–916. [Google Scholar] [CrossRef] [PubMed]
  153. Levy, D.N.; Refaeli, Y.; MacGregor, R.R.; Weiner, D.B. Serum Vpr Regulates Productive Infection and Latency of Human Immunodeficiency Virus Type 1. Proc. Natl. Acad. Sci. USA 1994, 91, 10873–10877. [Google Scholar] [CrossRef]
  154. Hoshino, S.; Konishi, M.; Mori, M.; Shimura, M.; Nishitani, C.; Kuroki, Y.; Koyanagi, Y.; Kano, S.; Itabe, H.; Ishizaka, Y. HIV-1 Vpr Induces TLR4/MyD88-Mediated IL-6 Production and Reactivates Viral Production from Latency. J. Leukoc. Biol. 2010, 87, 1133–1143. [Google Scholar] [CrossRef] [PubMed]
  155. Finkel, T.H.; Tudor-Williams, G.; Banda, N.K.; Cotion, M.F.; Curiel, T.; Monks, C.; Baba, T.W.; Ruprecht, R.M.; Kupfer, A. Apoptosis Occurs Predominantly in Bystander Cells and Not in Productively Infected Cells of HIV-and SIV-Infected Lymph Nodes. Nat. Med. 1995, 1, 129–134. [Google Scholar] [CrossRef]
  156. Richard, J.; Pham, T.N.Q.; Ishizaka, Y.; Cohen, É.A. Viral Protein R Upregulates Expression of ULBP2 on Uninfected Bystander Cells during HIV-1 Infection of Primary CD4+ T Lymphocytes. Virology 2013, 443, 248–256. [Google Scholar] [CrossRef] [PubMed]
  157. Ward, J.; Davis, Z.; DeHart, J.; Zimmerman, E.; Bosque, A.; Brunetta, E.; Mavilio, D.; Planelles, V.; Barker, E. HIV-1 Vpr Triggers Natural Killer Cell-Mediated Lysis of Infected Cells through Activation of the ATR-Mediated DNA Damage Response. PLoS Pathog. 2009, 5, e1000613. [Google Scholar] [CrossRef]
  158. Vassena, L.; Giuliani, E.; Matusali, G.; Cohen, É.A.; Doria, M. The Human Immunodeficiency Virus Type 1 Vpr Protein Upregulates PVR via Activation of the ATR-Mediated DNA Damage Response Pathway. J. Gen. Virol. 2013, 94, 2664–2669. [Google Scholar] [CrossRef]
  159. Lv, L.; Wang, Q.; Xu, Y.; Tsao, L.C.; Nakagawa, T.; Guo, H.; Su, L.; Xiong, Y. Vpr Targets TET2 for Degradation by CRL4 VprBP E3 Ligase to Sustain IL-6 Expression and Enhance HIV-1 Replication. Mol. Cell 2018, 70, 961–970.e5. [Google Scholar] [CrossRef]
  160. Jacquot, G.; Le Rouzic, E.; Maidou-Peindara, P.; Maizy, M.; Lefrère, J.J.; Daneluzzi, V.; Monteiro-Filho, C.M.R.; Hong, D.; Planelles, V.; Morand-Joubert, L.; et al. Characterization of the Molecular Determinants of Primary HIV-1 Vpr Proteins: Impact of the Q65R and R77Q Substitutions on Vpr Functions. PLoS ONE 2009, 4, e7514. [Google Scholar] [CrossRef]
  161. Soares, R.; Rocha, G.; Nogueira, C.; Meliço-Silvestre, A.; Gonçalves, T. R77Q and Q3R HIV1-VPR Mutations in an Otherwise Asymptomatic 5-year-old Child with Repeated Ear Infections. JMM Case Rep. 2014, 1, e002709. [Google Scholar] [CrossRef]
  162. Macreadie, I.G.; Castelli, L.A.; Hewisht, D.R.; Kirkpatricki, A.; Ward, A.C.; Azad, A.A. A Domain of Human Immunodeficiency Virus Type 1 Vpr Containing Repeated H(S/F)RIG Amino Acid Motifs Causes Cell Growth Arrest and Structural Defects. Proc. Natl. Acad. Sci. USA 1995, 92, 2770–2774. [Google Scholar] [CrossRef]
  163. Maudet, C.; Bertrand, M.; Le Rouzic, E.; Lahouassa, H.; Ayinde, D.; Nisole, S.; Goujon, C.; Cimarelli, A.; Margottin-Goguet, F.; Transy, C. Molecular Insight into How HIV-1 Vpr Protein Impairs Cell Growth through Two Genetically Distinct Pathways. J. Biol. Chem. 2011, 286, 23742–23752. [Google Scholar] [CrossRef] [PubMed]
  164. Rajan, D.; Wildum, S.; Rü, E.; Schindler, M.; Kirchhoff, F. Effect of R77Q, R77A and R80A Changes in Vpr on HIV-1 Replication and CD4 T Cell Depletion in Human Lymphoid Tissue Ex Vivo. AIDS 2006, 20, 831–836. [Google Scholar] [CrossRef] [PubMed]
  165. Byeon, I.J.L.; Calero, G.; Wu, Y.; Byeon, C.H.; Jung, J.; DeLucia, M.; Zhou, X.; Weiss, S.; Ahn, J.; Hao, C.; et al. Structure of HIV-1 Vpr in Complex with the Human Nucleotide Excision Repair Protein HHR23A. Nat. Commun. 2021, 12, 6864. [Google Scholar] [CrossRef] [PubMed]
  166. Bolton, D.L.; Lenardo, M.J. Vpr Cytopathicity Independent of G2/M Cell Cycle Arrest in Human Immunodeficiency Virus Type 1-Infected CD4 + T Cells. J. Virol. 2007, 81, 8878–8890. [Google Scholar] [CrossRef] [PubMed]
  167. Barnitz, R.A.; Chaigne-Delalande, B.; Bolton, D.L.; Lenardo, M.J. Exposed Hydrophobic Residues in Human Immunodeficiency Virus Type 1 Vpr Helix-1 Are Important for Cell Cycle Arrest and Cell Death. PLoS ONE 2011, 6, e24924. [Google Scholar] [CrossRef] [PubMed]
  168. Zhou, Y.I.; Ratner, L. Phosphorylation of Human Immunodeficiency Virus Type 1 Vpr Regulates Cell Cycle Arrest. J. Virol. 2000, 74, 6520–6527. [Google Scholar] [CrossRef]
  169. Barnitz, R.A.; Wan, F.; Tripuraneni, V.; Bolton, D.L.; Lenardo, M.J. Protein Kinase A Phosphorylation Activates Vpr-Induced Cell Cycle Arrest during Human Immunodeficiency Virus Type 1 Infection. J. Virol. 2010, 84, 6410–6424. [Google Scholar] [CrossRef]
  170. Hadi, K.; Walker, L.A.; Guha, D.; Murali, R.; Watkins, S.C.; Tarwater, P.; Srinivasan, A.; Ayyavoo, V. Human Immunodeficiency Virus Type 1 Vpr Polymorphisms Associated with Progressor and Nonprogressor Individuals Alter Vpr-Associated Functions. J. Gen. Virol. 2014, 95, 700–711. [Google Scholar] [CrossRef]
  171. Fritz, J.V.; Didier, P.; Clamme, J.P.; Schaub, E.; Muriaux, D.; Cabanne, C.; Morellet, N.; Bouaziz, S.; Darlix, J.L.; Mély, Y.; et al. Direct Vpr-Vpr Interaction in Cells Monitored by Two Photon Fluorescence Correlation Spectroscopy and Fluorescence Lifetime Imaging. Retrovirology 2008, 5, 87. [Google Scholar] [CrossRef]
  172. Somasundaran, M.; Sharkey, M.; Brichacek, B.; Luzuriaga, K.; Emerman, M.; Sullivan, J.L.; Stevenson, M. Evidence for a Cytopathogenicity Determinant in HIV-1 Vpr. Proc. Natl. Acad. Sci. USA 2002, 99, 9503–9508. [Google Scholar] [CrossRef]
  173. Yao, X.J.; Rougeau, N.; Duisit, G.; Lemay, J.; Cohen, É.A. Analysis of HIV-1 Vpr Determinants Responsible for Cell Growth Arrest in Saccharomyces cerevisiae. Retrovirology 2004, 1, 21. [Google Scholar] [CrossRef]
  174. Rhodes, D.I.; Ashton, L.; Solomon, A.; Carr, A.; Cooper, D.; Kaldor, J. Characterization of Three Nef-Defective Human Immunodeficiency Virus Type 1 Strains Associated with Long-Term Nonprogression. J. Virol. 2000, 74, 10581–10588. [Google Scholar] [CrossRef]
  175. Chen, M.; Elder, R.T.; Yu, M.; O’Gorman, M.G.; Selig, L.; Benarous, R.; Yamamoto, A.; Zhao, Y. Mutational Analysis of Vpr-Induced G2 Arrest, Nuclear Localization, and Cell Death in Fission Yeast. J. Virol. 1999, 73, 3236–3245. [Google Scholar] [CrossRef]
  176. Thotala, D.; Schafer, E.A.; Majumder, B.; Janket, M.L.; Wagner, M.; Srinivasan, A.; Watkins, S.; Ayyavoo, V. Structure-Functional Analysis of Human Immunodeficiency Virus Type 1 (HIV-1) Vpr: Role of Leucine Residues on Vpr-Mediated Transactivation and Virus Replication. Virology 2004, 328, 89–100. [Google Scholar] [CrossRef] [PubMed]
  177. Arora, S.; Verma, S.; Banerjea, A.C. HIV-1 Vpr Redirects Host Ubiquitination Pathway. J. Virol. 2014, 88, 9141–9152. [Google Scholar] [CrossRef] [PubMed]
  178. Bano, A.S.; Sood, V.; Neogi, U.; Goel, N.; Kuttiat, V.S.; Wanchu, A.; Banerjea, A.C. Genetic and Functional Characterization of Human Immunodeficiency Virus Type 1 VprC Variants from North India: Presence of Unique Recombinants with Mosaic Genomes from B, C and D Subtypes within the Open Reading Frame of Vpr. J. Gen. Virol. 2009, 90, 2768–2776. [Google Scholar] [CrossRef] [PubMed]
  179. Fritz, J.V.; Dujardin, D.; Godet, J.; Didier, P.; De Mey, J.; Darlix, J.-L.; Mély, Y.; de Rocquigny, H. HIV-1 Vpr Oligomerization but Not That of Gag Directs the Interaction between Vpr and Gag. J. Virol. 2010, 84, 1585–1596. [Google Scholar] [CrossRef] [PubMed]
  180. Yao, X.-J.; Subbramanian, R.A.; Rougeau, N.; Boisvert, O.; Bergeron, D.; Cohen, E.A. Mutagenic Analysis of Human Immunodeficiency Virus Type 1 Vpr: Role of a Predicted N-Terminal Alpha-Helical Structure in Vpr Nuclear Localization and Virion Incorporation. J. Virol. 1995, 69, 7032–7044. [Google Scholar] [CrossRef]
  181. Kino, T.; Gragerov, A.; Valentin, A.; Tsopanomihalou, M.; Ilyina-Gragerova, G.; Erwin-Cohen, R.; Chrousos, G.P.; Pavlakis, G.N. Vpr Protein of Human Immunodeficiency Virus Type 1 Binds to 14-3-3 Proteins and Facilitates Complex Formation with Cdc25C: Implications for Cell Cycle Arrest. J. Virol. 2005, 79, 2780–2787. [Google Scholar] [CrossRef] [PubMed]
  182. Sörgel, S.; Fraedrich, K.; Votteler, J.; Thomas, M.; Stamminger, T.; Schubert, U. Perinuclear Localization of the HIV-1 Regulatory Protein Vpr Is Important for Induction of G2-Arrest. Virology 2012, 432, 444–451. [Google Scholar] [CrossRef]
  183. Jacquot, G.; Le Rouzic, E.; David, A.; Mazzolini, J.; Bouchet, J.; Bouaziz, S.; Niedergang, F.; Pancino, G.; Benichou, S. Localization of HIV-1 Vpr to the Nuclear Envelope: Impact on Vpr Functions and Virus Replication in Macrophages. Retrovirology 2007, 4, 84. [Google Scholar] [CrossRef]
  184. Selig, L.; Pages, J.-C.; Tanchou, V.; Pré, S.; Ral, V.É.; Berlioz-Torrent, C.; Liu, L.X.; Erdtmann, L.; Darlix, J.-L.; Benarous, R.; et al. Interaction with the P6 Domain of the Gag Precursor Mediates Incorporation into Virions of Vpr and Vpx Proteins from Primate Lentiviruses. J. Virol. 1999, 73, 592–600. [Google Scholar] [CrossRef]
  185. Piller, S.C.; Ewart, G.D.; Jans, D.A.; Gage, P.W.; Cox, G.B. The Amino-Terminal Region of Vpr from Human Immunodeficiency Virus Type 1 Forms Ion Channels and Kills Neurons. J. Virol. 1999, 73, 4230–4238. [Google Scholar] [CrossRef] [PubMed]
  186. Gummuluru, S.; Emerman, M. Cell Cycle-and Vpr-Mediated Regulation of Human Immunodeficiency Virus Type 1 Expression in Primary and Transformed T-Cell Lines. J. Virol. 1999, 73, 5422–5430. [Google Scholar] [CrossRef]
  187. Selig, L.; Benichou, S.; Rogel, M.E.; Wu, L.I.; Vodicka, M.A.; Sire, J.; Benarous, R.; Emerman, M. Uracil DNA Glycosylase Specifically Interacts with Vpr of Both Human Immunodeficiency Virus Type 1 and Simian Immunodeficiency Virus of Sooty Mangabeys, but Binding Does Not Correlate with Cell Cycle Arrest. J. Virol. 1997, 71, 4842–4846. [Google Scholar] [CrossRef]
  188. Watanabe, N.; Nishihara, Y.; Yamaguchi, T.; Koito, A.; Miyoshi, H.; Kakeya, H.; Osada, H. Fumagillin Suppresses HIV-1 Infection of Macrophages through the Inhibition of Vpr Activity. FEBS Lett. 2006, 580, 2598–2602. [Google Scholar] [CrossRef]
  189. Ong, E.B.B.; Watanabe, N.; Saito, A.; Futamura, Y.; Abd El Galil, K.H.; Koito, A.; Najimudin, N.; Osada, H. Vipirinin, a Coumarin-Based HIV-1 Vpr Inhibitor, Interacts with a Hydrophobic Region of Vpr. J. Biol. Chem. 2011, 286, 14049–14056. [Google Scholar] [CrossRef]
  190. Bachand, F.; Yao, X.-J.; Hrimech, M.; Rougeau, N.; Cohen, E.A. Incorporation of Vpr into Human Immunodeficiency Virus Type 1 Requires a Direct Interaction with the P6 Domain of the P55 Gag Precursor*. J. Biol. Chem. 1999, 274, 9083–9091. [Google Scholar] [CrossRef] [PubMed]
  191. Di Marzio, P.; Choe, S.; Ebright, M.; Knoblauch, R.; Landau, N.R. Mutational Analysis of Cell Cycle Arrest, Nuclear Localization, and Virion Packaging of Human Immunodeficiency Virus Type 1 Vpr. J. Virol. 1995, 69, 7909–7916. [Google Scholar] [CrossRef] [PubMed]
  192. Venkatachari, N.J.; Walker, L.A.; Tastan, O.; Le, T.; Dempsey, T.M.; Li, Y.; Yanamala, N.; Srinivasan, A.; Klein-Seetharaman, J.; Montelaro, R.C.; et al. Human Immunodeficiency Virus Type 1 Vpr: Oligomerization Is an Essential Feature for Its Incorporation into Virus Particles. Virol. J. 2010, 7, 119. [Google Scholar] [CrossRef] [PubMed]
  193. Muthumani, K.; Choo, A.Y.; Zong, W.X.; Madesh, M.; Hwang, D.S.; Premkumar, A.; Thieu, K.P.; Emmanuel, J.; Kumar, S.; Thompson, C.B.; et al. The HIV-1 Vpr and Glucocorticoid Receptor Complex Is a Gain-of-Function Interaction That Prevents the Nuclear Localization of PARP-1. Nat. Cell Biol. 2006, 8, 170–179. [Google Scholar] [CrossRef] [PubMed]
  194. Mahalingam, S.; Ayyavoo, V.; Patel, M.; Kieber-Emmons, T.; Weiner, D.B. Nuclear Import, Virion Incorporation, and Cell Cycle Arrest/ Differentiation Are Mediated by Distinct Functional Domains of Human Immunodeficiency Virus Type 1 Vpr. J. Virol. 1997, 71, 6339–6347. [Google Scholar] [CrossRef] [PubMed]
  195. Mahalingam, S.; Ayyavoo, V.; Patel, M.; Kieber-Emmons, T.; Kao, G.D.; Muschel, R.J.; Weiner, D.B. HIV-1 Vpr Interacts with a Human 34-KDa Mov34 Homologue, a Cellular Factor Linked to the G2 M Phase Transition of the Mammalian Cell Cycle. Proc. Natl. Acad. Sci. USA 1998, 95, 3419–3424. [Google Scholar] [CrossRef] [PubMed]
  196. Stewart, S.A.; Poon, B.; Jowett, J.B.M.; Chen, I.S.Y. Human Immunodeficiency Virus Type 1 Vpr Induces Apoptosis Following Cell Cycle Arrest. J. Virol. 1997, 71, 5579–5592. [Google Scholar] [CrossRef]
  197. Chutiwitoonchai, N.; Siarot, L.; Takeda, E.; Shioda, T.; Ueda, M.; Aida, Y. HIV-1 Vpr Abrogates the Effect of TSG101 Overexpression to Support Virus Release. PLoS ONE 2016, 11, e163100. [Google Scholar] [CrossRef]
  198. Nie, Z.; Bergeron, D.; Subbramanian, R.A.; Yao, X.-J.; Checroune, F.; Rougeau, N.; Cohen, E.A. The Putative Alpha Helix 2 of Human Immunodeficiency Virus Type 1 Vpr Contains a Determinant Which Is Responsible for the Nuclear Translocation of Proviral DNA in Growth-Arrested Cells. J. Virol. 1998, 72, 4104–4115. [Google Scholar] [CrossRef]
  199. Zander, K.; Sherman, M.P.; Tessmer, U.; Bruns, K.; Wray, V.; Prechtel, A.T.; Schubert, E.; Henklein, P.; Luban, J.; Neidleman, J.; et al. Cyclophilin A Interacts with HIV-1 Vpr and Is Required for Its Functional Expression. J. Biol. Chem. 2003, 278, 43202–43213. [Google Scholar] [CrossRef]
  200. Solbak, S.M.; Wray, V.; Horvli, O.; Raae, A.J.; Flydal, M.I.; Henklein, P.; Henklein, P.; Nimtz, M.; Schubert, U.; Fossen, T. The Host-Pathogen Interaction of Human Cyclophilin A and HIV-1 Vpr Requires Specific N-Terminal and Novel C-Terminal Domains. BMC Struct. Biol. 2011, 11, 49. [Google Scholar] [CrossRef]
  201. Votteler, J.; Studtrucker, N.; Sörgel, S.; Münch, J.; Rücker, E.; Kirchhoff, F.; Schick, B.; Henklein, P.; Fossen, T.; Bruns, K.; et al. Proline 35 of Human Immunodeficiency Virus Type 1 (HIV-1) Vpr Regulates the Integrity of the N-Terminal Helix and the Incorporation of Vpr into Virus Particles and Supports the Replication of R5-Tropic HIV-1 in Human Lymphoid Tissue Ex Vivo. J. Virol. 2007, 81, 9572–9576. [Google Scholar] [CrossRef]
  202. Dampier, W.; Antell, G.C.; Aiamkitsumrit, B.; Nonnemacher, M.R.; Jacobson, J.M.; Pirrone, V.; Zhong, W.; Kercher, K.; Passic, S.; Williams, J.W.; et al. Specific Amino Acids in HIV-1 Vpr Are Significantly Associated with Differences in Patient Neurocognitive Status. J. Neurovirol. 2017, 23, 113–124. [Google Scholar] [CrossRef]
  203. Begum, N.A.; Izumi, N.; Nishikori, M.; Nagaoka, H.; Shinkura, R.; Honjo, T. Requirement of Non-Canonical Activity of Uracil DNA Glycosylase for Class Switch Recombination. J. Biol. Chem. 2007, 282, 731–742. [Google Scholar] [CrossRef] [PubMed]
  204. Miller, C.M.; Akiyama, H.; Agosto, L.M.; Emery, A.; Ettinger, C.R.; Swanstrom, R.I.; Henderson, A.J.; Gummuluru, S. Virion-Associated Vpr Alleviates a Postintegration Block to HIV-1 Infection of Dendritic Cells. J. Virol. 2017, 91, e00051-17. [Google Scholar] [CrossRef] [PubMed]
  205. Fenard, D.; Houzet, L.; Bernard, E.; Tupin, A.; Brun, S.; Mougel, M.; Devaux, C.; Chazal, N.; Briant, L. Uracil DNA Glycosylase 2 Negatively Regulates HIV-1 LTR Transcription. Nucleic Acids Res. 2009, 37, 6008–6018. [Google Scholar] [CrossRef] [PubMed]
  206. Weil, A.F.; Ghosh, D.; Zhou, Y.; Seiple, L.; Moira, A.M.; Spivak, A.M.; Siliciano, R.F.; Stivers, J.T. Uracil DNA Glycosylase Initiates Degradation of HIV-1 CDNA Containing Misincorporated DUTP and Prevents Viral Integration. Proc. Natl. Acad. Sci. USA 2013, 110, E448–E457. [Google Scholar] [CrossRef] [PubMed]
  207. Mansky, L.M.; Le Rouzic, E.; Benichou, S.; Gajary, L.C. Influence of Reverse Transcriptase Variants, Drugs, and Vpr on Human Immunodeficiency Virus Type 1 Mutant Frequencies. J. Virol. 2003, 77, 2071–2080. [Google Scholar] [CrossRef] [PubMed]
  208. Schröfelbauer, B.; Yu, Q.; Zeitlin, S.G.; Landau, N.R. Human Immunodeficiency Virus Type 1 Vpr Induces the Degradation of the UNG and SMUG Uracil-DNA Glycosylases. J. Virol. 2005, 79, 10978–10987. [Google Scholar] [CrossRef]
  209. Ahn, J.; Vu, T.; Novince, Z.; Guerrero-Santoro, J.; Rapic-Otrin, V.; Gronenborn, A.M. HIV-1 Vpr Loads Uracil DNA Glycosylase-2 onto DCAF1, a Substrate Recognition Subunit of a Cullin 4A-RING E3 Ubiquitin Ligase for Proteasome-Dependent Degradation. J. Biol. Chem. 2010, 285, 37333–37341. [Google Scholar] [CrossRef]
  210. Belzile, J.P.; Duisit, G.; Rougeau, N.; Mercier, J.; Finzi, A.; Cohen, É.A. HIV-1 Vpr-Mediated G2 Arrest Involves the DDB1-CUL4AVPRBP E3 Ubiquitin Ligase. PLoS Pathog. 2007, 3, e85. [Google Scholar] [CrossRef]
  211. Chen, R.; Le Rouzic, E.; Kearney, J.A.; Mansky, L.M.; Benichou, S. Vpr-Mediated Incorporation of UNG2 into HIV-1 Particles Is Required to Modulate the Virus Mutation Rate and for Replication in Macrophages. J. Biol. Chem. 2004, 279, 28419–28425. [Google Scholar] [CrossRef]
  212. Subbramanian, R.A.; Kessous-Elbaz, A.; Lodge, R.; Forget, J.; Yao, X.-J.; Bergeron, D.; Cohen, E.A. Human Immunodeficiency Virus Type 1 Vpr Is a Positive Regulator of Viral Transcription and Infectivity in Primary Human Macrophages. J. Exp. Med. 1998, 187, 1103–1111. [Google Scholar] [CrossRef] [PubMed]
  213. Zhao, L.J.; Jian, H.; Zhu, H. HIV-1 Auxiliary Regulatory Protein Vpr Promotes Ubiquitination and Turnover of Vpr Mutants Containing the L64P Mutation. FEBS Lett. 2004, 563, 170–178. [Google Scholar] [CrossRef] [PubMed]
  214. Schröfelbauer, B.; Hakata, Y.; Landau, N.R. HIV-1 Vpr Function Is Mediated by Interaction with the Damage-Specific DNA-Binding Protein DDB1. Proc. Natl. Acad. Sci. USA 2007, 104, 4130–4135. [Google Scholar] [CrossRef] [PubMed]
  215. Sherman, M.P.; de Noronha, C.M.C.; Eckstein, L.A.; Hataye, J.; Mundt, P.; Williams, S.A.F.; Neidleman, J.A.; Goldsmith, M.A.; Greene, W.C. Nuclear Export of Vpr Is Required for Efficient Replication of Human Immunodeficiency Virus Type 1 in Tissue Macrophages. J. Virol. 2003, 77, 7582–7589. [Google Scholar] [CrossRef] [PubMed]
  216. Kino, T.; Gragerov, A.; Kopp, J.B.; Stauber, R.H.; Pavlakis, G.N.; Chrousos, G.P. The HIV-1 Virion-Associated Protein Vpr Is a Coactivator of the Human Glucocorticoid Receptor. J. Exp. Med. 1999, 189, 51–61. [Google Scholar] [CrossRef] [PubMed]
  217. Belzile, J.P.; Abrahamyan, L.G.; Gérard, F.C.A.; Rougeau, N.; Cohen, É.A. Formation of Mobile Chromatin-Associated Nuclear Foci Containing HIV-1 Vpr and VPRBP Is Critical for the Induction of G2 Cell Cycle Arrest. PLoS Pathog. 2010, 6, e1001080. [Google Scholar] [CrossRef]
  218. Zhou, X.; DeLucia, M.; Ahn, J. SLX4-SLX1 Protein-Independent down-Regulation of MUS81-EME1 Protein by HIV-1 Viral Protein R (Vpr). J. Biol. Chem. 2016, 291, 16936–16947. [Google Scholar] [CrossRef]
  219. Sarmady, M.; Dampier, W.; Tozeren, A. HIV Protein Sequence Hotspots for Crosstalk with Host Hub Proteins. PLoS ONE 2011, 6, e23293. [Google Scholar] [CrossRef]
  220. Belzile, J.-P.; Richard, J.; Rougeau, N.; Xiao, Y.; Cohen, É.A. HIV-1 Vpr Induces the K48-Linked Polyubiquitination and Proteasomal Degradation of Target Cellular Proteins To Activate ATR and Promote G2 Arrest. J. Virol. 2010, 84, 3320–3330. [Google Scholar] [CrossRef]
  221. Berger, G.; Lawrence, M.; Hué, S.; Neil, S.J.D. G2/M Cell Cycle Arrest Correlates with Primate Lentiviral Vpr Interaction with the SLX4 Complex. J. Virol. 2015, 89, 230–240. [Google Scholar] [CrossRef]
  222. Hossain, D.; Barbosa, J.A.F.; Cohen, E.A.; Tsang, W.Y. HIV-1 Vpr Hijacks EDD-DYRK2-DDB1DCAF1 to Disrupt Centrosome Homeostasis. J. Biol. Chem. 2018, 293, 9448–9460. [Google Scholar] [CrossRef]
  223. Casey Klockow, L.; Sharifi, H.J.; Wen, X.; Flagg, M.; Furuya, A.K.M.; Nekorchuk, M.; De Noronha, C.M.C. The HIV-1 Protein Vpr Targets the Endoribonuclease Dicer for Proteasomal Degradation to Boost Macrophage Infection. Virology 2013, 444, 191–202. [Google Scholar] [CrossRef]
  224. Payne, E.H.; Ramalingam, D.; Fox, D.T.; Klotman, M.E. Polyploidy and Mitotic Cell Death Are Two Distinct HIV-1 Vpr-Driven Outcomes in Renal Tubule Epithelial Cells. J. Virol. 2018, 92, e01718-17. [Google Scholar] [CrossRef] [PubMed]
  225. Romani, B.; Baygloo, N.S.; Hamidi-Fard, M.; Aghasadeghi, M.R.; Allahbakhshi, E. HIV-1 Vpr Protein Induces Proteasomal Degradation of Chromatin-Associated Class I HDACs to Overcome Latent Infection of Macrophages. J. Biol. Chem. 2016, 291, 2696–2711. [Google Scholar] [CrossRef]
  226. Romani, B.; Baygloo, N.S.; Aghasadeghi, M.R.; Allahbakhshi, E. 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. 2015, 290, 17380–17389. [Google Scholar] [CrossRef]
  227. Le Rouzic, E.; Morel, M.; Ayinde, D.; Belaïdouni, N.; Letienne, J.; Transy, C.; Margottin-Goguet, F. Assembly with the Cul4A-DDB1DCAF1 Ubiquitin Ligase Protects HIV-1 Vpr from Proteasomal Degradation. J. Biol. Chem. 2008, 283, 21686–21692. [Google Scholar] [CrossRef]
  228. Kamata, M.; Aida, Y. Two Putative-Helical Domains of Human Immunodeficiency Virus Type 1 Vpr Mediate Nuclear Localization by at Least Two Mechanisms. J. Virol. 2000, 74, 7179–7186. [Google Scholar] [CrossRef] [PubMed]
  229. Masuda, M.; Nagai, Y.; Oshima, N.; Tanaka, K.; Murakami, H.; Igarashi, H.; Okayama, H. Genetic Studies with the Fission Yeast Schizosaccharomyces Pombe Suggest Involvement of Wee1, Ppa2, and Rad24 in Induction of Cell Cycle Arrest by Human Immunodeficiency Virus Type 1 Vpr. J. Virol. 2000, 74, 2636–2646. [Google Scholar] [CrossRef] [PubMed]
  230. Hashizume, C.; Kuramitsu, M.; Zhang, X.; Kurosawa, T.; Kamata, M.; Aida, Y. Human Immunodeficiency Virus Type 1 Vpr Interacts with Spliceosomal Protein SAP145 to Mediate Cellular Pre-MRNA Splicing Inhibition. Microbes Infect. 2007, 9, 490–497. [Google Scholar] [CrossRef]
  231. Kamata, M.; Nitahara-Kasahara, Y.; Miyamoto, Y.; Yoneda, Y.; Aida, Y. Importin-α Promotes Passage through the Nuclear Pore Complex of Human Immunodeficiency Virus Type 1 Vpr. J. Virol. 2005, 79, 3557–3564. [Google Scholar] [CrossRef]
  232. Jenkins, Y.; Sanchez, P.V.; Meyer, B.E.; Malim, M.H. Nuclear Export of Human Immunodeficiency Virus Type 1 Vpr Is Not Required for Virion Packaging. J. Virol. 2001, 75, 8348–8352. [Google Scholar] [CrossRef]
  233. Wang, H.; Guo, H.; Su, J.; Rui, Y.; Zheng, W.; Gao, W.; Zhang, W.; Li, Z.; Liu, G.; Markham, R.B.; et al. Inhibition of Vpx-Mediated SAMHD1 and Vpr-Mediated Host Helicase Transcription Factor Degradation by Selective Disruption of Viral CRL4 (DCAF1) E3 Ubiquitin Ligase Assembly. J. Virol. 2017, 91, e00225-17. [Google Scholar] [CrossRef]
  234. Goh, W.C.; Manel, N.; Emerman, M. The Human Immunodeficiency Virus Vpr Protein Binds Cdc25C: Implications for G2 Arrest. Virology 2004, 318, 337–349. [Google Scholar] [CrossRef] [PubMed]
  235. Caly, L.; Saksena, N.K.; Piller, S.C.; Jans, D.A. Impaired Nuclear Import and Viral Incorporation of Vpr Derived from a HIV Long-Term Non-Progressor. Retrovirology 2008, 5, 67. [Google Scholar] [CrossRef] [PubMed]
  236. Yoshizuka, N.; Yoshizuka-Chadani, Y.; Krishnan, V.; Zeichner, S.L. Human Immunodeficiency Virus Type 1 Vpr-Dependent Cell Cycle Arrest through a Mitogen-Activated Protein Kinase Signal Transduction Pathway. J. Virol. 2005, 79, 11366–11381. [Google Scholar] [CrossRef] [PubMed]
  237. Tan, L.; Ehrlich, E.; Yu, X.-F. DDB1 and Cul4A Are Required for Human Immunodeficiency Virus Type 1 Vpr-Induced G2 Arrest. J. Virol. 2007, 81, 10822–10830. [Google Scholar] [CrossRef] [PubMed]
  238. Bolton, D.L.; Barnitz, R.A.; Sakai, K.; Lenardo, M.J. 14-3-3 Theta Binding to Cell Cycle Regulatory Factors Is Enhanced by HIV-1 Vpr. Biol. Direct 2008, 3, 17. [Google Scholar] [CrossRef] [PubMed]
  239. Romani, B.; Glashoff, R.H.; Engelbrecht, S. Functional Integrity of Naturally Occurring Mutants of HIV-1 Subtype C Vpr. Virus Res. 2010, 153, 288–298. [Google Scholar] [CrossRef]
  240. Tcherepanova, I.; Starr, A.; Lackford, B.; Adams, M.D.; Routy, J.P.; Boulassel, M.R.; Calderhead, D.; Healey, D.; Nicolette, C. The Immunosuppressive Properties of the HIV Vpr Protein Are Linked to a Single Highly Conserved Residue, R90. PLoS ONE 2009, 4, e5853. [Google Scholar] [CrossRef]
Figure 1. Scheme illustrating Vpr’s involvement in G2 arrest induction and DNA damage response engagement. Inhibitory/deleterious associations, resulting in target protein blocking/degradation, are depicted as red blunt-ended lines. Activating/beneficial associations are depicted as green arrows. Associations confirmed in CD4+ T cells or physiologically relevant models (i.e., ex vivo cultured human lymphoid tissue, patient-derived samples, etc.) are depicted as full arrows, whereas dotted arrows illustrate associations yet to be characterized therein. Pro = prophase; Prm = prometaphase; Met = metaphase; Ana = anaphase; Tel = telophase. (Figure created with BioRender).
Figure 1. Scheme illustrating Vpr’s involvement in G2 arrest induction and DNA damage response engagement. Inhibitory/deleterious associations, resulting in target protein blocking/degradation, are depicted as red blunt-ended lines. Activating/beneficial associations are depicted as green arrows. Associations confirmed in CD4+ T cells or physiologically relevant models (i.e., ex vivo cultured human lymphoid tissue, patient-derived samples, etc.) are depicted as full arrows, whereas dotted arrows illustrate associations yet to be characterized therein. Pro = prophase; Prm = prometaphase; Met = metaphase; Ana = anaphase; Tel = telophase. (Figure created with BioRender).
Viruses 16 00420 g001
Figure 2. Scheme illustrating Vpr activities. Viral processes (numbered text, arrows) and components (objects) are depicted in purple. Host cell processes (text) and components (objects) are depicted in teal. Inhibitory/deleterious associations are depicted as red objects or blunt-ended lines. Activating/beneficial associations are depicted as green arrows. Associations confirmed in primary CD4+ T cells or physiologically relevant models (i.e., ex vivo cultured human lymphoid tissue, patient-derived samples) are depicted as full arrows, while dotted arrows illustrate associations yet to be characterized therein. (Figure created with BioRender).
Figure 2. Scheme illustrating Vpr activities. Viral processes (numbered text, arrows) and components (objects) are depicted in purple. Host cell processes (text) and components (objects) are depicted in teal. Inhibitory/deleterious associations are depicted as red objects or blunt-ended lines. Activating/beneficial associations are depicted as green arrows. Associations confirmed in primary CD4+ T cells or physiologically relevant models (i.e., ex vivo cultured human lymphoid tissue, patient-derived samples) are depicted as full arrows, while dotted arrows illustrate associations yet to be characterized therein. (Figure created with BioRender).
Viruses 16 00420 g002
Table 1. Summary of Vpr mutations with an altered phenotype and/or clinical relevance. HAND = HIV-associated neurocognitive disorder; LTNP = long-term non-progressor; RP = rapid progressor; C-ter = carboxyl-terminus; N-ter = amino-terminus; α1 = alpha-helix 1; α2 = alpha-helix 2; α3 = alpha-helix 3, HLT = human lymphoid tissue.
Table 1. Summary of Vpr mutations with an altered phenotype and/or clinical relevance. HAND = HIV-associated neurocognitive disorder; LTNP = long-term non-progressor; RP = rapid progressor; C-ter = carboxyl-terminus; N-ter = amino-terminus; α1 = alpha-helix 1; α2 = alpha-helix 2; α3 = alpha-helix 3, HLT = human lymphoid tissue.
MutationDomainPhenotypeClinical
Association
Characterized InReference
Q3RN-terLower viral production
(in tandem with R77Q)
Lower cytopathic potential
LTNPHeLa
S. cerevisiae
Patient PBMCs
CD4+ T cells
[161,171,172,173]
P5AN-terReduced oligomerization
Reduced transactivation
NoneCemM7, Jurkat
HeLa
CD4+ T cells
Macrophages
[120]
P10AN-terReduced NFAT activation
Reduced transactivation
NoneCemM7, Jurkat
HeLa
CD4+ T cells
Macrophages
[120]
P14AN-terReduced transactivation
Diminished G2 arrest
NoneCemM7, Jurkat
HeLa
CD4+ T cells
Macrophages
[120]
N16HN-terNo change with respect to WTLTNPPatient PBMCs[174]
W18A/S/E/Rα1Altered virion packaging and delivery
Diminished G2 arrest
Lower cytopathic potential
NoneJurkat
S. pombe.
[167,175]
T19Aα1No change with respect to WTLTNPJurkat, HeLa
COS-7
PBMCs
[170]
E21/24Qα1Altered subcellular localization
Reduced NFAT activation
No oligomerization
No PARP1 translocation
No transactivation
NoneCemM7, Jurkat
HeLa
CD4+ T cells
Macrophages
[120]
L22A/S/Eα1Altered subcellular localization
Altered virion packaging and delivery
Diminished G2 arrest
No oligomerization (A)
Reduced PARP1 translocation (A)
Increased GRE activation (A)
Increased transactivation (A)
Nuclear aggregation (A)
NoneCemM7, Jurkat
HeLa, CV1
HEK293
CD4+ T cells
Macrophages
[120,167,176,177,178]
L23F/Aα1Altered subcellular localization
No oligomerization
Altered virion packaging and delivery
No PARP1 translocation
Diminished G2 arrest
Diminished interaction with 14-3-3η
Abolished interaction with hCGI
NoneCemM7, Jurkat
HeLa, CV1
COS-7, MT4
CD4+ T cells
Macrophages
S. cerevisiae
HEK293
[120,171,173,176,179,180,181,182,183]
E24G/Qα1Lower cytopathic potential
Impaired interaction with UNG2
Diminished G2 arrest
Altered Na+ permeability
NoneS. pombe
Hippocampal neurons
CD4+ T cells
[175,184,185,186]
E25Kα1Lower cytopathic potential
Altered virion packaging and delivery
Altered subcellular localization
Impaired interaction with UNG2
Diminished G2 arrest
Resistance to Fumagilin
NoneJurkat
S. pombe
S. cerevisiae
COS-7, MT4
HEK293, HeLa
[58,173,175,180,184,187,188,189,190]
K27Mα1Aggregation
Loss of G2 arrest
Lower cytopathic potential
Reduced oligomerization
Reduced NFAT activation
Reduced transactivation
No PARP1 translocation
Altered subcellular localization
Abolished interaction with hCG1
No induction of TNF-α
NoneCemM7, Jurkat
J-Lat, HEK293
HeLa, MT4C5
CD4+ T cells
Macrophages
[120,142,163,183]
A30L/P/F/Sα1Loss of G2 arrest
No oligomerization
Altered virion packaging and delivery
Altered subcellular localization
Unable to interact with nucleoporins
No Iκκα stimulation
NoneHEK293, COS-7
HeLa, RD
Jurkat, CV1
SupT1, MT4
S. cerevisiae
[32,110,173,180,183,190,191,192,193,194,195,196,197,198]
P35Aα1-α2
loop
Altered subcellular localization
Altered virion packaging and delivery
No oligomerization
Diminished G2 arrest
Lower cytopathic potential
Reduced NFAT activation
Reduced PARP1 translocation
Reduced transactivation
Reduced interaction with CypA
NoneCemM7, Jurkat
HeLa, HEK293
CD4+ T cells
Macrophages
Structural studies
[120,182,191,199,200,201]
R36Wα1-α2
loop
Aggregation
Altered virion packaging and delivery
Increased oligomerization
Increased viral replication
Diminished G2 arrest
RPJurkat, HeLa
HEK293, COS
PBMCs
[170,176,191]
L39Gα2Altered virion packaging and delivery
Altered subcellular localization
NoneHeLa, CV1
HEK293
[176]
G41N/A55α2High neurocognitive deficitHANDPatient PBMCs[202]
G41N/N28Sα2Inability to sequester APC1NoneCD4+ T cells
Macrophages
[71]
G41S/I37α2Low neurocognitive deficitHANDPatient PBMCs[202]
L42A/Gα2Reduced GRE activation
Altered virion packaging and delivery
Diminished G2 arrest
NoneHeLa, CV1
HEK293
[176]
H46Wα2Altered subcellular localization
(in tandem with H45W)
None CD4+ T cells[35]
Y50Aα2Diminished G2 arrestNoneHeLa, Jurkat[163]
W54Rα3Lower cytopathic potential
Impaired interaction with UNG2
Impaired interaction with SMUG
Increased viral mutation rates
Abolished interaction with hCG1
Impaired IgA class switch reccombination
NoneHEK293, HeLa
MT4C5, J-Lat
CH12F3-2, HT-29
MAGIC-5B
S. pombe
CD4+ T cells
Dendritic cells
Structural studies
[7,21,23,25,76,142,158,183,187,203,204,205,206,207,208,209,210,211]
V57Lα3Altered virion packaging and delivery
Altered subcellular localization
None MT4[198]
A59Pα3Altered virion packaging and delivery
Altered subcellular localization
Loss of G2 arrest
NoneHeLa, RD
Jurkat, CV1
[193,194,195]
R62P/N/Dα3Diminished G2 arrest
Unable to form nuclear foci
Altered subcellular localization
Diminished interaction with DCAF1
NoneS. cerevisiae
HEK293
MT4C5, J-Lat
CD4+ T cells
Structural studies
[7,142,173,177,212]
I63E/K/F/Tα3Lower cytopathic potential
Diminished G2 arrest
Reduced oligomerization
Altered virion packaging and delivery
Altered subcellular localization
(in tandem with L64R)
NoneJurkat, SupT1
COS-7, MT4
[166,171,173,180,198]
L64P/A/Sα3Unable to interact with DCAF1
Altered virion packaging and delivery
Altered subcellular localization
Loss of G2 arrest
Loss of cytopathic potential
No oligomerization
No PARP1 translocation
No NFAT activation
No GR activation
No transactivation
Increased Vpr turnover (P)
NoneCemM7, Jurkat
HeLa, CV1, RD
CD4+ T cells
Macrophages
HEK293, TZM-bl
[120,176,177,178,194,210,213,214,215,216]
64-68Aα3Altered subcellular localization
Reduced oligomerization
No PARP1 translocation
No NFAT activation
No transactivation
No virion incorporation
Loss of G2 arrest
Loss of cytopathic potential
NoneCemM7, Jurkat
HeLa
CD4+ T cells
Macrophages
[120]
Q65Rα3Unable to interact with DCAF1
Altered subcellular localization
Unable to form nuclear foci
Reduced oligomerization
Loss of G2 arrest
Loss of cytopathic potential
Diminished viral replication
Reduced induction of TNF-α
Reduced vDNA nuclear import
Diminished accumulation of γ-H2AX
Increased Vpr turnover
LTNPHEK293, MT4
CEM.SS, HeLa
MT4C5, J-Lat
CD4+ T cells
Patient PBMCs
Macrophages
Dendritic cells
Structural studies
[5,26,142,160,163,198,204,210,217,218,219,220,221,222,223,224,225,226,227]
L67E/A/S/Pα3Diminished G2 arrest
Lower cytopathic potential
Altered subcellular localization
Altered virion packaging and delivery
Reduced oligomerization
Reduced GRE activation
Abrogated splicing takeover (P)
No transactivation
NoneJurkat, HeLa
SupT1, CV1, RD
HEK293, TZM-bl
[35,166,171,176,177,179,194,195,228,229,230,231]
L68M/A/Sα3Reduced oligomerization
Lower viral production
Diminished G2 arrest
Altered subcellular localization
Reduced vDNA nuclear import (A)
RPTZM-bl
Jurkat, HeLa
COS, CV1
HEK293, RD
PBMCs
[170,176,177,192,215,232]
I70Sα3Altered subcellular localization
Reduced oligomerization
NoneJurkat, SupT1[166]
H71R/C/Y/Aα3Diminished G2 arrest
No oligomerization
Altered virion packaging and delivery
Lower cytopathic potential
Altered subcellular localization
Diminished viral replication
Impaired interaction with UNG2
Diminished interaction with DCAF1
Diminished interaction with Cdc25C
Increased NF-κB stimulation (Y)
Increased viral mutation rates
Diminished accumulation of γ-H2AX (A)
NoneHEK293, HeLa
RD, Jurkat
CV1, TZM-bl
S. pombe
CD4+ T cells
Dendritic cells
[24,25,59,79,175,186,187,192,193,194,195,204,207,209,233,234]
F72L/Aα3Altered subcellular localization
Diminished G2 arrest (A)
(in tandem with R73A)
LTNPPatient PBMCs[235,236]
R73A/Sα3Diminished interaction with hHR23A
No transactivation
NoneJurkat, HEK293
Structural studies
[58,165]
I74Eα3Diminished G2 arrest
Lower cytopathic potential
NoneJurkat, SupT1[166]
G75Aα3No oligomerization
Altered virion packaging and delivery
Altered subcellular localization
Diminished G2 arrest
NoneHEK293, HeLa
RD, Jurkat
CV1
[191,193,194,195]
C76A/Sα3Diminished interaction with DCAF1
No oligomerization
No NFAT activation
Lower cytopathic potential
Loss of G2 arrest
Reduced PARP1 translocation
Reduced transactivation
Altered virion packaging and delivery
NoneHEK293
CemM7, Jurkat
HeLa, RD
CD4+ T cells
Macrophages
[120,194,195,233,237]
R77Qα3Lower viral production
Lower cytopathic potential
Loss of G2 arrest
Reduced PARP1 translocation
No Iκκα stimulation
LTNPCemM7, Jurkat
HeLa, J-Lat
MT4C5, HEK293
S. cerevisiae
CD4+ T cells
Macrophages
Patient PBMCs
[110,120,160,164,173]
R77Aα3Lower viral production
Reduced PARP1 translocation
LTNPCemM7, Jurkat
HeLa
CD4+ T cells
Macrophages
Ex vivo HLT
[120,164]
S79A/IC-terLack of phosphorylation
Loss of G2 arrest
Increased interaction with UNG2
No transactivation
(in tandem with R80K)
No induction of TNF-α
No Iκκβ stimulation
Abolished 14-3-3θ association to targets
Altered subcellular localization
NoneJ-Lat, MT4C5, CV1
Jurkat, PBMCs
HEK293, HeLa
S. cerevisiae
Macrophages
CD4+ T cells
[58,110,142,163,168,169,173,187,193,210,238]
R80AC-terLower viral production
Loss of G2 arrest
No NFAT activation
No induction of TNF-α
No transactivation
Diminished viral replication
Diminished interaction with hHR23A
Reduction in whole-cell ubiquitination
Abolished interaction with 14-3-3η
Abolished 14-3-3θ association to targets
Unable to form nuclear foci
Unable to activate unintegrated provirus
Reduced interaction with CypA
Reduced vDNA nuclear import
Diminished accumulation of γ-H2AX
LTNPSupT1, MT4
CemM7, CEMx174
Jurkat, HeLa
J-Lat, MT4C5
HEK293, TZM-bl
S. cerevisiae
CD4+ T cells
Macrophages
Ex vivo HLT
Structural studies
[23,26,58,79,120,142,158,164,165,166,169,173,177,181,183,191,196,198,200,210,212,216,217,218,221,225,226,236,238]
R85Y/PC-terAltered subcellular localization
Lower viral production
Increased oligomerization
Diminished G2 arrest
RPJurkat, HeLa, COS
PBMCs, HEK293
[170,239]
R90NC-terNo change with respect to WTLTNPJurkat, HeLa, COS
PBMCs
[170]
R90KC-terSuppressed IL-12 production
Lower cytopathic potential
Loss of G2 arrest
Reduced transactivation
Impaired IgA class switch recombination
LTNPHEK293, HeLa
CH12F3-2
Jurkat
S. pombe
Dendritic cells
[23,59,171,175,183,187,203,211,240]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Vanegas-Torres, C.A.; Schindler, M. HIV-1 Vpr Functions in Primary CD4+ T Cells. Viruses 2024, 16, 420. https://0-doi-org.brum.beds.ac.uk/10.3390/v16030420

AMA Style

Vanegas-Torres CA, Schindler M. HIV-1 Vpr Functions in Primary CD4+ T Cells. Viruses. 2024; 16(3):420. https://0-doi-org.brum.beds.ac.uk/10.3390/v16030420

Chicago/Turabian Style

Vanegas-Torres, Carlos Alberto, and Michael Schindler. 2024. "HIV-1 Vpr Functions in Primary CD4+ T Cells" Viruses 16, no. 3: 420. https://0-doi-org.brum.beds.ac.uk/10.3390/v16030420

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

Article Metrics

Back to TopTop