Next Article in Journal
Epidemiologic and Genomic Evidence for Zoonotic Transmission of SARS-CoV-2 among People and Animals on a Michigan Mink Farm, United States, 2020
Next Article in Special Issue
HIV Expression in Infected T Cell Clones
Previous Article in Journal
Dietary Vitamin D Mitigates Coronavirus-Induced Lung Inflammation and Damage in Mice
Previous Article in Special Issue
HIV-1 Remission: Accelerating the Path to Permanent HIV-1 Silencing
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Breaking the Silence: Regulation of HIV Transcription and Latency on the Road to a Cure

by
Natasha N. Duggan
1,
Tatjana Dragic
1,
Sumit K. Chanda
1,* and
Lars Pache
2,*
1
Department of Immunology and Microbiology, Scripps Research, La Jolla, CA 92037, USA
2
NCI Designated Cancer Center, Sanford Burnham Prebys Medical Discovery Institute, La Jolla, CA 92037, USA
*
Authors to whom correspondence should be addressed.
Viruses 2023, 15(12), 2435; https://doi.org/10.3390/v15122435
Submission received: 21 November 2023 / Revised: 12 December 2023 / Accepted: 13 December 2023 / Published: 15 December 2023
(This article belongs to the Special Issue Regulation of HIV-1 Transcription and Latency)

Abstract

:
Antiretroviral therapy (ART) has brought the HIV/AIDS epidemic under control, but a curative strategy for viral eradication is still needed. The cessation of ART results in rapid viral rebound from latently infected CD4+ T cells, showing that control of viral replication alone does not fully restore immune function, nor does it eradicate viral reservoirs. With a better understanding of factors and mechanisms that promote viral latency, current approaches are primarily focused on the permanent silencing of latently infected cells (“block and lock”) or reactivating HIV-1 gene expression in latently infected cells, in combination with immune restoration strategies to eliminate HIV infected cells from the host (“shock and kill”). In this review, we provide a summary of the current, most promising approaches for HIV-1 cure strategies, including an analysis of both latency-promoting agents (LPA) and latency-reversing agents (LRA) that have shown promise in vitro, ex vivo, and in human clinical trials to reduce the HIV-1 reservoir.

1. Introduction

Human immunodeficiency virus type 1 (HIV-1) continues to pose a global health challenge. Currently, 38 million individuals are living with HIV-1, with an additional 1–2 million new infections annually [1]. Antiretroviral therapy (ART), when taken as instructed, can effectively suppress viral replication and has substantially reduced HIV-1-related mortality, transforming the infection into a manageable chronic condition. However, people living with HIV (PLWH) on long-term ART still have a higher risk of non-AIDS-related morbidities and greater mortality than adults not living with HIV [2]. Chronic inflammation and immune dysfunction persist despite ART, contributing not only to the maintenance of latent HIV but also the prevalence of clinical comorbidities [3]. In addition, ART is associated with genotoxicity and metabolic changes that place the aging PLWH population at greater risk of developing osteoporosis and fractures, renal and metabolic disorders, central nervous system disorders, cardiovascular disease, liver disease and chronic inflammation [4,5,6,7,8]. However, PLWH cannot circumvent ART because its cessation results in rapid viral rebound, showing that control of viral replication alone does not eradicate the infection [9,10,11,12]. This is due to the establishment and persistence of viral reservoirs, where HIV-1 remains transcriptionally silent and therefore impervious to the immune system and circulating antivirals [13,14]. Long-lived memory CD4+ T cells are the major reservoir of transcriptionally quiescent proviruses, but a range of other cell types, including myeloid cells, mast cells, natural killer (NK) cells, microglia, and dendritic cells (DC) have also been shown to act as HIV-1 reservoirs [15,16,17,18,19]. Thus, the field has focused on eliminating or permanently silencing viral reservoirs, coupled with immune restoration, as a curative strategy for HIV-1 infection [20,21,22].
Several approaches have emerged as frontrunners in the quest to eradicate HIV-1 infection from the host. The “shock and kill” approach aims to reactivate latent reservoirs, making them susceptible to immune responses or targeted elimination therapies. However, the low efficacy of current latency-reversing agents (LRAs) has raised questions about the definition of a cure and whether durable suppression of viral replication in the absence of ART (remission) is sufficient for long-term viral control. In contrast, “block and lock” approaches aim to use latency-promoting agents (LPAs) to “block” cells into latency and thereby “lock” viral transcription to preclude the need for ART [23]. These approaches may also affect immune function, and if locked latency cannot be achieved, patients will have to remain on ART. The effective development and implementation of any of these curative approaches requires a deep understanding of the host transcriptional machinery that HIV-1 relies on for both latency and viral gene expression. Additionally, the transcriptional machinery put in play may vary by cell type and activation status. In this review, we aim to provide an overview of the regulation of HIV-1 transcription and examine the current state of approaches to overcome the challenges posed by HIV-1 latency (Table 1).

2. HIV Transcription and Establishment of Latency

Upon HIV-1 entry into target cells, viral RNA is reverse transcribed, and the resulting DNA integrates into the host genome [24,25]. The 5′ long terminal repeat (5′LTR) drives proviral transcription and comprises a modulatory regulatory element, an enhancer, and a promoter in the U3 region, followed by the transactivation response element (TAR) in the R region [26]. This short stem-loop structure binds the viral Trans-Activator of Transcription (Tat) protein that is critical for HIV-1 gene expression [27]. Transcription activators that bind to the HIV-1 5′LTR include Sp1, NF-κB, the AP-1 complex composed of Jun and Fos protein family members, and NFAT proteins [26,28,29,30,31,32,33]. Cellular repressors or transcriptional silencers also bind the HIV-1 5′LTR, including NELF, YY1, and AP4 [34,35,36,37]. When HIV-1-infected activated CD4+ T cells transition to a long-lived resting memory state, proviral gene expression can be repressed by the absence of positive transcriptional regulators or the inhibition of their binding to the 5′LTR [26,31,38,39,40]. Additionally, several groups found that HIV-1 integration into transcriptionally active host genes promotes latency [41,42,43,44,45] by proximal promoter interference with transcription from the 5′LTR [26].
Following the binding of transcription factors, the cellular transcription machinery is recruited to the HIV promoter, including TBP (TATA-binding protein), which serves as a platform for the assembly of the RNA polymerase II pre-initiation complex [46,47,48]. The elongation of the viral transcript by RNA polymerase II is regulated by the interaction between Tat and TAR. Tat recruits the positive transcription elongation factor b (P-TEFb), which phosphorylates RNA polymerase II and other factors associated with the elongation complex, promoting processive transcription [26,49,50,51]. The autoregulation of Tat is very sensitive, and minor changes in transcription initiation rates are enough to restrict Tat production and thus halt elongation [52]. In latently infected cells, this transcriptional feedback mechanism is disrupted, resulting in a decrease in Tat below threshold levels [52,53]. There are also several negative regulatory factors that can interfere with the recruitment or function of Tat, leading to the repression of HIV-1 transcription [54,55]. For example, HEXIM1, 7SK snRNA, and DSIF (DRB sensitivity-inducing factor) sequester P-TEFb and prevent its interaction with the Tat/TAR complex [27,49,51,56].
Epigenetic mechanisms add another layer of control of HIV-1 expression and latency. In vitro studies have reported that the methylation of CpG sites found within the proviral promoter can silence the transcription of HIV-1 genes and may contribute to the maintenance of latency [26,40,57]. Other epigenetic factors such as nucleosome positioning and chromatin remodeling influence the accessibility of the TAR region and the efficiency of Tat-mediated transcriptional activation [58,59,60]. Histone modifications including deacetylation, methylation, and hypoacetylation in the vicinity of the HIV-1 promoter contribute to transcriptional repression and latency [61,62]. These epigenetic processes also influence the elongation process [59,63]. The termination of HIV-1 transcription is primarily regulated by general mechanisms such as attenuation, elongation, and, in some cases, read-through. Variability in premature termination contributes to the generation of diverse viral RNA transcripts, which oppose the formation of full-length transcripts, thus negatively impacting HIV-1 gene expression [26,57].
The establishment of HIV-1 latency is regulated by cellular processes that restrict the access of the transcription machinery to the HIV-1 promoter region [52,64]. The availability of transcriptional regulators and their interplay with epigenetic modifications of the provirus largely depend on the activation and differentiation status of infected cells and determine the outcome of HIV-1 expression and the establishment of latency. HIV latency is promoted by the activation of the phosphoinositide 3-kinase (PI3K)/Akt/mammalian target of rapamycin (mTOR) pathway in CD4+ T cells. This mechanism has been shown to promote HIV latency by suppressing viral transcription and promoting cell survival [65]. Repression of provirus expression also occurs through blocking the phosphorylation of CDK9, a P-TEFb complex member that is a cofactor for Tat-mediated transcription. Conversely, latency can be reversed through the modulation of cellular signaling pathways involved in T cell activation, such as the NF-κB and NFAT pathways [16,45,66,67].

3. “Block and Lock” Approach

The HIV-1 integration complex generally favors open, transcriptionally active chromatin, with a majority (approx. 69%) of integration sites being in active genes or in regional hotspots [41]. All current HIV-1 cure strategies adopting the block and lock approach rely on inducing epigenetic or transcriptional silencing of the HIV-1 5′LTR to prevent viral gene expression [68]. Moreover, the strategy seeks to permanently lock the viral transcriptional machinery in a suppressed state, thus precluding the need for ART [69].

3.1. RNA-Induced Silencing

Heterochromatin is transcriptionally silent, and its formation or maintenance can be induced by short interfering (si) or short hairpin (sh) RNA molecules targeting specific sequences [70]. The general mechanism is reviewed in Vansant et al. [71]. si/shRNAs, among which PromA was the first to be developed, target the distinctive tandem NF-κB sites within the HIV-1 promoter [72]. PromA siRNA effectively triggers robust transcriptional gene silencing mediated by the sustained recruitment of key factors such as Argonaute 1 (AGO1), responsible for target gene silencing, as well as histone deacetylase-1 (HDAC1) and histone methyl-transferases that induce the formation of heterochromatin [73,74]. This silencing was observed both in vitro and ex vivo in human bone marrow-derived CD34+ cells [70,75,76]. Furthermore, cells expressing PromA siRNA were found to be resistant to reactivation stimuli such as an anti-CD28 antibody and LRAs including TNF, SAHA, Bryostatin, and Chaetocin [70], though the addition of the HDAC inhibitor trichostatin-A partially restored HIV-1 transcription [77]. LTR-362 is another siRNA that targets the tandem NF-κB sites in the HIV-1 promoter that is effective in cell culture but not in HIV-1 infected humanized mice [68,78]. Differences in NF-κB sequences render clade B-targeting siRNAs less effective against clade C viruses [79], which is why S4-siRNA was developed to target the unique NF-κB binding sequences in HIV-1 subtype C, responsible for more than half of all global HIV-1 infections. There was a significant reduction in viral RNA (vRNA) levels when TZM-bl cells were transfected with S4-siRNA in vitro. siRNAs targeting other sequences in the U3 region of the 5′LTR have also been shown to induce target gene silencing but have not been pursued beyond initial in vitro studies [74]. The main hurdle to the clinical application of RNA-induced epigenetic silencing is the sustained delivery of si/shRNAs to all or most reservoir cells [80]. An additional major issue is the potential for off-target effects, such as the inadvertent targeting of genes with homology to the siRNAs [80,81]. Long non-coding RNAs (lncRNAs) are another important class of RNAs involved in transcription and gene modulation and are capable of both repressing and promoting gene expression [82,83,84]. In vitro studies in J-Lat cells have shown that HIV-1-encoded lncRNA can induce transcriptional silencing by chromatin-remodeling via the recruitment of DNMT3a, EZH2, and HDAC-1 to the virus promoter region of the 5′LTR [85]. The suppression of HIV-1 gene expression by lncRNA has been reported in multiple studies [86,87,88,89,90,91,92]. It is notable that NRON, an lncRNA expressed in resting CD4+ T cells, directly links Tat to ubiquitin/proteasome components, including CUL4B and PSMD11, thus facilitating Tat degradation [89]. Therefore, the manipulation of NRON expression in PLWH could be a novel approach for developing latency-reversing as well as latency-promoting agents.

3.2. Inhibition of Tat Function

The viral Tat protein stimulates HIV-1 RNA elongation by recruiting and activating RNA polymerase II [93]. Tat also recruits histone acetyltransferases (HATs) to the viral promoter region, leading to the activation of HIV-1 transcription [94,95]. Blocking Tat function, therefore, is a viable latency-promoting strategy. To this end, Nullbasic was developed as a trans-dominant Tat mutant with a mutated TAR-binding region. The mutant competes with endogenous, wild type Tat, thereby inhibiting HIV-1 transcription by RNA polymerase II through interaction with the positive transcription elongation factor (P-TEFb) and causing epigenetic silencing of the HIV-1 promoter [96]. Nullbasic also inhibits Rev-dependent viral mRNA transport from the nucleus by binding to DEAD/H-box helicase 1 (DDX1) [97] and inhibiting reverse transcription, leading to accelerated uncoating kinetics post infection and defective viral DNA synthesis [96,98]. In vivo studies in NSG mice with primary CD4+ T cell engraftment showed undetectable viral RNA levels only 14 days after treatment with Nullbasic [99]. However, introducing Nullbasic into all or most of the reservoir cells would face the same hurdles as any gene-therapy approach. A more viable option to modulate Tat function, therefore, could be the development of small molecule inhibitors. Cortistatins, steroid-like alkaloids isolated from the marine sponge Corticium simplex, represent such a class of compounds [100]. Didehydro-cortistatin A (dCA) inhibits TAR/Tat binding and blocks HIV-1 replication at concentrations as low as 1 nM [101]. Over time, the inhibition of Tat-dependent HIV-1 transcription by dCA results in the accumulation of epigenetic modifications in the nucleosome directly downstream of the HIV-1 promoter, restricting RNA polymerase II recruitment and elongation [101]. As such, dCA prompts the viral promoter into deep transcriptional latency, refractory to viral reactivation by cytokines, HDAC inhibitors, and protein kinase C (PKC) activators [102]. Additionally, dCA was found to enhance the recruitment of the repressive BAF complex, further contributing to the suppression of viral expression [103]. In patient-derived cell models and bone marrow/liver/thymus (BLT) humanized mouse models of HIV-1 latency, dCA effectively delays and decreases viral rebound [104] and is one of the most advanced block and lock approaches having progressed to ex vivo studies in non-human primate cells [105]. Importantly, when assessing any “block and lock” strategy, two key factors must be considered: (1) is the compound targeting and reaching all latently infected cells and (2) how long does transcriptional latency last?

4. “Shock and Kill” Strategy for Curing HIV-1 Infection

A major challenge to curing HIV is its persistence in long-lived, quiescent, CD4+ memory T cells. An effective therapy must completely remove or disable these viral reservoirs. The “shock and kill” strategy aims to reactivate latent pro-viral genomes in the presence of antiretroviral therapy and expose cells that are actively expressing viral proteins to immune clearance or treatments designed to target and kill these cells [106]. At the heart of this approach are latency-reversing agents (LRAs), small molecules or immunomodulatory treatments that trigger the expression of viral genes in latently infected cells (Figure 1). Several LRAs have been described, including T cell stimulatory agents, kinase activators, and chromatin modifiers, but these regimens have largely proven to be ineffective in clinical trials to date due to incomplete penetrance and great variability in their reactivation potential on an individual, cellular, and provirus basis [107,108,109,110,111,112,113,114,115,116,117]. This heterogeneity in reactivation is thought to arise from the complex interaction between regulatory cis-acting elements near the site of proviral integration, the extracellular environment, as well as transcriptional regulators available in the cell [118]. Additionally, many LRAs that activate proviral expression also lead to widespread immune activation or induce severe adverse effects, making them unsuitable for clinical use [119]. By understanding the complex interplay between intrinsic host factors and virus reactivation by LRAs, we can pave the way for safe and more effective and targeted interventions to eliminate HIV.

4.1. Epigenetic Modifiers

4.1.1. Long Non-Coding RNAs (lncRNAs)

As stated in the previous section, lncRNAs (long non-coding RNAs) can repress or promote gene expression. In particular, the lncRNA HEAL upregulates transcription by forming a complex with the RNA-binding protein FUS, which binds the HIV promoter and facilitates the recruitment of the histone acetyltransferase p300 [120]. This recruitment activates proviral transcription by promoting increased acetylation of histone H3K27 and P-TEFb enrichment at the HIV-1 promoter [120]. Another lncRNA, MALAT1, reverses the epigenetic silencing of HIV-1 transcription by interacting with the chromatin modulator polycomb repressive complex 2 (PRC2), disrupting its recruitment to the HIV-1 LTR promoter [121]. As a result, the methylation of histone H3 on lysine 27 (H3K27me3) through enhancer of zeste homolog 2 (EZH2), a core component of PRC2, is prevented. Methylation of H3K27me3 recruits HDACs that promote heterochromatin formation and thus latency. Consequently, preventing H3K27me3 methylation relieves the epigenetic silencing of HIV-1 transcription and promotes latency reversal [121]. However, using lncRNAs to induce latency reversal has the same caveats as the RNA-induced silencing approaches described above, namely a need for sustained delivery of the lncRNAs to the targeted reservoir cells and high target specificity to avoid the potential for off-target effects due to sequence homologies.

4.1.2. Histone Deacetylase Inhibitors (HDACi)

Histone acetylation relaxes euchromatin, rendering promoters more accessible to transcription factors and RNA polymerase II. Histone deacetylase (HDAC) removes acetyl groups from lysine residues in the NH2 terminal tails of core histones, resulting in a more closed chromatin structure and repression of gene transcription [1]. In humans, there are altogether four classes of HDACs with 18 members: class I and II can be targeted by HDAC inhibitors (HDACi), an emerging class of anticancer drugs [122,123,124,125,126,127,128,129,130,131,132,133]. As reviewed in Li et al., HDACi were originally designed to increase global transcription levels of tumor suppressor genes, thereby exerting an anti-proliferative effect [134]. Since quiescent HIV-1-infected T-cells express high levels of HDAC, HDACi were subsequently tested as LRAs and found to promote viral reactivation as a result of increased HIV-1 promoter accessibility [1,135]. HDACi vary greatly in LRA efficacy depending on the latency model they are applied to [136,137,138,139,140,141,142]. For example, while MRK-1 had modest activity both in primary cell models and JLat clones, the HDACi vorinostat acted similarly to MRK-1 in primary models but showed poor activity in JLat cells [142]. Nonselective pan-HDACi, such as vorinostat and panobinostat, which inhibit many HDAC class isoforms, have a greater potential for toxicity. More selective HDACi that specifically target class I HDACs, like entinostat and romidepsin, may be able to act as LRAs with reduced off-target effects, provided they can exhibit the same potency as the pan-HDACi [143]. In the last decade, four different HDACi were approved by the US-FDA for the treatment of several cancers, and three of these, romidepsin, panobinostat, and vorinostat, have been tested as LRAs in clinical trials (NCT02850016, NCT01680094, NCT01365065, NCT02513901, NCT00289952, NCT01933594, NCT01319383). Panobinostat was well tolerated and caused a 3.5-fold increase in cell-associated HIV-1 RNA in aviremic PLWH on ART but did not decrease reservoir size [113]. Romidepsin showed good latency reversal activity in an in vitro T cell model with an EC50 of 4.5 nM [144], but outcomes in clinical trials were inconsistent and ranged from no effect to a moderate increase in HIV-1 transcription [112,115]. Vorinostat disrupted HIV-1 latency in individuals on ART, leading to a median 4.6-fold increase in cell-associated unspliced HIV-1 RNA in resting memory CD4+ T cells, but it did not result in lower HIV-1 RNA levels in study participants upon analytical treatment interruption (ATI) [107,111]. Moreover, combinations of Vorinostat and Romidepsin with broadly neutralizing antibodies (VRC07-523LS and 3BNC117, respectively) showed no impact on reservoir size in a clinical trial [108,109]. Additional HDACi have shown promising LRA efficacy in vitro/ex vivo but have not yet progressed to clinical trials as LRAs, including Givinosat [145], currently in phase III clinical trials for the treatment of Duchenne Muscular Dystrophy [135]; Belinostat [136], FDA-approved for peripheral T-cell lymphoma [146,147]; and Entinostat, which is being tested in clinical trials for breast cancer [135]. Mocetinostat [135], developed to treat several blood cancers [148], activated latent HIV-1 expression ex vivo in patient cells but did not proceed to clinical trials due to cytotoxicity [146]. An in vitro study of oxamflatin suggested that a small therapeutic window may limit proviral expression at sub-cytotoxic concentrations [149]. Notably, the newly developed HDACi CC-4a has been reported to reactivate latent HIV-1 as well as induce apoptosis in infected cells in vitro [150], suggesting the potential to be both a “shock” and “kill” agent. Overall, despite promising in vitro and ex vivo data, HDACi have performed poorly as single agents in vivo. However, future studies may investigate the potential benefit of incorporating HDACi into an LRA cocktail to leverage synergies by targeting multiple latency mechanisms.

4.1.3. Histone Methyltransferases (HMT) Inhibitors (HMTi)

Histone methylation marks, such as H3K9me3 (mediated by Suv39H1) and H3K27me3 (mediated by G9-alpha), create a chromatin environment that restricts transcriptional activity [151]. These repressive marks are often found in the HIV-1 promoter, contributing to the establishment and maintenance of latency [152]. HMTis disrupt the deposition of repressive methylation marks, leading to a more relaxed chromatin structure, and enhanced accessibility of the viral promoter to the transcriptional machinery [153,154]. Given the potential for contrasting effects arising from different histone methylation patterns, achieving latency reversal through HTMi treatment will demand a high specificity for distinct methyltransferases to precisely target the intended epigenetic mark and avoid off-target effects [155]. Ex vivo experiments in resting CD4+ T cells isolated from ART-suppressed PLWH found that chaetocin, an inhibitor of Suv39H1, as well as DZNep and BIX-01294, inhibitors of G9-alpha, significantly increased virus production [156]. However, another study failed to reproduce high levels of latency reversal by these agents [152]. Further investigations found that DZNep was cytotoxic at levels required for latency reversal. However, lower doses in combination with other LRAs such as HDACi may be an option to achieve a therapeutic window. Due to their cytotoxicity, the progress of these drugs to clinical trials as LRAs has been halted pending results from ongoing oncology clinical trials [157].

4.1.4. Bromodomain and Extra-Terminal Domain Inhibitors (BETi)

BET proteins are epigenetic readers that play a vital role in gene expression [158,159,160]. Each of the four protein family members, BRD2, BRD3, BRD4, and BRDT, comprises two N-terminal bromodomains (BD1 and BD2) that bind acetylated lysine residues on histones [158,160]. BRD4 relaxes chromatin [161] and recruits transcription factors, the Mediator complex, and RNA Pol II to active promoter and enhancer regions [162,163,164]. BRD4 also binds P-TEFb, which activates transcription initiation and elongation [165]. During active HIV-1 transcription, BRD4 and Tat compete for binding to P-TEFb, which is available in limited supply [166,167]. In latency, high levels of BRD4 outcompete Tat for P-TEFb binding and inhibit Tat-mediated transcriptional transactivation while increasing basal transcription [168,169]. The roles of the other BET proteins in HIV-1 expression are less well defined. However, BRD2 was shown to promote latency and suppress HIV-1 transcription in a Tat-independent manner [170] by recruiting repressor complexes to the LTR [171]. In in vitro shRNA BRD2 knockdown experiments, HIV-1 reactivation was comparable to treatment with BET inhibitor JQ1 [170].
BET inhibitors (BETi) are small molecules that displace BRDs from chromatin by binding to their bromodomains BD1 and BD2 [172]. This enables binding of Tat to P-TEFb, thereby activating HIV-1 transcription from the LTR [153,173,174]. Like many other categories of LRAs, BETi were originally designed as cancer therapeutics [175,176,177,178,179] that were subsequently shown to reverse HIV-1 latency [180]. Several BETi have shown varying degrees of latency reversal in vitro and ex vivo [168,169,180,181,182,183,184,185,186,187]. Well established pan-BETi that universally target BET proteins, such as JQ1 and I-BET151, have shown strong latency reversal activity in vitro, but these results have not translated to ex vivo and in vivo studies at non-toxic concentrations [168,170,180,182,183]. For example, JQ1 potently reversed HIV-1 latency in multiple cell lines but had only modest activity when tested in resting CD4+ T cells from PLWH [168,169,180,181,182]. I-BET151 preferentially reactivated HIV-1 in monocytic cells over T cells when tested in humanized mice, with no p24 detected in CD4+ T cells [183]. OTX015 induced a 2-fold increase in HIV-transcription in resting CD4+ T cells from PLWH [184]. More recently, BRD4-selective BETi have been developed that preferentially target BRD4 over other BET family proteins. CPI-203 targets the BD1 bromodomain of BRD4 and was found to be more potent as LRA than JQ1 in a J-Lat model [183]. Studies in resting CD4+ T cells from PLWH indicated that CPI-203 may offer a significant therapeutic window because cytotoxicity was observed only at concentrations more than 100 times greater than its effective concentration of 1 uM [185]. MMQO, a BRD4-specific functional mimic of JQ1, and RVX-208, a BRD4(BD2)-selective BETi, were ten-fold less potent than JQ1 in CD4+ T cells from PLWH but are expected to have a greater therapeutic index [186,187]. In vivo mouse studies suggest that BD2-selective BETi are less toxic and do not induce widespread immune activation or a cytokine storm in comparison to pan-BETi or BD1-selective BETi [188,189,190], but they are also less potent. Because effective dose cytotoxicity has been a limiting factor in the successful in vivo LRA activity of BETi, this new generation of BD2-selective BETi may offer a therapeutic window that allows achieving latency reversal in vivo.

4.2. Activators/Inhibitors of Inducible Host Factors

4.2.1. Toll-Like Receptor (TLR) Agonists

TLRs are pattern recognition receptors expressed on the cell surface or within the endosomal structures of various innate and adaptive immune cells, including B cells, macrophages, dendritic cells (DCs), and T cells [191,192,193,194]. Upon recognizing specific pathogen-associated molecular patterns (PAMPs), TLRs trigger signaling pathways that result in the expression of innate antiviral factors as well as the interferon response [192]. The stimulation of TLRs results in the upregulation of proinflammatory cytokines and immune modulators through the NF-κB and MAPK pathways, which are also involved in HIV-1 expression [195]. Following observations that microbial PAMPs induce HIV-1 transcription, TLR agonists have been studied as potential LRAs [195,196]. The latency-reversing activity of TLR agonists can be direct, by activating signaling pathways in CD4+ T cells, or indirect, by activating other immune cells to release cytokines or IFNs that in turn mediate latency reversal in infected CD4+ T cells [197,198].
A range of agonists targeting different TLRs have been investigated in vitro and ex vivo (reviewed in [195,196]). Compounds that activate TLR3, TLR7, and TLR9 are the most advanced candidates that have progressed to clinical trials in ART-suppressed study participants. In a clinical study of the synthetic double-stranded RNA molecule poly-ICLC, a TLR-3 agonist, transient innate immune stimulation was reported. However, no signs of HIV-1 latency reversal or changes in the size of the viral reservoir were observed [199]. The TLR9 agonist MGN1703 induced detectable viral RNA in plasma of 6 out of 15 study participants in a first clinical trial treating ART-suppressed PLWH but caused no reduction in reservoir size [116]. A second trial of MGN1703 with prolonged treatment duration also did not reduce the size of the viral reservoir or affect viral rebound upon ATI [117].
The TLR7 agonists GS-986 and GS-9620 (Vesatolimod) have been studied in SHIV- and SIV-infected rhesus macaques before advancing to clinical trials. In the first NHP study, treatment of SIV-infected rhesus macaques with a combination of GS-986 and the therapeutic vaccine Ad26/MVA resulted in a reduction in viral DNA levels and delayed viral rebound upon ATI [200]. A second study by the same authors that combined GS-9620 with the broadly neutralizing antibody (bnAb) PGT121 to treat SHIV-infected rhesus macaques also reported delayed rebound in a majority of animals [201]. Further, Lin et al. treated SIV-infected rhesus macaques with repeated doses of GS-986 or GS-9620, reporting detectable RNA in all treated animals in the first phase of the study and a diminished response in a second intervention period [202]. A reduction in the inducible SIV reservoir was reported as well. However, more recently, other groups failed to detect spikes in plasma viral RNA levels or differences in rebound kinetics following ATI in GS-9620-treated animals [203,204]. It has been proposed that differences in the duration and timing of ART initiation may be responsible for the different outcomes of the studies, suggesting that the induction of viral transcription by GS-9620 may be highly sensitive to the characteristics of the latent reservoir [203]. In a phase I clinical trial, GS-9620 was well tolerated, safe, and reversed latency in HIV-1-infected individuals, though the size of the latent viral reservoir remained unaffected (NCT02858401) [114]. A second phase I trial of GS-9620 is ongoing (NCT03060447). Thus, while some studies conducted in NHP showed promise for TLR7 agonists as LRAs, it remains to be seen whether these results can be reproduced in a clinical setting. Pending results from the clinical trial NCT03060447 promise to provide additional insights.

4.2.2. Activators of Canonical NF-κB Signaling

Upon infection, HIV-1 activates NF-κB through various signaling pathways, leading to its translocation into the nucleus. The binding of NF-κB to specific enhancer elements in the viral LTR then promotes the transcription of viral genes [205]. When HIV-1 is in a latent state, NF-κB binding to the 5′LTR can reactivate viral gene expression [206,207]. Thus, regulating NF-κB signaling has been investigated as a latency reversal strategy. Protein kinase C (PKC) agonists, which activate the canonical NF-κB pathway, are among the most effective LRAs in vitro described to date. PKC is a family of serine/threonine kinases that are activated by diacylglycerol (DAG), a phospholipases C (PLC) metabolite. PKCs phosphorylate their cellular substrates, including IκB, which is essential for the activation of the canonical NF-κB pathway [142,208]. Natural and synthetic PKCa, including phorbol ester phorbol myristate acetate (PMA), Prostatin, dipeptidyl peptidase (DPP), Bryostatins, diacylglycerol (DAG) analogs, and Ingenol derivatives, all activate the NF-κB pathway to induce HIV-1 transcription [209]. Although PMA is an effective LRA, a clinical trial for its use as a cancer therapeutic highlighted severe adverse effects, thereby excluding it from clinical use [207]. Similarly, Prostratin potently reactivated HIV-1 production in latently infected rCD4+ T cells from PLWH, but caused substantial cytotoxicity, equivalent to 0.04 uM of PMA at >10 uM [210]. Additionally, the activation of the canonical NF-κB pathway by these compounds was shown to induce widespread immune activation [211,212]. Bryostatin-1, the most well studied member of the Bryostatins, a class of macrocyclic lactones, was found to be a significantly more potent LRA, activating HIV-1 expression in THP-p89 cells at low nanomolar concentrations [213,214]. Unlike compounds that reactivate HIV-1 by targeting PKC-α, β1, β2, or γ, Bryostatin-1 activates both PKC-α and PKC δ to reverse latency in vitro [215,216]. The activation of PKC δ enhances HIV replication in the presence of sub-optimal concentrations of Tat by mediating phosphorylation of Nef [213,217,218]. Additionally, unlike other isoforms, PKC δ does not require calcium for activation [215]. An unexpected effect of Bryostatin-1 is that treated cells are more resistant to apoptosis, as shown by the ERK1/2-dependent phosphorylation of anti-apoptotic BCL2 [219,220]. This could potentially interfere with the activity of additional kill agents used in combination to deplete reservoir cells following viral reactivation with this compound. While one clinical trial reported a single dose of Bryostatin-1 to be well tolerated in aviremic HIV-1-infected individuals on ART, the plasma concentrations of the drug achieved in this study were insufficient to activate PKC and significantly below the levels required for HIV-1 latency reversal [110]. Other clinical trials evaluating Bryostatin-1 for cancer indications at higher doses reported severe adverse effects, indicating that this compound may not be tolerable at concentrations effective for latency reversal [221].
Ingenol, another PKCa, is extracted from the plant Euphorbia peplus and has long been used in traditional medicine to treat skin conditions and certain cancers. More recently, natural and man-made derivatives of Ingenol have been used as treatments for skin cancers [222] and were found to act through the PKC/NF-κB pathway [223]. With regard to HIV-1, novel semi-synthetic Ingenols have been developed that exhibit optimized LRA activity [224,225]. Among these compounds, Ingenol-3-mebutate (Ingenol-3-angelate) and Ingenol B have demonstrated latency-reversing activity in multiple systems, including a study treating SIV-infected rhesus macaques with a combination of Ingenol B and Vorinostat that resulted in one of two animals exhibiting increased viral loads, albeit in the presence of systemic inflammation markers [226,227,228]. A newly designed stabilized Ingenol B derivative, GSK445A, induced latency reversal in CD4+ T cells from both ART-suppressed humans and rhesus macaques at concentrations above 10 nM. When tested in vivo in ART-suppressed SIV infected macaques, the drug was well-tolerated by most animals, and three out of four monkeys showed a modest but measurable increase in plasma SIV RNA after three doses [229]. Lastly, the PKC activator Gnidimacrin, a daphnane dieterpene that is a potent anti-cancer agent was also found to reverse latency in chronically infected cell lines, ACH-2 and U1, at picomolar concentration [230,231,232]. In a viral outgrowth assay treating PBMCs from PLWH, a reduced frequency of latently infected cells was reported, which the authors attributed to the induction of a strong CTL response by Gnidimacrin [233]. Overall, PKCa, in particular Bryostatins and Ingenols, have demonstrated potent LRA activity but are also linked to systemic T cell activation, cytotoxicity, and adverse effects that result in a small therapeutic window and limit their clinical use. In the absence of novel molecules with a reduced risk of adverse effects, PKCa could potentially be utilized at sub-toxic doses in synergistic combinations with other LRAs. In addition to PKCa, disulfiram [bis(diethylthiocarbamoyl) disulfide], an FDA-approved drug for the treatment of alcoholism, has been identified as an LRA that promotes HIV-1 transcription via canonical NF-κB signaling [234]. Disulfiram mediates the depletion of the phosphatase and tensin homolog (PTEN) protein, which activates the PI3K/Akt pathway and results in the activation of NF-κB transcription factors [235]. Two clinical trials conducted to evaluate latency reversal by disulfiram, including a phase 2 dose escalation study, showed that the drug was well tolerated but induced only modest increases in HIV-1 transcription, with a two-fold increase in cell associated unspliced HIV-1 RNA and no reduction in the reservoir size [236,237]. A study assessing combinations of disulfiram with PKCa or HDACi did not find evidence of synergy between the LRAs ex vivo [238].

4.2.3. Second Mitochondria-Derived Activator of Caspases (Smac) Mimetics

In addition to activation through PKCa-mediated canonical NF-κB signaling, HIV transcription can also be induced by the non-canonical NF-κB pathway (ncNF-κB) [239]. Unlike the canonical pathway, which is characterized by the rapid onset of broad and transient activation of genes, ncNF-κB signaling induces the activation of a more limited set of genes with slower, longer-lasting kinetics [240]. The slower onset, more persistent activity, and higher functional selectivity of ncNF-κB have been found to induce latency reversal while limiting toxicity [239,241,242]. The ncNF-κB pathway is activated through a subset of tumor necrosis factor receptors (TNFRs), including lymphotoxin beta receptor (LTbR) and CD40 [240]. In the absence of receptor ligation, TRAF3, in complex with TRAF2, cIAP1, and cIAP2, constitutively degrades NF-κB-inducing kinase (NIK) and prevents ncNF-κB pathway activation. Upon receptor stimulation, TRAF3 is degraded, leading to NIK accumulation, IKKa activation, and p100 cleavage to p52, which translocates into the nucleus along with RELB [240]. In addition to receptor stimulation, the ncNF-κB pathway can also be activated through the degradation of cIAP1 mediated by the second mitochondria-derived activator of caspases (Smac) protein that targets the inhibitor of apoptosis protein (IAP) family [243,244]. Smac mimetics are small molecules mimicking a sequence of Smac and were originally designed as cancer therapeutics to compete with XIAP for caspase binding, thereby promoting apoptosis. Smac mimetics can also bind to cIAP1 and cIAP2 and allosterically activate their E3 ubiquitin ligase activity, leading to their autoubiquitination and subsequent proteasomal degradation.
Our group previously identified cIAP1 as a negative regulator of HIV-1 transcription due to its inhibition of non-canonical NF-κB signaling [239]. We were able to demonstrate that Smac mimetics act as LRAs by mediating the ubiquitination and degradation of cIAP1, which leads to the binding of non-canonical NF-κB to the HIV-1 5′LTR [239]. Several Smac mimetics, including LCL161, Debio-1143 (Xevinapant), Ciapavir, and AZD5582, have since been shown to exhibit LRA activity in vitro, ex vivo, and in vivo [239,241,242,245]. Studies comparing the LRA activity of different compounds have found that Smac mimetics with a bivalent structure, having the ability to bind two domains of IAP proteins in cis or in trans, appear to exhibit superior latency reversal activity compared to monovalent molecules [239,241]. Importantly, these studies have demonstrated the LRA activity of the Smac mimetic Ciapavir in a humanized BLT mouse model and of AZD5582 in both a humanized mouse model and SIV-infected rhesus macaques, showing increases in viral RNA levels in treated, ART-suppressed animals in the absence of widespread immune activation. Moreover, a recent study combined the Smac mimetic AZD5582 with SIV Env-specific Rhesus monoclonal antibodies (RhmAbs) ± N-803 (an IL-15 superagonist) to treat SIV-infected, ART-suppressed adolescent rhesus macaques [246]. Beyond demonstrating latency reversal in most Smac mimetic-treated animals, the authors reported a reduction in the lymph node viral reservoir, evidenced by lower levels of total and replication-competent SIV-DNA in lymph node-derived CD4+ T cells in animals treated with a combination of Smac mimetic and RhmAbs. While it is unlikely that AZD5582 will progress into the clinic due to potential toxicity issues associated with this particular molecule, the data demonstrate clear proof of concept that Smac mimetics, as part of a shock and kill approach, can be employed to reduce reservoir size. Thus, Smac mimetics currently represent one of the most promising classes of LRAs, able to reverse HIV-1 latency in vivo in the absence of systemic immune activation and with minimal adverse effects and showing potential for reservoir depletion when combined with an antibody treatment. The development of novel Smac mimetics that enable latency reversal without inducing immune activation will be an important step towards the use of these compounds as a cure treatment in the clinic.

4.3. Latency-Reversing Agents in Combination

While different LRAs have been shown to reactivate HIV-1 expression in various cellular and animal models, none have succeeded at effectively reducing, let alone clearing, the latent viral reservoir to date. The success of any cure strategy will depend on reaching all latently infected cells, and combination treatments to target multiple distinct mechanisms may be required to broadly and effectively reverse latency [238]. Because most LRAs were originally designed to induce cancer cell death, many exhibit some level of cytotoxicity at the concentrations required for robust latency reversal as single agents. Since current ART regimens allow for the management of HIV-1 with a high quality of life and life expectancies comparable to people without HIV, adverse effects of any curative treatment for HIV-1 need to be minimal [247]. Combining such LRAs to leverage synergistic effects by targeting multiple mechanisms would allow the use of sub-toxic concentrations of the drugs. Several studies have investigated combining different classes of LRAs. PKCas, particularly Bryostatin, Prostatin, and Ingenol, were found to synergize with HDACi in several cell lines and ex vivo in resting CD4+ T cells from PLWH [135,229,238,248,249,250]. The combination of these LRAs allowed up to a 10-fold increase in the efficacy of latency reversal at significantly lower concentrations of the individual agents, thereby increasing their therapeutic window. The HDACi CC-4a, which was shown to reactivate HIV and induce apoptosis, synergized with the PKCa Prostratin without triggering widespread immune activation [150]. PKCas have also been found to synergize with BET inhibitors, particularly JQ1, in vitro and ex vivo [206,227,238,251,252,253], but result in widespread immune activation, and varying levels of cytotoxicity at levels required for substantial latency reversal [170,185,186,187,254,255]. It is critical for any combination treatments to maintain a balance between viral reactivation and immune activation that ensures LRA efficacy while avoiding immune-related adverse effects. Interestingly, the immune activation typically caused by PKCas decreased when newly designed PKCa 10-methyl-apog-1 (10MA-1) was combined with the BETi JQ1 [256]. Smac mimetics have been shown to synergize with HDACi and BETi as well but did not show significant synergy in combination with PKCas [239,241,242,245,257]. While combinations of Smac mimetics and BET inhibitors showed excellent latency reversal activity in vitro, the efficacy of these combinations could not be replicated ex vivo [242,245,253,258,259]. Smac mimetics also exhibited synergy in combination with HDACi in vitro and ex vivo. However, these treatment combinations have not yet been evaluated in vivo [241,242,245,260].
A recent study combining a Smac mimetic with the IL-15 superagonist N-803 and RhmAbs reported latency reversal and a reduction in reservoir size in SIV-infected rhesus macaques upon treatment [246]. While the impact of N-803 on latency reversal in this study was modest, the use of IL-15 or N-803 as a part of cure strategies has been studied extensively (reviewed in [261]). In vivo studies and clinical trials have indicated that IL-15 or N-803 alone are unlikely to provide sufficient latency reversal activity but have highlighted their potential as a component of combination treatments. This is supported by observations of immune restorative effects of IL-15 or N-803 treatment that may support the clearance of reservoir cells [261,262,263].
The small molecule 3-Hydroxy-1,2,3-benzotriazin-4(3H)-one (HODHBt) has been shown to act as an LRA ex vivo in cells from PLWH by enhancing STAT5 activation and its binding to the HIV LTR, promoting viral transcription [264]. Further, in combination with IL-15, STAT5 activation by HODHBt mediated increased latency reversal and enhanced the immune effector functions of NK cells and CD8 T cells targeting HIV-1-infected cells, leading to a reduction in intact proviruses ex vivo [263,265,266].
A new LRA mechanism affecting transcriptional regulation has been proposed by a recent study reporting that iPAF1C, an inhibitor of the polymerase-associated factor 1 complex (PAF1C), reduces the genome-wide chromatin occupancy of PAF1C and thereby induces the release of promoter-proximal paused RNA Pol II [267]. The molecule reactivated latent proviruses ex vivo in cells from PLWH and was shown to enhance viral reactivation by several LRAs, including the BETi JQ1. iPAF1C has been proposed to remove a block to transcriptional elongation, thereby promoting synergistic latency reversal when combined with an LRA that stimulates transcription initiation [267].
The synergistic effects of LRA combination therapies on the latent reservoir show great promise but will require careful assessment of their effects on the immune response. Multiple studies have implicated certain LRAs, in particular HDACi and PKC, to impact CD8+ T cell function, indicating that these effects will have to be further examined to ensure and maintain an effective CD8+ T cell response during treatment [268]. In general, combining LRA treatments will require a thorough evaluation of potential drug–drug interactions that may result in unanticipated adverse effects not observed in individual drug treatments to ensure the safety of these regimens.

4.4. Kill Agents

HIV-1 infection kills its target cells and causes immune dysregulation, which further complicates elimination of the virus, even in the presence of an effective latency reversal agent. Numerous studies have shown that latency reversal alone is insufficient to eliminate reservoir cells, indicating that an additional immune effector component must be coupled with LRA treatment [269]. To this end, a wide range of immune-based therapies are being investigated, including therapeutic vaccines, antibodies, and the enhancement of T cell function (reviewed in [270]). Stem-cell transplants from a CCR5-negative donor have been successful in eliminating intact HIV-1 in a small number of patients, including Timothy Ray Brown, known as the ‘Berlin Patient’ and Adam Castillejo, known as the ‘London Patient’ [271,272]. While the significant risks and complications associated with stem-cell transplantation make it unsuitable for broad clinical use as an HIV-1 cure, gene editing of T cells to disrupt CCR5 coreceptor expression represents an alternative to protect CD4+ T cells from infection [273]. Clinical trials with CCR5 gene-edited CD4+ T cells in PLWH have reported increased CD4+ T cell counts [274] and delayed, though not prevented, viral rebound [275]. The development of a CCR5 gene-edited memory stem cell-like CD4+ T cell subset has been proposed to enable the long-term sustained reconstitution of CD4+ T cells and the decay of the viral reservoir [276]. However, the potential effects of CCR5 editing on immune responses to other pathogens, as well as the risk posed by CXCR4-tropic HIV-1 strains, must be taken into consideration when evaluating these therapeutic strategies [277].
Chimeric antigen receptor (CAR)-T cells that recognize and eliminate infected cells are effectors that have been studied extensively as “kill agents” in the context of HIV-1 latency. Originally designed as a cancer therapeutic [278], this technology was applied to HIV-1 elimination strategies in the 1990s by modifying cytolytic CD8+ T cells to express CD4 with an MHC-independent intracellular signal transduction domain. This allowed for CD8+ T cell-mediated cytolysis despite HIV-1-dependent downregulation of MHC-1 [279]. Despite in vitro studies showing that CAR-T cells specifically targeted and lysed gp120-expressing cells, no control of viral infection was observed in clinical trials [279]. Costimulatory domains were added (including CD28 and IL-12) to increase lymphocyte activation and attract other innate immune cells [280,281]. Clinical trials of these constructs failed to show efficacy even as they exhibited strong safety profiles [282]. The most recent generation of CAR-Ts has a dual CAR-T construct designed to target two gp120 epitopes, preventing viral interaction with CCR5. Clinical trials of this construct are ongoing [283]. Several obstacles have been encountered in the design of CAR-T therapy, including their infection by HIV-1, viral escape, and a need to broadly target the viral reservoir. These hurdles must be overcome for CAR-T therapy to be effective, as well as combining this therapy with an effective “shock” approach [284].
Beyond cell and gene therapy, further approaches are being investigated to eliminate reservoir cells. Antibody-derived bispecific molecules that recognize HIV-1-infected cells based on monoclonal envelope-targeted antibodies have been engineered to recruit cytotoxic T cells or NK cells that mediate antibody-dependent cellular cytotoxicity (ADCC) (reviewed in [285]). Immune checkpoint molecules, including those targeting programmed cell death-1 (PD-1), are known to contribute to the establishment and maintenance of HIV-1 latency [286]. Studies investigating whether immune checkpoint inhibitors may act to disrupt the viral reservoir not only showed evidence of latency reversal in a clinical trial [287], but PD-1 blockers have also been indicated to enhance the immune clearance of reservoir cells, which may be due to a proliferation and activation of HIV-specific CD8+ T cells [288,289,290,291]. TGF-β signaling has been found to enhance the establishment and maintenance of HIV-1 latency [292]. Consequently, TGF-β inhibition was proposed as a strategy to target the latent reservoir. This has been demonstrated by recent studies that reported increased latency reversal in SIV-infected rhesus macaques by the TGF-β inhibitor galunisertib [293] and a decrease in the viral reservoir size, evidenced by reduced cell-associated viral DNA levels and likely resulting from the stimulation of SIV-specific immune responses [294].
Additional kill strategies have been proposed that induce apoptosis in latently infected cells following LRA treatment by targeting specific signaling pathways. An example is the inhibition of the PI3K/Akt pathway, which is critical to controlling the cell cycle and promoting cell survival. PI3K antagonists could induce apoptosis in HIV-1-infected cells as part of a shock and kill treatment [66,295,296]. Also being investigated as inducers of apoptosis in the context of latency are Bcl-2 antagonists such as venetoclax, FDA-approved for treatment of myeloid and lymphocytic leukemia, or navitoclax [297,298]. Further, the inhibition of Jak-STAT signaling through the treatment of PLWH on ART with the Jak1/2 inhibitor ruxolitinib in a clinical trial resulted in a significant decrease in Bcl-2 expression [299], which has been associated with reservoir reduction following latency reversal [300]. A decrease in HIV-1-DNA-harboring T cells upon ex vivo treatment with Jak inhibitors ruxolitinib and tofacitinib has been reported [301]. Together, these data suggest that a combination of Jak inhibitors with LRAs could promote reservoir depletion [299]. Lastly, RIG-I inducers and Smac mimetics have been proposed to exhibit both LRA and pro-apoptotic activities that may both reactivate latent HIV-1 and induce cell death, though further studies are needed to confirm these activities [302,303,304].
An important caveat for the development of kill agents is that these approaches will require an effective LRA treatment to allow for the assessment of their efficacy to eliminate reactivated reservoir cells. Therefore, and due to the combined effects that certain treatments have on both the latent reservoir and the immune system, the development of “shock” and “kill” treatments is closely connected and mutually dependent.
Lastly, the disruption of integrated proviral DNA through CRISPR/Cas gene editing has emerged as an alternative approach to eliminate the viral reservoir (reviewed in [305]). While studies investigating gene editing approaches as an HIV-1 cure strategy have shown promise, the need for effective delivery methods and the risk of off-target effects, viral escape, and immunogenicity continue to pose significant challenges for this strategy.

5. Conclusions

With HIV/AIDS continuing to impose an immense global burden, the urgency for an effective cure is paramount. Overcoming the hurdle of long-lived latently infected CD4+ T cells is necessary to eliminate the virus in PLWH or achieve viral remission in the absence of ART. HIV cure therapy is particularly crucial for populations with limited access to healthcare as it offers a potential path to alleviating the burdens imposed by viral persistence. In environments where healthcare resources are scarce, the long-term management of HIV through traditional antiretroviral therapy can present significant challenges, including accessibility, adherence, and cost. A cure, which eliminates the need for lifelong treatment, would not only improve individual health outcomes but also reduce the strain on healthcare systems. However, as HIV antiretroviral treatment has transformed the landscape of managing the virus, providing individuals on ART with a high quality of life and normal life expectancy, the pursuit of curative therapies for HIV must adhere to exceptionally high levels of safety. Therapeutic intervention aiming at reservoir elimination and a complete cure must not compromise the safety and efficacy of existing ART regimens. In the past decade, several approaches, including “Shock and Kill” and “Block and Lock”, have advanced to the pre-clinical or clinical trial stage but have not yet translated into a successful cure. Nevertheless, recent studies have shown encouraging progress. However, the further development of these approaches, and likely the combination of multiple treatments, will be necessary to successfully eliminate the latent HIV-1 reservoir from PLWH and achieve a complete cure.

Author Contributions

Conceptualization, N.N.D., T.D. and L.P.; writing—original draft preparation, N.N.D. and L.P.; writing—review and editing, N.N.D., T.D., S.K.C. and L.P.; visualization, L.P.; supervision, S.K.C. and L.P.; funding acquisition, S.K.C. and L.P. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Institute of Allergy and Infectious Diseases (NIAID) of the National Institutes of Health (NIH/NIAID). N.N.D. is supported by training grant T32 AI007344, and T.D., S.K.C., and L.P. are supported by grant UM1 AI164561 (Reversing Immune Dysfunction for HIV-1 Eradication (RID-HIV), part of the Martin Delaney Collaboratories for HIV Cure Research (MDC)).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Archin, N.M.; Margolis, D.M. Emerging strategies to deplete the HIV reservoir. Curr. Opin. Infect. Dis. 2014, 27, 29–35. [Google Scholar] [CrossRef] [PubMed]
  2. Zicari, S.; Sessa, L.; Cotugno, N.; Ruggiero, A.; Morrocchi, E.; Concato, C.; Rocca, S.; Zangari, P.; Manno, E.C.; Palma, P. Immune Activation, Inflammation, and Non-AIDS Co-Morbidities in HIV-Infected Patients under Long-Term ART. Viruses 2019, 11, 200. [Google Scholar] [CrossRef]
  3. Cai, C.W.; Sereti, I. Residual immune dysfunction under antiretroviral therapy. Semin. Immunol. 2021, 51, 101471. [Google Scholar] [CrossRef]
  4. Alshehri, A.M. Metabolic syndrome and cardiovascular risk. J. Fam. Community Med. 2010, 17, 73–78. [Google Scholar] [CrossRef]
  5. Bosho, D.D.; Dube, L.; Mega, T.A.; Adare, D.A.; Tesfaye, M.G.; Eshetie, T.C. Prevalence and predictors of metabolic syndrome among people living with human immunodeficiency virus (PLWHIV). Diabetol. Metab. Syndr. 2018, 10, 10. [Google Scholar] [CrossRef]
  6. Gami, A.S.; Witt, B.J.; Howard, D.E.; Erwin, P.J.; Gami, L.A.; Somers, V.K.; Montori, V.M. Metabolic syndrome and risk of incident cardiovascular events and death: A systematic review and meta-analysis of longitudinal studies. J. Am. Coll. Cardiol. 2007, 49, 403–414. [Google Scholar] [CrossRef]
  7. Zevin, A.S.; McKinnon, L.; Burgener, A.; Klatt, N.R. Microbial translocation and microbiome dysbiosis in HIV-associated immune activation. Curr. Opin. HIV AIDS 2016, 11, 182–190. [Google Scholar] [CrossRef] [PubMed]
  8. Raposo, M.A.; Armiliato, G.N.A.; Guimaraes, N.S.; Caram, C.A.; Silveira, R.D.S.; Tupinambas, U. Metabolic disorders and cardiovascular risk in people living with HIV/AIDS without the use of antiretroviral therapy. Rev. Soc. Bras. Med. Trop. 2017, 50, 598–606. [Google Scholar] [CrossRef]
  9. Deeks, S.G.; Wrin, T.; Liegler, T.; Hoh, R.; Hayden, M.; Barbour, J.D.; Hellmann, N.S.; Petropoulos, C.J.; McCune, J.M.; Hellerstein, M.K.; et al. Virologic and immunologic consequences of discontinuing combination antiretroviral-drug therapy in HIV-infected patients with detectable viremia. N. Engl. J. Med. 2001, 344, 472–480. [Google Scholar] [CrossRef] [PubMed]
  10. Persaud, D.; Pierson, T.; Ruff, C.; Finzi, D.; Chadwick, K.R.; Margolick, J.B.; Ruff, A.; Hutton, N.; Ray, S.; Siliciano, R.F. A stable latent reservoir for HIV-1 in resting CD4(+) T lymphocytes in infected children. J. Clin. Investig. 2000, 105, 995–1003. [Google Scholar] [CrossRef] [PubMed]
  11. Chun, T.W.; Stuyver, L.; Mizell, S.B.; Ehler, L.A.; Mican, J.A.; Baseler, M.; Lloyd, A.L.; Nowak, M.A.; Fauci, A.S. Presence of an inducible HIV-1 latent reservoir during highly active antiretroviral therapy. Proc. Natl. Acad. Sci. USA 1997, 94, 13193–13197. [Google Scholar] [CrossRef] [PubMed]
  12. Finzi, D.; Blankson, J.; Siliciano, J.D.; Margolick, J.B.; Chadwick, K.; Pierson, T.; Smith, K.; Lisziewicz, J.; Lori, F.; Flexner, C.; et al. Latent infection of CD4+ T cells provides a mechanism for lifelong persistence of HIV-1, even in patients on effective combination therapy. Nat. Med. 1999, 5, 512–517. [Google Scholar] [CrossRef] [PubMed]
  13. Finzi, D.; Hermankova, M.; Pierson, T.; Carruth, L.M.; Buck, C.; Chaisson, R.E.; Quinn, T.C.; Chadwick, K.; Margolick, J.; Brookmeyer, R.; et al. Identification of a reservoir for HIV-1 in patients on highly active antiretroviral therapy. Science 1997, 278, 1295–1300. [Google Scholar] [CrossRef]
  14. Barton, K.; Winckelmann, A.; Palmer, S. HIV-1 Reservoirs during Suppressive Therapy. Trends Microbiol. 2016, 24, 345–355. [Google Scholar] [CrossRef]
  15. Alexaki, A.; Liu, Y.; Wigdahl, B. Cellular reservoirs of HIV-1 and their role in viral persistence. Curr. HIV Res. 2008, 6, 388–400. [Google Scholar] [CrossRef]
  16. Lutz, C.T.; Karapetyan, A.; Al-Attar, A.; Shelton, B.J.; Holt, K.J.; Tucker, J.H.; Presnell, S.R. Human NK cells proliferate and die in vivo more rapidly than T cells in healthy young and elderly adults. J. Immunol. 2011, 186, 4590–4598. [Google Scholar] [CrossRef]
  17. Patel, A.A.; Ginhoux, F.; Yona, S. Monocytes, macrophages, dendritic cells and neutrophils: An update on lifespan kinetics in health and disease. Immunology 2021, 163, 250–261. [Google Scholar] [CrossRef] [PubMed]
  18. Banga, R.; Procopio, F.A.; Lana, E.; Gladkov, G.T.; Roseto, I.; Parsons, E.M.; Lian, X.; Armani-Tourret, M.; Bellefroid, M.; Gao, C.; et al. Lymph node dendritic cells harbor inducible replication-competent HIV despite years of suppressive ART. Cell Host Microbe 2023, 31, 1714–1731. [Google Scholar] [CrossRef]
  19. Tang, Y.; Chaillon, A.; Gianella, S.; Wong, L.M.; Li, D.; Simermeyer, T.L.; Porrachia, M.; Ignacio, C.; Woodworth, B.; Zhong, D.; et al. Brain microglia serve as a persistent HIV reservoir despite durable antiretroviral therapy. J. Clin. Investig. 2023, 133, e167417. [Google Scholar] [CrossRef]
  20. Parbie, P.K.; Mizutani, T.; Ishizaka, A.; Kawana-Tachikawa, A.; Runtuwene, L.R.; Seki, S.; Abana, C.Z.; Kushitor, D.; Bonney, E.Y.; Ofori, S.B.; et al. Dysbiotic Fecal Microbiome in HIV-1 Infected Individuals in Ghana. Front. Cell Infect. Microbiol. 2021, 11, 646467. [Google Scholar] [CrossRef]
  21. Barre-Sinoussi, F.; Ross, A.L.; Delfraissy, J.F. Past, present and future: 30 years of HIV research. Nat. Rev. Microbiol. 2013, 11, 877–883. [Google Scholar] [CrossRef] [PubMed]
  22. International, A.S.S.W.G.o.H.I.V.C.; Deeks, S.G.; Autran, B.; Berkhout, B.; Benkirane, M.; Cairns, S.; Chomont, N.; Chun, T.W.; Churchill, M.; Di Mascio, M.; et al. Towards an HIV cure: A global scientific strategy. Nat. Rev. Immunol. 2012, 12, 607–614. [Google Scholar] [CrossRef]
  23. Darcis, G.; Van Driessche, B.; Van Lint, C. HIV Latency: Should We Shock or Lock? Trends Immunol. 2017, 38, 217–228. [Google Scholar] [CrossRef] [PubMed]
  24. Whitcomb, J.M.; Kumar, R.; Hughes, S.H. Sequence of the circle junction of human immunodeficiency virus type 1: Implications for reverse transcription and integration. J. Virol. 1990, 64, 4903–4906. [Google Scholar] [CrossRef]
  25. Coffin, J.M.; Hughes, S.H.; Varmus, H.E. (Eds.) Retroviruses; Cold Spring Harbor: New York, NY, USA, 1997. [Google Scholar]
  26. Schiralli Lester, G.M.; Henderson, A.J. Mechanisms of HIV Transcriptional Regulation and Their Contribution to Latency. Mol. Biol. Int. 2012, 2012, 614120. [Google Scholar] [CrossRef]
  27. Bannwarth, S.; Gatignol, A. HIV-1 TAR RNA: The target of molecular interactions between the virus and its host. Curr. HIV Res. 2005, 3, 61–71. [Google Scholar] [CrossRef]
  28. Kilareski, E.M.; Shah, S.; Nonnemacher, M.R.; Wigdahl, B. Regulation of HIV-1 transcription in cells of the monocyte-macrophage lineage. Retrovirology 2009, 6, 118. [Google Scholar] [CrossRef]
  29. Pereira, L.A.; Bentley, K.; Peeters, A.; Churchill, M.J.; Deacon, N.J. A compilation of cellular transcription factor interactions with the HIV-1 LTR promoter. Nucleic Acids Res. 2000, 28, 663–668. [Google Scholar] [CrossRef]
  30. Rohr, O.; Marban, C.; Aunis, D.; Schaeffer, E. Regulation of HIV-1 gene transcription: From lymphocytes to microglial cells. J. Leukoc. Biol. 2003, 74, 736–749. [Google Scholar] [CrossRef]
  31. Li, G.M.Y.; Franza, B.R., Jr. In vitro study of functional involvement of Sp1, NF-kappa B/Rel, and AP1 in phorbol 12-myristate 13-acetate-mediated HIV-1 long terminal repeat activation. J. Biol. Chem. 1994, 269, 30616–30619. [Google Scholar] [CrossRef]
  32. Majello, B.; De Luca, P.; Hagen, G.; Suske, G.; Lania, L. Different members of the Sp1 multigene family exert opposite transcriptional regulation of the long terminal repeat of HIV-1. Nucleic Acids Res. 1994, 22, 4914–4921. [Google Scholar] [CrossRef] [PubMed]
  33. Perkins, N.D.; Edwards, N.L.; Duckett, C.S.; Agranoff, A.B.; Schmid, R.M.; Nabel, G.J. A cooperative interaction between NF-kappa B and Sp1 is required for HIV-1 enhancer activation. EMBO J. 1993, 12, 3551–3558. [Google Scholar] [CrossRef] [PubMed]
  34. Yamaguchi, Y.; Takagi, T.; Wada, T.; Yano, K.; Furuya, A.; Sugimoto, S.; Hasegawa, J.; Handa, H. NELF, a multisubunit complex containing RD, cooperates with DSIF to repress RNA polymerase II elongation. Cell 1999, 97, 41–51. [Google Scholar] [CrossRef] [PubMed]
  35. He, G.; Margolis, D.M. Counterregulation of chromatin deacetylation and histone deacetylase occupancy at the integrated promoter of human immunodeficiency virus type 1 (HIV-1) by the HIV-1 repressor YY1 and HIV-1 activator Tat. Mol. Cell Biol. 2002, 22, 2965–2973. [Google Scholar] [CrossRef]
  36. Bernhard, W.; Barreto, K.; Raithatha, S.; Sadowski, I. An upstream YY1 binding site on the HIV-1 LTR contributes to latent infection. PLoS ONE 2013, 8, e77052. [Google Scholar] [CrossRef] [PubMed]
  37. Imai, K.; Okamoto, T. Transcriptional repression of human immunodeficiency virus type 1 by AP-4. J. Biol. Chem. 2006, 281, 12495–12505. [Google Scholar] [CrossRef]
  38. Canonne-Hergaux, F.; Aunis, D.; Schaeffer, E. Interactions of the transcription factor AP-1 with the long terminal repeat of different human immunodeficiency virus type 1 strains in Jurkat, glial, and neuronal cells. J. Virol. 1995, 69, 6634–6642. [Google Scholar] [CrossRef]
  39. Roebuck, K.A.; Gu, D.S.; Kagnoff, M.F. Activating protein-1 cooperates with phorbol ester activation signals to increase HIV-1 expression. AIDS 1996, 10, 819–826. [Google Scholar] [CrossRef]
  40. Roebuck, K.A.; Saifuddin, M. Regulation of HIV-1 transcription. Gene Expr. 1999, 8, 67–84. [Google Scholar]
  41. Schroder, A.R.; Shinn, P.; Chen, H.; Berry, C.; Ecker, J.R.; Bushman, F. HIV-1 integration in the human genome favors active genes and local hotspots. Cell 2002, 110, 521–529. [Google Scholar] [CrossRef]
  42. Bushman, F.; Lewinski, M.; Ciuffi, A.; Barr, S.; Leipzig, J.; Hannenhalli, S.; Hoffmann, C. Genome-wide analysis of retroviral DNA integration. Nat. Rev. Microbiol. 2005, 3, 848–858. [Google Scholar] [CrossRef]
  43. Lewinski, M.K.; Bisgrove, D.; Shinn, P.; Chen, H.; Hoffmann, C.; Hannenhalli, S.; Verdin, E.; Berry, C.C.; Ecker, J.R.; Bushman, F.D. Genome-wide analysis of chromosomal features repressing human immunodeficiency virus transcription. J. Virol. 2005, 79, 6610–6619. [Google Scholar] [CrossRef]
  44. Lewinski, M.K.; Yamashita, M.; Emerman, M.; Ciuffi, A.; Marshall, H.; Crawford, G.; Collins, F.; Shinn, P.; Leipzig, J.; Hannenhalli, S.; et al. Retroviral DNA integration: Viral and cellular determinants of target-site selection. PLoS Pathog. 2006, 2, e60. [Google Scholar] [CrossRef]
  45. Duverger, A.; Jones, J.; May, J.; Bibollet-Ruche, F.; Wagner, F.A.; Cron, R.Q.; Kutsch, O. Determinants of the establishment of human immunodeficiency virus type 1 latency. J. Virol. 2009, 83, 3078–3093. [Google Scholar] [CrossRef]
  46. Jones, K.A.; Kadonaga, J.T.; Luciw, P.A.; Tjian, R. Activation of the AIDS retrovirus promoter by the cellular transcription factor, Sp1. Science 1986, 232, 755–759. [Google Scholar] [CrossRef]
  47. Shaw, J.P.; Utz, P.J.; Durand, D.B.; Toole, J.J.; Emmel, E.A.; Crabtree, G.R. Identification of a putative regulator of early T cell activation genes. Science 1988, 241, 202–205. [Google Scholar] [CrossRef]
  48. Osborn, L.; Kunkel, S.; Nabel, G.J. Tumor necrosis factor alpha and interleukin 1 stimulate the human immunodeficiency virus enhancer by activation of the nuclear factor kappa B. Proc. Natl. Acad. Sci. USA 1989, 86, 2336–2340. [Google Scholar] [CrossRef]
  49. Ni, Z.; Saunders, A.; Fuda, N.J.; Yao, J.; Suarez, J.R.; Webb, W.W.; Lis, J.T. P-TEFb is critical for the maturation of RNA polymerase II into productive elongation in vivo. Mol. Cell Biol. 2008, 28, 1161–1170. [Google Scholar] [CrossRef]
  50. Ping, Y.H.; Rana, T.M. DSIF and NELF interact with RNA polymerase II elongation complex and HIV-1 Tat stimulates P-TEFb-mediated phosphorylation of RNA polymerase II and DSIF during transcription elongation. J. Biol. Chem. 2001, 276, 12951–12958. [Google Scholar] [CrossRef]
  51. Peterlin, B.M.; Price, D.H. Controlling the elongation phase of transcription with P-TEFb. Mol. Cell 2006, 23, 297–305. [Google Scholar] [CrossRef] [PubMed]
  52. Karn, J. The molecular biology of HIV latency: Breaking and restoring the Tat-dependent transcriptional circuit. Curr. Opin. HIV AIDS 2011, 6, 4–11. [Google Scholar] [CrossRef]
  53. Siliciano, R.F.; Greene, W.C. HIV latency. Cold Spring Harb. Perspect. Med. 2011, 1, a007096. [Google Scholar] [CrossRef] [PubMed]
  54. Arya, S.K.; Gallo, R.C.; Hahn, B.H.; Shaw, G.M.; Popovic, M.; Salahuddin, S.Z.; Wong-Staal, F. Homology of genome of AIDS-associated virus with genomes of human T-cell leukemia viruses. Science 1984, 225, 927–930. [Google Scholar] [CrossRef] [PubMed]
  55. Rosen, C.A.; Sodroski, J.G.; Haseltine, W.A. The location of cis-acting regulatory sequences in the human T cell lymphotropic virus type III (HTLV-III/LAV) long terminal repeat. Cell 1985, 41, 813–823. [Google Scholar] [CrossRef] [PubMed]
  56. Michels, A.A.; Fraldi, A.; Li, Q.; Adamson, T.E.; Bonnet, F.; Nguyen, V.T.; Sedore, S.C.; Price, J.P.; Price, D.H.; Lania, L.; et al. Binding of the 7SK snRNA turns the HEXIM1 protein into a P-TEFb (CDK9/cyclin T) inhibitor. EMBO J. 2004, 23, 2608–2619. [Google Scholar] [CrossRef] [PubMed]
  57. 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] [PubMed]
  58. Pumfery, A.; Deng, L.; Maddukuri, A.; de la Fuente, C.; Li, H.; Wade, J.D.; Lambert, P.; Kumar, A.; Kashanchi, F. Chromatin remodeling and modification during HIV-1 Tat-activated transcription. Curr. HIV Res. 2003, 1, 343–362. [Google Scholar] [CrossRef]
  59. Machida, S.; Depierre, D.; Chen, H.C.; Thenin-Houssier, S.; Petitjean, G.; Doyen, C.M.; Takaku, M.; Cuvier, O.; Benkirane, M. Exploring histone loading on HIV DNA reveals a dynamic nucleosome positioning between unintegrated and integrated viral genome. Proc. Natl. Acad. Sci. USA 2020, 117, 6822–6830. [Google Scholar] [CrossRef]
  60. Rafati, H.; Parra, M.; Hakre, S.; Moshkin, Y.; Verdin, E.; Mahmoudi, T. Repressive LTR nucleosome positioning by the BAF complex is required for HIV latency. PLoS Biol. 2011, 9, e1001206. [Google Scholar] [CrossRef]
  61. Boekhoudt, G.H.; Guo, Z.; Beresford, G.W.; Boss, J.M. Communication between NF-kappa B and Sp1 controls histone acetylation within the proximal promoter of the monocyte chemoattractant protein 1 gene. J. Immunol. 2003, 170, 4139–4147. [Google Scholar] [CrossRef]
  62. Bhatt, D.; Ghosh, S. Regulation of the NF-kappaB-Mediated Transcription of Inflammatory Genes. Front. Immunol. 2014, 5, 71. [Google Scholar] [CrossRef] [PubMed]
  63. Petty, E.; Pillus, L. Balancing chromatin remodeling and histone modifications in transcription. Trends Genet. 2013, 29, 621–629. [Google Scholar] [CrossRef] [PubMed]
  64. Ajasin, D.; Eugenin, E.A. HIV-1 Tat: Role in Bystander Toxicity. Front. Cell Infect. Microbiol. 2020, 10, 61. [Google Scholar] [CrossRef]
  65. Besnard, E.; Hakre, S.; Kampmann, M.; Lim, H.W.; Hosmane, N.N.; Martin, A.; Bassik, M.C.; Verschueren, E.; Battivelli, E.; Chan, J.; et al. The mTOR Complex Controls HIV Latency. Cell Host Microbe 2016, 20, 785–797. [Google Scholar] [CrossRef]
  66. Pasquereau, S.; Herbein, G. CounterAKTing HIV: Toward a “Block and Clear” Strategy? Front. Cell Infect. Microbiol. 2022, 12, 827717. [Google Scholar] [CrossRef]
  67. Cary, D.C.; Fujinaga, K.; Peterlin, B.M. Molecular mechanisms of HIV latency. J. Clin. Investig. 2016, 126, 448–454. [Google Scholar] [CrossRef] [PubMed]
  68. Ahlenstiel, C.L.; Symonds, G.; Kent, S.J.; Kelleher, A.D. Block and Lock HIV Cure Strategies to Control the Latent Reservoir. Front. Cell Infect. Microbiol. 2020, 10, 424. [Google Scholar] [CrossRef]
  69. Li, C.; Mori, L.P.; Lyu, S.; Bronson, R.; Getzler, A.J.; Pipkin, M.E.; Valente, S.T. The chaperone protein p32 stabilizes HIV-1 Tat and strengthens the p-TEFb/RNAPII/TAR complex promoting HIV transcription elongation. Proc. Natl. Acad. Sci. USA 2023, 120, e2217476120. [Google Scholar] [CrossRef]
  70. Mendez, C.; Ledger, S.; Petoumenos, K.; Ahlenstiel, C.; Kelleher, A.D. RNA-induced epigenetic silencing inhibits HIV-1 reactivation from latency. Retrovirology 2018, 15, 67. [Google Scholar] [CrossRef]
  71. Vansant, G.; Bruggemans, A.; Janssens, J.; Debyser, Z. Block-And-Lock Strategies to Cure HIV Infection. Viruses 2020, 12, 84. [Google Scholar] [CrossRef]
  72. Surabhi, R.M.; Gaynor, R.B. RNA interference directed against viral and cellular targets inhibits human immunodeficiency Virus Type 1 replication. J. Virol. 2002, 76, 12963–12973. [Google Scholar] [CrossRef] [PubMed]
  73. Suzuki, K.; Shijuuku, T.; Fukamachi, T.; Zaunders, J.; Guillemin, G.; Cooper, D.; Kelleher, A. Prolonged transcriptional silencing and CpG methylation induced by siRNAs targeted to the HIV-1 promoter region. J. RNAi Gene Silencing 2005, 1, 66–78. [Google Scholar] [PubMed]
  74. Suzuki, K.; Ishida, T.; Yamagishi, M.; Ahlenstiel, C.; Swaminathan, S.; Marks, K.; Murray, D.; McCartney, E.M.; Beard, M.R.; Alexander, M.; et al. Transcriptional gene silencing of HIV-1 through promoter targeted RNA is highly specific. RNA Biol. 2011, 8, 1035–1046. [Google Scholar] [CrossRef] [PubMed]
  75. Suzuki, K.; Ahlenstiel, C.; Marks, K.; Kelleher, A.D. Promoter Targeting RNAs: Unexpected Contributors to the Control of HIV-1 Transcription. Mol. Ther. Nucleic Acids 2015, 4, e222. [Google Scholar] [CrossRef] [PubMed]
  76. Tsukamoto, T. Transcriptional gene silencing limits CXCR4-associated depletion of bone marrow CD34+ cells in HIV-1 infection. AIDS 2018, 32, 1737–1747. [Google Scholar] [CrossRef]
  77. Yamagishi, M.; Ishida, T.; Miyake, A.; Cooper, D.A.; Kelleher, A.D.; Suzuki, K.; Watanabe, T. Retroviral delivery of promoter-targeted shRNA induces long-term silencing of HIV-1 transcription. Microbes Infect. 2009, 11, 500–508. [Google Scholar] [CrossRef]
  78. Weinberg, M.S.; Villeneuve, L.M.; Ehsani, A.; Amarzguioui, M.; Aagaard, L.; Chen, Z.X.; Riggs, A.D.; Rossi, J.J.; Morris, K.V. The antisense strand of small interfering RNAs directs histone methylation and transcriptional gene silencing in human cells. RNA 2006, 12, 256–262. [Google Scholar] [CrossRef]
  79. Singh, A.; Palanichamy, J.K.; Ramalingam, P.; Kassab, M.A.; Bhagat, M.; Andrabi, R.; Luthra, K.; Sinha, S.; Chattopadhyay, P. Long-term suppression of HIV-1C virus production in human peripheral blood mononuclear cells by LTR heterochromatization with a short double-stranded RNA. J. Antimicrob. Chemother. 2014, 69, 404–415. [Google Scholar] [CrossRef]
  80. Hu, B.; Zhong, L.; Weng, Y.; Peng, L.; Huang, Y.; Zhao, Y.; Liang, X.J. Therapeutic siRNA: State of the art. Signal Transduct. Target. Ther. 2020, 5, 101. [Google Scholar] [CrossRef]
  81. Martinez, M.A. Progress in the therapeutic applications of siRNAs against HIV-1. Methods Mol. Biol. 2009, 487, 343–368. [Google Scholar] [CrossRef]
  82. Ray, R.M.; Morris, K.V. Long Non-coding RNAs Mechanisms of Action in HIV-1 Modulation and the Identification of Novel Therapeutic Targets. Noncoding RNA 2020, 6, 12. [Google Scholar] [CrossRef]
  83. Kopp, F.; Mendell, J.T. Functional Classification and Experimental Dissection of Long Noncoding RNAs. Cell 2018, 172, 393–407. [Google Scholar] [CrossRef] [PubMed]
  84. Ramirez, P.W.; Pantoja, C.; Beliakova-Bethell, N. An Evaluation on the Role of Non-Coding RNA in HIV Transcription and Latency: A Review. HIV AIDS 2023, 15, 115–134. [Google Scholar] [CrossRef] [PubMed]
  85. Saayman, S.; Ackley, A.; Turner, A.W.; Famiglietti, M.; Bosque, A.; Clemson, M.; Planelles, V.; Morris, K.V. An HIV-encoded antisense long noncoding RNA epigenetically regulates viral transcription. Mol. Ther. 2014, 22, 1164–1175. [Google Scholar] [CrossRef] [PubMed]
  86. Imam, H.; Bano, A.S.; Patel, P.; Holla, P.; Jameel, S. The lncRNA NRON modulates HIV-1 replication in a NFAT-dependent manner and is differentially regulated by early and late viral proteins. Sci. Rep. 2015, 5, 8639. [Google Scholar] [CrossRef] [PubMed]
  87. Kobayashi-Ishihara, M.; Yamagishi, M.; Hara, T.; Matsuda, Y.; Takahashi, R.; Miyake, A.; Nakano, K.; Yamochi, T.; Ishida, T.; Watanabe, T. HIV-1-encoded antisense RNA suppresses viral replication for a prolonged period. Retrovirology 2012, 9, 38. [Google Scholar] [CrossRef]
  88. Li, H.; Chi, X.; Li, R.; Ouyang, J.; Chen, Y. A Novel lncRNA, AK130181, Contributes to HIV-1 Latency by Regulating Viral Promoter-Driven Gene Expression in Primary CD4(+) T Cells. Mol. Ther. Nucleic Acids 2020, 20, 754–763. [Google Scholar] [CrossRef]
  89. Li, J.; Chen, C.; Ma, X.; Geng, G.; Liu, B.; Zhang, Y.; Zhang, S.; Zhong, F.; Liu, C.; Yin, Y.; et al. Long noncoding RNA NRON contributes to HIV-1 latency by specifically inducing tat protein degradation. Nat. Commun. 2016, 7, 11730. [Google Scholar] [CrossRef] [PubMed]
  90. Trypsteen, W.; White, C.H.; Mukim, A.; Spina, C.A.; De Spiegelaere, W.; Lefever, S.; Planelles, V.; Bosque, A.; Woelk, C.H.; Vandekerckhove, L.; et al. Long non-coding RNAs and latent HIV—A search for novel targets for latency reversal. PLoS ONE 2019, 14, e0224879. [Google Scholar] [CrossRef]
  91. Wang, H.; Liu, Y.; Huan, C.; Yang, J.; Li, Z.; Zheng, B.; Wang, Y.; Zhang, W. NF-kappaB-Interacting Long Noncoding RNA Regulates HIV-1 Replication and Latency by Repressing NF-kappaB Signaling. J. Virol. 2020, 94, e01057-20. [Google Scholar] [CrossRef]
  92. Zhang, Q.; Chen, C.Y.; Yedavalli, V.S.; Jeang, K.T. NEAT1 long noncoding RNA and paraspeckle bodies modulate HIV-1 posttranscriptional expression. mBio 2013, 4, e00596-12. [Google Scholar] [CrossRef]
  93. Rice, A.P. The HIV-1 Tat Protein: Mechanism of Action and Target for HIV-1 Cure Strategies. Curr. Pharm. Des. 2017, 23, 4098–4102. [Google Scholar] [CrossRef]
  94. Nekhai, S.; Jeang, K.T. Transcriptional and post-transcriptional regulation of HIV-1 gene expression: Role of cellular factors for Tat and Rev. Future Microbiol. 2006, 1, 417–426. [Google Scholar] [CrossRef]
  95. Vardabasso, C.; Manganaro, L.; Lusic, M.; Marcello, A.; Giacca, M. The histone chaperone protein Nucleosome Assembly Protein-1 (hNAP-1) binds HIV-1 Tat and promotes viral transcription. Retrovirology 2008, 5, 8. [Google Scholar] [CrossRef] [PubMed]
  96. Meredith, L.W.; Sivakumaran, H.; Major, L.; Suhrbier, A.; Harrich, D. Potent inhibition of HIV-1 replication by a Tat mutant. PLoS ONE 2009, 4, e7769. [Google Scholar] [CrossRef]
  97. Lin, M.H.; Sivakumaran, H.; Apolloni, A.; Wei, T.; Jans, D.A.; Harrich, D. Nullbasic, a potent anti-HIV tat mutant, induces CRM1-dependent disruption of HIV rev trafficking. PLoS ONE 2012, 7, e51466. [Google Scholar] [CrossRef] [PubMed]
  98. Jin, H.; Li, D.; Sivakumaran, H.; Lor, M.; Rustanti, L.; Cloonan, N.; Wani, S.; Harrich, D. Shutdown of HIV-1 Transcription in T Cells by Nullbasic, a Mutant Tat Protein. mBio 2016, 7, e00518-16. [Google Scholar] [CrossRef]
  99. Jin, H.; Sun, Y.; Li, D.; Lin, M.H.; Lor, M.; Rustanti, L.; Harrich, D. Strong In Vivo Inhibition of HIV-1 Replication by Nullbasic, a Tat Mutant. mBio 2019, 10, e01769-19. [Google Scholar] [CrossRef] [PubMed]
  100. Mousseau, G.; Valente, S.T. Didehydro-Cortistatin A: A new player in HIV-therapy? Expert. Rev. Anti Infect. Ther. 2016, 14, 145–148. [Google Scholar] [CrossRef]
  101. Mousseau, G.; Aneja, R.; Clementz, M.A.; Mediouni, S.; Lima, N.S.; Haregot, A.; Kessing, C.F.; Jablonski, J.A.; Thenin-Houssier, S.; Nagarsheth, N.; et al. Resistance to the Tat Inhibitor Didehydro-Cortistatin A Is Mediated by Heightened Basal HIV-1 Transcription. mBio 2019, 10, e01750-18. [Google Scholar] [CrossRef]
  102. Mousseau, G.; Kessing, C.F.; Fromentin, R.; Trautmann, L.; Chomont, N.; Valente, S.T. The Tat Inhibitor Didehydro-Cortistatin A Prevents HIV-1 Reactivation from Latency. mBio 2015, 6, e00465. [Google Scholar] [CrossRef] [PubMed]
  103. Li, C.; Mousseau, G.; Valente, S.T. Tat inhibition by didehydro-Cortistatin A promotes heterochromatin formation at the HIV-1 long terminal repeat. Epigenetics Chromatin 2019, 12, 23. [Google Scholar] [CrossRef]
  104. Kessing, C.F.; Nixon, C.C.; Li, C.; Tsai, P.; Takata, H.; Mousseau, G.; Ho, P.T.; Honeycutt, J.B.; Fallahi, M.; Trautmann, L.; et al. In Vivo Suppression of HIV Rebound by Didehydro-Cortistatin A, a “Block-and-Lock” Strategy for HIV-1 Treatment. Cell Rep. 2017, 21, 600–611. [Google Scholar] [CrossRef] [PubMed]
  105. Mediouni, S.; Chinthalapudi, K.; Ekka, M.K.; Usui, I.; Jablonski, J.A.; Clementz, M.A.; Mousseau, G.; Nowak, J.; Macherla, V.R.; Beverage, J.N.; et al. Didehydro-Cortistatin A Inhibits HIV-1 by Specifically Binding to the Unstructured Basic Region of Tat. mBio 2019, 10, e02662-18. [Google Scholar] [CrossRef] [PubMed]
  106. Siliciano, J.D.; Siliciano, R.F. Recent developments in the search for a cure for HIV-1 infection: Targeting the latent reservoir for HIV-1. J. Allergy Clin. Immunol. 2014, 134, 12–19. [Google Scholar] [CrossRef] [PubMed]
  107. Archin, N.M.; Liberty, A.L.; Kashuba, A.D.; Choudhary, S.K.; Kuruc, J.D.; Crooks, A.M.; Parker, D.C.; Anderson, E.M.; Kearney, M.F.; Strain, M.C.; et al. Administration of vorinostat disrupts HIV-1 latency in patients on antiretroviral therapy. Nature 2012, 487, 482–485. [Google Scholar] [CrossRef] [PubMed]
  108. Gay, C.L.; James, K.S.; Tuyishime, M.; Falcinelli, S.D.; Joseph, S.B.; Moeser, M.J.; Allard, B.; Kirchherr, J.L.; Clohosey, M.; Raines, S.L.M.; et al. Stable Latent HIV Infection and Low-level Viremia Despite Treatment with the Broadly Neutralizing Antibody VRC07-523LS and the Latency Reversal Agent Vorinostat. J. Infect. Dis. 2022, 225, 856–861. [Google Scholar] [CrossRef] [PubMed]
  109. Gruell, H.; Gunst, J.D.; Cohen, Y.Z.; Pahus, M.H.; Malin, J.J.; Platten, M.; Millard, K.G.; Tolstrup, M.; Jones, R.B.; Conce Alberto, W.D.; et al. Effect of 3BNC117 and romidepsin on the HIV-1 reservoir in people taking suppressive antiretroviral therapy (ROADMAP): A randomised, open-label, phase 2A trial. Lancet Microbe 2022, 3, e203–e214. [Google Scholar] [CrossRef]
  110. Gutierrez, C.; Serrano-Villar, S.; Madrid-Elena, N.; Perez-Elias, M.J.; Martin, M.E.; Barbas, C.; Ruiperez, J.; Munoz, E.; Munoz-Fernandez, M.A.; Castor, T.; et al. Bryostatin-1 for latent virus reactivation in HIV-infected patients on antiretroviral therapy. AIDS 2016, 30, 1385–1392. [Google Scholar] [CrossRef]
  111. Kroon, E.; Ananworanich, J.; Pagliuzza, A.; Rhodes, A.; Phanuphak, N.; Trautmann, L.; Mitchell, J.L.; Chintanaphol, M.; Intasan, J.; Pinyakorn, S.; et al. A randomized trial of vorinostat with treatment interruption after initiating antiretroviral therapy during acute HIV-1 infection. J. Virus Erad. 2020, 6, 100004. [Google Scholar] [CrossRef]
  112. McMahon, D.K.; Zheng, L.; Cyktor, J.C.; Aga, E.; Macatangay, B.J.; Godfrey, C.; Para, M.; Mitsuyasu, R.T.; Hesselgesser, J.; Dragavon, J.; et al. A Phase 1/2 Randomized, Placebo-Controlled Trial of Romidespin in Persons with HIV-1 on Suppressive Antiretroviral Therapy. J. Infect. Dis. 2021, 224, 648–656. [Google Scholar] [CrossRef]
  113. Rasmussen, T.A.; Tolstrup, M.; Brinkmann, C.R.; Olesen, R.; Erikstrup, C.; Solomon, A.; Winckelmann, A.; Palmer, S.; Dinarello, C.; Buzon, M.; et al. Panobinostat, a histone deacetylase inhibitor, for latent-virus reactivation in HIV-infected patients on suppressive antiretroviral therapy: A phase 1/2, single group, clinical trial. Lancet HIV 2014, 1, e13–e21. [Google Scholar] [CrossRef] [PubMed]
  114. Riddler, S. Vesatolimod (GS-9620) is Safe and Pharmacodynamically Active in HIV-Infected Individuals. In Proceedings of the International AIDS Society Conference on HIV Science, Mexico City, Mexico, 21–24 July 2019. [Google Scholar]
  115. Sogaard, O.S.; Graversen, M.E.; Leth, S.; Olesen, R.; Brinkmann, C.R.; Nissen, S.K.; Kjaer, A.S.; Schleimann, M.H.; Denton, P.W.; Hey-Cunningham, W.J.; et al. The Depsipeptide Romidepsin Reverses HIV-1 Latency In Vivo. PLoS Pathog. 2015, 11, e1005142. [Google Scholar] [CrossRef]
  116. Vibholm, L.; Schleimann, M.H.; Hojen, J.F.; Benfield, T.; Offersen, R.; Rasmussen, K.; Olesen, R.; Dige, A.; Agnholt, J.; Grau, J.; et al. Short-Course Toll-Like Receptor 9 Agonist Treatment Impacts Innate Immunity and Plasma Viremia in Individuals With Human Immunodeficiency Virus Infection. Clin. Infect. Dis. 2017, 64, 1686–1695. [Google Scholar] [CrossRef] [PubMed]
  117. Vibholm, L.K.; Konrad, C.V.; Schleimann, M.H.; Frattari, G.; Winckelmann, A.; Klastrup, V.; Jensen, N.M.; Jensen, S.S.; Schmidt, M.; Wittig, B.; et al. Effects of 24-week Toll-like receptor 9 agonist treatment in HIV type 1+ individuals. AIDS 2019, 33, 1315–1325. [Google Scholar] [CrossRef] [PubMed]
  118. Dahabieh, M.S.; Battivelli, E.; Verdin, E. Understanding HIV latency: The road to an HIV cure. Annu. Rev. Med. 2015, 66, 407–421. [Google Scholar] [CrossRef]
  119. Marsden, M.D.; Loy, B.A.; Wu, X.; Ramirez, C.M.; Schrier, A.J.; Murray, D.; Shimizu, A.; Ryckbosch, S.M.; Near, K.E.; Chun, T.W.; et al. In vivo activation of latent HIV with a synthetic bryostatin analog effects both latent cell “kick” and “kill” in strategy for virus eradication. PLoS Pathog. 2017, 13, e1006575. [Google Scholar] [CrossRef]
  120. Chao, T.C.; Zhang, Q.; Li, Z.; Tiwari, S.K.; Qin, Y.; Yau, E.; Sanchez, A.; Singh, G.; Chang, K.; Kaul, M.; et al. The Long Noncoding RNA HEAL Regulates HIV-1 Replication through Epigenetic Regulation of the HIV-1 Promoter. mBio 2019, 10, e02016-19. [Google Scholar] [CrossRef]
  121. Qu, D.; Sun, W.W.; Li, L.; Ma, L.; Sun, L.; Jin, X.; Li, T.; Hou, W.; Wang, J.H. Long noncoding RNA MALAT1 releases epigenetic silencing of HIV-1 replication by displacing the polycomb repressive complex 2 from binding to the LTR promoter. Nucleic Acids Res. 2019, 47, 3013–3027. [Google Scholar] [CrossRef]
  122. Taunton, J.; Hassig, C.A.; Schreiber, S.L. A mammalian histone deacetylase related to the yeast transcriptional regulator Rpd3p. Science 1996, 272, 408–411. [Google Scholar] [CrossRef]
  123. Yang, W.M.; Inouye, C.; Zeng, Y.; Bearss, D.; Seto, E. Transcriptional repression by YY1 is mediated by interaction with a mammalian homolog of the yeast global regulator RPD3. Proc. Natl. Acad. Sci. USA 1996, 93, 12845–12850. [Google Scholar] [CrossRef]
  124. Hu, E.; Chen, Z.; Fredrickson, T.; Zhu, Y.; Kirkpatrick, R.; Zhang, G.F.; Johanson, K.; Sung, C.M.; Liu, R.; Winkler, J. Cloning and characterization of a novel human class I histone deacetylase that functions as a transcription repressor. J. Biol. Chem. 2000, 275, 15254–15264. [Google Scholar] [CrossRef] [PubMed]
  125. Grozinger, C.M.; Hassig, C.A.; Schreiber, S.L. Three proteins define a class of human histone deacetylases related to yeast Hda1p. Proc. Natl. Acad. Sci. USA 1999, 96, 4868–4873. [Google Scholar] [CrossRef] [PubMed]
  126. Kao, H.Y.; Downes, M.; Ordentlich, P.; Evans, R.M. Isolation of a novel histone deacetylase reveals that class I and class II deacetylases promote SMRT-mediated repression. Genes. Dev. 2000, 14, 55–66. [Google Scholar] [CrossRef]
  127. Zhou, X.; Richon, V.M.; Rifkind, R.A.; Marks, P.A. Identification of a transcriptional repressor related to the noncatalytic domain of histone deacetylases 4 and 5. Proc. Natl. Acad. Sci. USA 2000, 97, 1056–1061. [Google Scholar] [CrossRef]
  128. Kao, H.Y.; Lee, C.H.; Komarov, A.; Han, C.C.; Evans, R.M. Isolation and characterization of mammalian HDAC10, a novel histone deacetylase. J. Biol. Chem. 2002, 277, 187–193. [Google Scholar] [CrossRef] [PubMed]
  129. Gao, L.; Cueto, M.A.; Asselbergs, F.; Atadja, P. Cloning and functional characterization of HDAC11, a novel member of the human histone deacetylase family. J. Biol. Chem. 2002, 277, 25748–25755. [Google Scholar] [CrossRef]
  130. Glozak, M.A.; Seto, E. Acetylation/deacetylation modulates the stability of DNA replication licensing factor Cdt1. J. Biol. Chem. 2009, 284, 11446–11453. [Google Scholar] [CrossRef]
  131. Brachmann, C.B.; Sherman, J.M.; Devine, S.E.; Cameron, E.E.; Pillus, L.; Boeke, J.D. The SIR2 gene family, conserved from bacteria to humans, functions in silencing, cell cycle progression, and chromosome stability. Genes. Dev. 1995, 9, 2888–2902. [Google Scholar] [CrossRef]
  132. Frye, R.A. Characterization of five human cDNAs with homology to the yeast SIR2 gene: Sir2-like proteins (sirtuins) metabolize NAD and may have protein ADP-ribosyltransferase activity. Biochem. Biophys. Res. Commun. 1999, 260, 273–279. [Google Scholar] [CrossRef]
  133. Du, J.; Zhou, Y.; Su, X.; Yu, J.J.; Khan, S.; Jiang, H.; Kim, J.; Woo, J.; Kim, J.H.; Choi, B.H.; et al. Sirt5 is a NAD-dependent protein lysine demalonylase and desuccinylase. Science 2011, 334, 806–809. [Google Scholar] [CrossRef] [PubMed]
  134. Li, Y.; Seto, E. HDACs and HDAC Inhibitors in Cancer Development and Therapy. Cold Spring Harb. Perspect. Med. 2016, 6, a026831. [Google Scholar] [CrossRef] [PubMed]
  135. Zaikos, T.D.; Painter, M.M.; Sebastian Kettinger, N.T.; Terry, V.H.; Collins, K.L. Class 1-Selective Histone Deacetylase (HDAC) Inhibitors Enhance HIV Latency Reversal while Preserving the Activity of HDAC Isoforms Necessary for Maximal HIV Gene Expression. J. Virol. 2018, 92, e02110-17. [Google Scholar] [CrossRef] [PubMed]
  136. Matalon, S.; Rasmussen, T.A.; Dinarello, C.A. Histone deacetylase inhibitors for purging HIV-1 from the latent reservoir. Mol. Med. 2011, 17, 466–472. [Google Scholar] [CrossRef] [PubMed]
  137. Ying, H.; Zhang, Y.; Lin, S.; Han, Y.; Zhu, H.Z. Histone deacetylase inhibitor Scriptaid reactivates latent HIV-1 promoter by inducing histone modification in in vitro latency cell lines. Int. J. Mol. Med. 2010, 26, 265–272. [Google Scholar] [CrossRef]
  138. Yin, H.; Zhang, Y.; Zhou, X.; Zhu, H. Histonedeacetylase inhibitor Oxamflatin increase HIV-1 transcription by inducing histone modification in latently infected cells. Mol. Biol. Rep. 2011, 38, 5071–5078. [Google Scholar] [CrossRef]
  139. Mates, J.M.; de Silva, S.; Lustberg, M.; Van Deusen, K.; Baiocchi, R.A.; Wu, L.; Kwiek, J.J. A Novel Histone Deacetylase Inhibitor, AR-42, Reactivates HIV-1 from Chronically and Latently Infected CD4(+) T-cells. Retrovirology 2015, 7, 1–5. [Google Scholar] [CrossRef]
  140. Kiernan, R.E.; Vanhulle, C.; Schiltz, L.; Adam, E.; Xiao, H.; Maudoux, F.; Calomme, C.; Burny, A.; Nakatani, Y.; Jeang, K.T.; et al. HIV-1 tat transcriptional activity is regulated by acetylation. EMBO J. 1999, 18, 6106–6118. [Google Scholar] [CrossRef]
  141. Divsalar, D.N.; Simoben, C.V.; Schonhofer, C.; Richard, K.; Sippl, W.; Ntie-Kang, F.; Tietjen, I. Novel Histone Deacetylase Inhibitors and HIV-1 Latency-Reversing Agents Identified by Large-Scale Virtual Screening. Front. Pharmacol. 2020, 11, 905. [Google Scholar] [CrossRef]
  142. Spina, C.A.; Anderson, J.; Archin, N.M.; Bosque, A.; Chan, J.; Famiglietti, M.; Greene, W.C.; Kashuba, A.; Lewin, S.R.; Margolis, D.M.; et al. An in-depth comparison of latent HIV-1 reactivation in multiple cell model systems and resting CD4+ T cells from aviremic patients. PLoS Pathog. 2013, 9, e1003834. [Google Scholar] [CrossRef]
  143. Boateng, A.T.; Abaidoo-Myles, A.; Bonney, E.Y.; Kyei, G.B. Isoform-Selective Versus Nonselective Histone Deacetylase Inhibitors in HIV Latency Reversal. AIDS Res. Hum. Retroviruses 2022, 38, 615–621. [Google Scholar] [CrossRef]
  144. Wei, D.G.; Chiang, V.; Fyne, E.; Balakrishnan, M.; Barnes, T.; Graupe, M.; Hesselgesser, J.; Irrinki, A.; Murry, J.P.; Stepan, G.; et al. Histone deacetylase inhibitor romidepsin induces HIV expression in CD4 T cells from patients on suppressive antiretroviral therapy at concentrations achieved by clinical dosing. PLoS Pathog. 2014, 10, e1004071. [Google Scholar] [CrossRef] [PubMed]
  145. Banga, R.; Procopio, F.A.; Cavassini, M.; Perreau, M. In Vitro Reactivation of Replication-Competent and Infectious HIV-1 by Histone Deacetylase Inhibitors. J. Virol. 2016, 90, 1858–1871. [Google Scholar] [CrossRef] [PubMed]
  146. Shultz, M.; Fan, J.; Chen, C.; Cho, Y.S.; Davis, N.; Bickford, S.; Buteau, K.; Cao, X.; Holmqvist, M.; Hsu, M.; et al. The design, synthesis and structure-activity relationships of novel isoindoline-based histone deacetylase inhibitors. Bioorg. Med. Chem. Lett. 2011, 21, 4909–4912. [Google Scholar] [CrossRef] [PubMed]
  147. New Drugs/Drug News. Pharm. Ther. 2014, 39, 539–578.
  148. Younes, A.; Oki, Y.; Bociek, R.G.; Kuruvilla, J.; Fanale, M.; Neelapu, S.; Copeland, A.; Buglio, D.; Galal, A.; Besterman, J.; et al. Mocetinostat for relapsed classical Hodgkin’s lymphoma: An open-label, single-arm, phase 2 trial. Lancet Oncol. 2011, 12, 1222–1228. [Google Scholar] [CrossRef]
  149. Huber, K.; Doyon, G.; Plaks, J.; Fyne, E.; Mellors, J.W.; Sluis-Cremer, N. Inhibitors of histone deacetylases: Correlation between isoform specificity and reactivation of HIV type 1 (HIV-1) from latently infected cells. J. Biol. Chem. 2011, 286, 22211–22218. [Google Scholar] [CrossRef]
  150. Lu, W.; Yang, C.; Xu, X.; Chen, C.; Hou, X.; Fang, H.; Liu, S. A novel selective histone deacetylase I inhibitor CC-4a activates latent HIV-1 through NF-kappaB pathway. Life Sci. 2021, 267, 118427. [Google Scholar] [CrossRef]
  151. Blankstein, A.; Rubinstein, E.; Ezra, E.; Lokiec, F.; Caspi, I.; Horoszowski, H. Disc space infection and vertebral osteomyelitis as a complication of percutaneous lateral discectomy. Clin. Orthop. Relat. Res. 1987, 225, 234–237. [Google Scholar] [CrossRef]
  152. Friedman, J.; Cho, W.K.; Chu, C.K.; Keedy, K.S.; Archin, N.M.; Margolis, D.M.; Karn, J. Epigenetic silencing of HIV-1 by the histone H3 lysine 27 methyltransferase enhancer of Zeste 2. J. Virol. 2011, 85, 9078–9089. [Google Scholar] [CrossRef]
  153. Maina, E.K.; Adan, A.A.; Mureithi, H.; Muriuki, J.; Lwembe, R.M. A Review of Current Strategies Towards the Elimination of Latent HIV-1 and Subsequent HIV-1 Cure. Curr. HIV Res. 2021, 19, 14–26. [Google Scholar] [CrossRef]
  154. Kumar, A.; Darcis, G.; Van Lint, C.; Herbein, G. Epigenetic control of HIV-1 post integration latency: Implications for therapy. Clin. Epigenetics 2015, 7, 103. [Google Scholar] [CrossRef]
  155. Boehm, D.; Ott, M. Host Methyltransferases and Demethylases: Potential New Epigenetic Targets for HIV Cure Strategies and Beyond. AIDS Res. Hum. Retroviruses 2017, 33, S8–S22. [Google Scholar] [CrossRef]
  156. Bouchat, S.; Gatot, J.S.; Kabeya, K.; Cardona, C.; Colin, L.; Herbein, G.; De Wit, S.; Clumeck, N.; Lambotte, O.; Rouzioux, C.; et al. Histone methyltransferase inhibitors induce HIV-1 recovery in resting CD4(+) T cells from HIV-1-infected HAART-treated patients. AIDS 2012, 26, 1473–1482. [Google Scholar] [CrossRef]
  157. Margolis, D.M.; Archin, N.M.; Cohen, M.S.; Eron, J.J.; Ferrari, G.; Garcia, J.V.; Gay, C.L.; Goonetilleke, N.; Joseph, S.B.; Swanstrom, R.; et al. Curing HIV: Seeking to Target and Clear Persistent Infection. Cell 2020, 181, 189–206. [Google Scholar] [CrossRef]
  158. Wu, S.Y.; Chiang, C.M. The double bromodomain-containing chromatin adaptor Brd4 and transcriptional regulation. J. Biol. Chem. 2007, 282, 13141–13145. [Google Scholar] [CrossRef]
  159. Zaware, N.; Zhou, M.M. Bromodomain biology and drug discovery. Nat. Struct. Mol. Biol. 2019, 26, 870–879. [Google Scholar] [CrossRef] [PubMed]
  160. Cheung, K.L.; Kim, C.; Zhou, M.M. The Functions of BET Proteins in Gene Transcription of Biology and Diseases. Front. Mol. Biosci. 2021, 8, 728777. [Google Scholar] [CrossRef] [PubMed]
  161. Wu, S.Y.; Lee, A.Y.; Lai, H.T.; Zhang, H.; Chiang, C.M. Phospho switch triggers Brd4 chromatin binding and activator recruitment for gene-specific targeting. Mol. Cell 2013, 49, 843–857. [Google Scholar] [CrossRef] [PubMed]
  162. Sims, R.J., 3rd; Belotserkovskaya, R.; Reinberg, D. Elongation by RNA polymerase II: The short and long of it. Genes Dev. 2004, 18, 2437–2468. [Google Scholar] [CrossRef]
  163. Whyte, W.A.; Orlando, D.A.; Hnisz, D.; Abraham, B.J.; Lin, C.Y.; Kagey, M.H.; Rahl, P.B.; Lee, T.I.; Young, R.A. Master transcription factors and mediator establish super-enhancers at key cell identity genes. Cell 2013, 153, 307–319. [Google Scholar] [CrossRef] [PubMed]
  164. Yang, Z.; Yik, J.H.; Chen, R.; He, N.; Jang, M.K.; Ozato, K.; Zhou, Q. Recruitment of P-TEFb for stimulation of transcriptional elongation by the bromodomain protein Brd4. Mol. Cell 2005, 19, 535–545. [Google Scholar] [CrossRef]
  165. Itzen, F.; Greifenberg, A.K.; Bosken, C.A.; Geyer, M. Brd4 activates P-TEFb for RNA polymerase II CTD phosphorylation. Nucleic Acids Res. 2014, 42, 7577–7590. [Google Scholar] [CrossRef]
  166. Bisgrove, D.A.; Mahmoudi, T.; Henklein, P.; Verdin, E. Conserved P-TEFb-interacting domain of BRD4 inhibits HIV transcription. Proc. Natl. Acad. Sci. USA 2007, 104, 13690–13695. [Google Scholar] [CrossRef]
  167. Asamitsu, K.; Fujinaga, K.; Okamoto, T. HIV Tat/P-TEFb Interaction: A Potential Target for Novel Anti-HIV Therapies. Molecules 2018, 23, 933. [Google Scholar] [CrossRef] [PubMed]
  168. Li, Z.; Guo, J.; Wu, Y.; Zhou, Q. The BET bromodomain inhibitor JQ1 activates HIV latency through antagonizing Brd4 inhibition of Tat-transactivation. Nucleic Acids Res. 2013, 41, 277–287. [Google Scholar] [CrossRef] [PubMed]
  169. Zhu, J.; Gaiha, G.D.; John, S.P.; Pertel, T.; Chin, C.R.; Gao, G.; Qu, H.; Walker, B.D.; Elledge, S.J.; Brass, A.L. Reactivation of latent HIV-1 by inhibition of BRD4. Cell Rep. 2012, 2, 807–816. [Google Scholar] [CrossRef] [PubMed]
  170. Boehm, D.; Calvanese, V.; Dar, R.D.; Xing, S.; Schroeder, S.; Martins, L.; Aull, K.; Li, P.C.; Planelles, V.; Bradner, J.E.; et al. BET bromodomain-targeting compounds reactivate HIV from latency via a Tat-independent mechanism. Cell Cycle 2013, 12, 452–462. [Google Scholar] [CrossRef]
  171. Boehm, D.; Conrad, R.J.; Ott, M. Bromodomain proteins in HIV infection. Viruses 2013, 5, 1571–1586. [Google Scholar] [CrossRef] [PubMed]
  172. Shi, J.; Vakoc, C.R. The mechanisms behind the therapeutic activity of BET bromodomain inhibition. Mol. Cell 2014, 54, 728–736. [Google Scholar] [CrossRef] [PubMed]
  173. Devaiah, B.N.; Gegonne, A.; Singer, D.S. Bromodomain 4: A cellular Swiss army knife. J. Leukoc. Biol. 2016, 100, 679–686. [Google Scholar] [CrossRef] [PubMed]
  174. Salahong, T.; Schwartz, C.; Sungthong, R. Are BET Inhibitors yet Promising Latency-Reversing Agents for HIV-1 Reactivation in AIDS Therapy? Viruses 2021, 13, 1026. [Google Scholar] [CrossRef] [PubMed]
  175. Baldan, F.; Allegri, L.; Lazarevic, M.; Catia, M.; Milosevic, M.; Damante, G.; Milasin, J. Biological and molecular effects of bromodomain and extra-terminal (BET) inhibitors JQ1, IBET-151, and IBET-762 in OSCC cells. J. Oral Pathol. Med. 2019, 48, 214–221. [Google Scholar] [CrossRef] [PubMed]
  176. Bauer, K.; Berger, D.; Zielinski, C.C.; Valent, P.; Grunt, T.W. Hitting two oncogenic machineries in cancer cells: Cooperative effects of the multi-kinase inhibitor ponatinib and the BET bromodomain blockers JQ1 or dBET1 on human carcinoma cells. Oncotarget 2018, 9, 26491–26506. [Google Scholar] [CrossRef] [PubMed]
  177. Bechter, O.; Schoffski, P. Make your best BET: The emerging role of BET inhibitor treatment in malignant tumors. Pharmacol. Ther. 2020, 208, 107479. [Google Scholar] [CrossRef] [PubMed]
  178. Chen, P.; Yang, Y.; Yang, L.; Tian, J.; Zhang, F.; Zhou, J.; Zhang, H. 3-Hydroxyisoindolin-1-one derivates: Synthesis by palladium-catalyzed CH activation as BRD4 inhibitors against human acute myeloid leukemia (AML) cells. Bioorg. Chem. 2019, 86, 119–125. [Google Scholar] [CrossRef] [PubMed]
  179. Chen, Y.; Xu, L.; Mayakonda, A.; Huang, M.L.; Kanojia, D.; Tan, T.Z.; Dakle, P.; Lin, R.Y.; Ke, X.Y.; Said, J.W.; et al. Bromodomain and extraterminal proteins foster the core transcriptional regulatory programs and confer vulnerability in liposarcoma. Nat. Commun. 2019, 10, 1353. [Google Scholar] [CrossRef]
  180. Banerjee, C.; Archin, N.; Michaels, D.; Belkina, A.C.; Denis, G.V.; Bradner, J.; Sebastiani, P.; Margolis, D.M.; Montano, M. BET bromodomain inhibition as a novel strategy for reactivation of HIV-1. J. Leukoc. Biol. 2012, 92, 1147–1154. [Google Scholar] [CrossRef]
  181. Conrad, R.J.; Fozouni, P.; Thomas, S.; Sy, H.; Zhang, Q.; Zhou, M.M.; Ott, M. The Short Isoform of BRD4 Promotes HIV-1 Latency by Engaging Repressive SWI/SNF Chromatin-Remodeling Complexes. Mol. Cell 2017, 67, 1001–1012. [Google Scholar] [CrossRef] [PubMed]
  182. Bullen, C.K.; Laird, G.M.; Durand, C.M.; Siliciano, J.D.; Siliciano, R.F. New ex vivo approaches distinguish effective and ineffective single agents for reversing HIV-1 latency in vivo. Nat. Med. 2014, 20, 425–429. [Google Scholar] [CrossRef]
  183. Li, G.; Zhang, Z.; Reszka-Blanco, N.; Li, F.; Chi, L.; Ma, J.; Jeffrey, J.; Cheng, L.; Su, L. Specific Activation In Vivo of HIV-1 by a Bromodomain Inhibitor from Monocytic Cells in Humanized Mice under Antiretroviral Therapy. J. Virol. 2019, 93, e00233-19. [Google Scholar] [CrossRef]
  184. Lu, P.; Qu, X.; Shen, Y.; Jiang, Z.; Wang, P.; Zeng, H.; Ji, H.; Deng, J.; Yang, X.; Li, X.; et al. The BET inhibitor OTX015 reactivates latent HIV-1 through P-TEFb. Sci. Rep. 2016, 6, 24100. [Google Scholar] [CrossRef]
  185. Liang, T.; Zhang, X.; Lai, F.; Lin, J.; Zhou, C.; Xu, X.; Tan, X.; Liu, S.; Li, L. A novel bromodomain inhibitor, CPI-203, serves as an HIV-1 latency-reversing agent by activating positive transcription elongation factor b. Biochem. Pharmacol. 2019, 164, 237–251. [Google Scholar] [CrossRef]
  186. Abner, E.; Stoszko, M.; Zeng, L.; Chen, H.C.; Izquierdo-Bouldstridge, A.; Konuma, T.; Zorita, E.; Fanunza, E.; Zhang, Q.; Mahmoudi, T.; et al. A New Quinoline BRD4 Inhibitor Targets a Distinct Latent HIV-1 Reservoir for Reactivation from Other “Shock” Drugs. J. Virol. 2018, 92, e02056-17. [Google Scholar] [CrossRef] [PubMed]
  187. Lu, P.; Shen, Y.; Yang, H.; Wang, Y.; Jiang, Z.; Yang, X.; Zhong, Y.; Pan, H.; Xu, J.; Lu, H.; et al. BET inhibitors RVX-208 and PFI-1 reactivate HIV-1 from latency. Sci. Rep. 2017, 7, 16646. [Google Scholar] [CrossRef] [PubMed]
  188. Filippakopoulos, P.; Qi, J.; Picaud, S.; Shen, Y.; Smith, W.B.; Fedorov, O.; Morse, E.M.; Keates, T.; Hickman, T.T.; Felletar, I.; et al. Selective inhibition of BET bromodomains. Nature 2010, 468, 1067–1073. [Google Scholar] [CrossRef] [PubMed]
  189. Tyler, D.S.; Vappiani, J.; Caneque, T.; Lam, E.Y.N.; Ward, A.; Gilan, O.; Chan, Y.C.; Hienzsch, A.; Rutkowska, A.; Werner, T.; et al. Click chemistry enables preclinical evaluation of targeted epigenetic therapies. Science 2017, 356, 1397–1401. [Google Scholar] [CrossRef]
  190. Picaud, S.; Wells, C.; Felletar, I.; Brotherton, D.; Martin, S.; Savitsky, P.; Diez-Dacal, B.; Philpott, M.; Bountra, C.; Lingard, H.; et al. RVX-208, an inhibitor of BET transcriptional regulators with selectivity for the second bromodomain. Proc. Natl. Acad. Sci. USA 2013, 110, 19754–19759. [Google Scholar] [CrossRef]
  191. Iwasaki, A.; Medzhitov, R. Toll-like receptor control of the adaptive immune responses. Nat. Immunol. 2004, 5, 987–995. [Google Scholar] [CrossRef]
  192. Janeway, C.A., Jr.; Medzhitov, R. Innate immune recognition. Annu. Rev. Immunol. 2002, 20, 197–216. [Google Scholar] [CrossRef]
  193. Medzhitov, R.; Janeway, C.A., Jr. Innate immunity: The virtues of a nonclonal system of recognition. Cell 1997, 91, 295–298. [Google Scholar] [CrossRef]
  194. Medzhitov, R.; Preston-Hurlburt, P.; Janeway, C.A., Jr. A human homologue of the Drosophila Toll protein signals activation of adaptive immunity. Nature 1997, 388, 394–397. [Google Scholar] [CrossRef]
  195. Martinsen, J.T.; Gunst, J.D.; Hojen, J.F.; Tolstrup, M.; Sogaard, O.S. The Use of Toll-Like Receptor Agonists in HIV-1 Cure Strategies. Front. Immunol. 2020, 11, 1112. [Google Scholar] [CrossRef] [PubMed]
  196. Macedo, A.B.; Novis, C.L.; Bosque, A. Targeting Cellular and Tissue HIV Reservoirs With Toll-Like Receptor Agonists. Front. Immunol. 2019, 10, 2450. [Google Scholar] [CrossRef]
  197. Liu, G.; Zhao, Y. Toll-like receptors and immune regulation: Their direct and indirect modulation on regulatory CD4+ CD25+ T cells. Immunology 2007, 122, 149–156. [Google Scholar] [CrossRef] [PubMed]
  198. Kawai, T.; Akira, S. The role of pattern-recognition receptors in innate immunity: Update on Toll-like receptors. Nat. Immunol. 2010, 11, 373–384. [Google Scholar] [CrossRef] [PubMed]
  199. Saxena, M.; Sabado, R.L.; La Mar, M.; Mohri, H.; Salazar, A.M.; Dong, H.; Correa Da Rosa, J.; Markowitz, M.; Bhardwaj, N.; Miller, E. Poly-ICLC, a TLR3 Agonist, Induces Transient Innate Immune Responses in Patients With Treated HIV-Infection: A Randomized Double-Blinded Placebo Controlled Trial. Front. Immunol. 2019, 10, 725. [Google Scholar] [CrossRef]
  200. Borducchi, E.N.; Cabral, C.; Stephenson, K.E.; Liu, J.; Abbink, P.; Ng’ang’a, D.; Nkolola, J.P.; Brinkman, A.L.; Peter, L.; Lee, B.C.; et al. Ad26/MVA therapeutic vaccination with TLR7 stimulation in SIV-infected rhesus monkeys. Nature 2016, 540, 284–287. [Google Scholar] [CrossRef]
  201. Borducchi, E.N.; Liu, J.; Nkolola, J.P.; Cadena, A.M.; Yu, W.H.; Fischinger, S.; Broge, T.; Abbink, P.; Mercado, N.B.; Chandrashekar, A.; et al. Antibody and TLR7 agonist delay viral rebound in SHIV-infected monkeys. Nature 2018, 563, 360–364. [Google Scholar] [CrossRef]
  202. Lim, S.Y.; Osuna, C.E.; Hraber, P.T.; Hesselgesser, J.; Gerold, J.M.; Barnes, T.L.; Sanisetty, S.; Seaman, M.S.; Lewis, M.G.; Geleziunas, R.; et al. TLR7 agonists induce transient viremia and reduce the viral reservoir in SIV-infected rhesus macaques on antiretroviral therapy. Sci. Transl. Med. 2018, 10, eaao4521. [Google Scholar] [CrossRef]
  203. Del Prete, G.Q.; Alvord, W.G.; Li, Y.; Deleage, C.; Nag, M.; Oswald, K.; Thomas, J.A.; Pyle, C.; Bosche, W.J.; Coalter, V.; et al. TLR7 agonist administration to SIV-infected macaques receiving early initiated cART does not induce plasma viremia. JCI Insight 2019, 4, e127717. [Google Scholar] [CrossRef] [PubMed]
  204. Bekerman, E.; Hesselgesser, J.; Carr, B.; Nagel, M.; Hung, M.; Wang, A.; Stapleton, L.; von Gegerfelt, A.; Elyard, H.A.; Lifson, J.D.; et al. PD-1 Blockade and TLR7 Activation Lack Therapeutic Benefit in Chronic Simian Immunodeficiency Virus-Infected Macaques on Antiretroviral Therapy. Antimicrob. Agents Chemother. 2019, 63, e01163-19. [Google Scholar] [CrossRef] [PubMed]
  205. Griffin, G.E.; Leung, K.; Folks, T.M.; Kunkel, S.; Nabel, G.J. Activation of HIV gene expression during monocyte differentiation by induction of NF-kappa B. Nature 1989, 339, 70–73. [Google Scholar] [CrossRef] [PubMed]
  206. Jiang, G.; Dandekar, S. Targeting NF-kappaB signaling with protein kinase C agonists as an emerging strategy for combating HIV latency. AIDS Res. Hum. Retroviruses 2015, 31, 4–12. [Google Scholar] [CrossRef]
  207. McKernan, L.N.; Momjian, D.; Kulkosky, J. Protein Kinase C: One Pathway towards the Eradication of Latent HIV-1 Reservoirs. Adv. Virol. 2012, 2012, 805347. [Google Scholar] [CrossRef] [PubMed]
  208. Hurley, J.H.; Grobler, J.A. Protein kinase C and phospholipase C: Bilayer interactions and regulation. Curr. Opin. Struct. Biol. 1997, 7, 557–565. [Google Scholar] [CrossRef] [PubMed]
  209. Goel, G.; Makkar, H.P.; Francis, G.; Becker, K. Phorbol esters: Structure, biological activity, and toxicity in animals. Int. J. Toxicol. 2007, 26, 279–288. [Google Scholar] [CrossRef]
  210. Rullas, J.; Bermejo, M.; Garcia-Perez, J.; Beltan, M.; Gonzalez, N.; Hezareh, M.; Brown, S.J.; Alcami, J. Prostratin induces HIV activation and downregulates HIV receptors in peripheral blood lymphocytes. Antivir. Ther. 2004, 9, 545–554. [Google Scholar] [CrossRef]
  211. Philip, P.A.; Rea, D.; Thavasu, P.; Carmichael, J.; Stuart, N.S.; Rockett, H.; Talbot, D.C.; Ganesan, T.; Pettit, G.R.; Balkwill, F.; et al. Phase I study of bryostatin 1: Assessment of interleukin 6 and tumor necrosis factor alpha induction in vivo. The Cancer Research Campaign Phase I Committee. J. Natl. Cancer Inst. 1993, 85, 1812–1818. [Google Scholar] [CrossRef]
  212. Jayson, G.C.; Crowther, D.; Prendiville, J.; McGown, A.T.; Scheid, C.; Stern, P.; Young, R.; Brenchley, P.; Chang, J.; Owens, S.; et al. A phase I trial of bryostatin 1 in patients with advanced malignancy using a 24 hour intravenous infusion. Br. J. Cancer 1995, 72, 461–468. [Google Scholar] [CrossRef]
  213. Mehla, R.; Bivalkar-Mehla, S.; Zhang, R.; Handy, I.; Albrecht, H.; Giri, S.; Nagarkatti, P.; Nagarkatti, M.; Chauhan, A. Bryostatin modulates latent HIV-1 infection via PKC and AMPK signaling but inhibits acute infection in a receptor independent manner. PLoS ONE 2010, 5, e11160. [Google Scholar] [CrossRef]
  214. Curreli, F.; Ahmed, S.; Victor, S.M.B.; Debnath, A.K. Identification of Combinations of Protein Kinase C Activators and Histone Deacetylase Inhibitors That Potently Reactivate Latent HIV. Viruses 2020, 12, 609. [Google Scholar] [CrossRef]
  215. Steinberg, S.F. Structural basis of protein kinase C isoform function. Physiol. Rev. 2008, 88, 1341–1378. [Google Scholar] [CrossRef] [PubMed]
  216. von Burstin, V.A.; Xiao, L.; Kazanietz, M.G. Bryostatin 1 inhibits phorbol ester-induced apoptosis in prostate cancer cells by differentially modulating protein kinase C (PKC) delta translocation and preventing PKCdelta-mediated release of tumor necrosis factor-alpha. Mol. Pharmacol. 2010, 78, 325–332. [Google Scholar] [CrossRef] [PubMed]
  217. Suman Ranjan Das, S.J. Biology of the HIV Nef protein. Indian J. Med. Res. 2005, 121, 315–332. [Google Scholar]
  218. Wolf, D.; Witte, V.; Laffert, B.; Blume, K.; Stromer, E.; Trapp, S.; d’Aloja, P.; Schurmann, A.; Baur, A.S. HIV-1 Nef associated PAK and PI3-kinases stimulate Akt-independent Bad-phosphorylation to induce anti-apoptotic signals. Nat. Med. 2001, 7, 1217–1224. [Google Scholar] [CrossRef]
  219. Deng, X.; Ruvolo, P.; Carr, B.; May, W.S., Jr. Survival function of ERK1/2 as IL-3-activated, staurosporine-resistant Bcl2 kinases. Proc. Natl. Acad. Sci. USA 2000, 97, 1578–1583. [Google Scholar] [CrossRef] [PubMed]
  220. French, A.J.; Natesampillai, S.; Krogman, A.; Correia, C.; Peterson, K.L.; Alto, A.; Chandrasekar, A.P.; Misra, A.; Li, Y.; Kaufmann, S.H.; et al. Reactivating latent HIV with PKC agonists induces resistance to apoptosis and is associated with phosphorylation and activation of BCL2. PLoS Pathog. 2020, 16, e1008906. [Google Scholar] [CrossRef]
  221. Morgan, R.J., Jr.; Leong, L.; Chow, W.; Gandara, D.; Frankel, P.; Garcia, A.; Lenz, H.J.; Doroshow, J.H. Phase II trial of bryostatin-1 in combination with cisplatin in patients with recurrent or persistent epithelial ovarian cancer: A California cancer consortium study. Investig. New Drugs 2012, 30, 723–728. [Google Scholar] [CrossRef]
  222. Jiang, G.; Maverakis, E.; Cheng, M.Y.; Elsheikh, M.M.; Deleage, C.; Mendez-Lagares, G.; Shimoda, M.; Yukl, S.A.; Hartigan-O’Connor, D.J.; Thompson, G.R., 3rd; et al. Disruption of latent HIV in vivo during the clearance of actinic keratosis by ingenol mebutate. JCI Insight 2019, 4, e126027. [Google Scholar] [CrossRef]
  223. Fujiwara, M.; Okamoto, M.; Ijichi, K.; Tokuhisa, K.; Hanasaki, Y.; Katsuura, K.; Uemura, D.; Shigeta, S.; Konno, K.; Yokota, T.; et al. Upregulation of HIV-1 replication in chronically infected cells by ingenol derivatives. Arch. Virol. 1998, 143, 2003–2010. [Google Scholar] [CrossRef]
  224. Pandelo Jose, D.; Bartholomeeusen, K.; da Cunha, R.D.; Abreu, C.M.; Glinski, J.; da Costa, T.B.; Bacchi Rabay, A.F.; Pianowski Filho, L.F.; Dudycz, L.W.; Ranga, U.; et al. Reactivation of latent HIV-1 by new semi-synthetic ingenol esters. Virology 2014, 462–463, 328–339. [Google Scholar] [CrossRef] [PubMed]
  225. Abreu, C.M.; Price, S.L.; Shirk, E.N.; Cunha, R.D.; Pianowski, L.F.; Clements, J.E.; Tanuri, A.; Gama, L. Dual role of novel ingenol derivatives from Euphorbia tirucalli in HIV replication: Inhibition of de novo infection and activation of viral LTR. PLoS ONE 2014, 9, e97257. [Google Scholar] [CrossRef]
  226. Warrilow, D.; Gardner, J.; Darnell, G.A.; Suhrbier, A.; Harrich, D. HIV type 1 inhibition by protein kinase C modulatory compounds. AIDS Res. Hum. Retroviruses 2006, 22, 854–864. [Google Scholar] [CrossRef] [PubMed]
  227. Jiang, G.; Mendes, E.A.; Kaiser, P.; Wong, D.P.; Tang, Y.; Cai, I.; Fenton, A.; Melcher, G.P.; Hildreth, J.E.; Thompson, G.R.; et al. Synergistic Reactivation of Latent HIV Expression by Ingenol-3-Angelate, PEP005, Targeted NF-kB Signaling in Combination with JQ1 Induced p-TEFb Activation. PLoS Pathog. 2015, 11, e1005066. [Google Scholar] [CrossRef] [PubMed]
  228. Gama, L.; Abreu, C.M.; Shirk, E.N.; Price, S.L.; Li, M.; Laird, G.M.; Pate, K.A.; Wietgrefe, S.W.; O’Connor, S.L.; Pianowski, L.; et al. Reactivation of simian immunodeficiency virus reservoirs in the brain of virally suppressed macaques. AIDS 2017, 31, 5–14. [Google Scholar] [CrossRef] [PubMed]
  229. Okoye, A.A.; Fromentin, R.; Takata, H.; Brehm, J.H.; Fukazawa, Y.; Randall, B.; Pardons, M.; Tai, V.; Tang, J.; Smedley, J.; et al. The ingenol-based protein kinase C agonist GSK445A is a potent inducer of HIV and SIV RNA transcription. PLoS Pathog. 2022, 18, e1010245. [Google Scholar] [CrossRef]
  230. Asada, Y.; Sukemori, A.; Watanabe, T.; Malla, K.J.; Yoshikawa, T.; Li, W.; Koike, K.; Chen, C.H.; Akiyama, T.; Qian, K.; et al. Stelleralides A-C, novel potent anti-HIV daphnane-type diterpenoids from Stellera chamaejasme L. Org. Lett. 2011, 13, 2904–2907. [Google Scholar] [CrossRef]
  231. Huang, L.; Ho, P.; Yu, J.; Zhu, L.; Lee, K.H.; Chen, C.H. Picomolar dichotomous activity of gnidimacrin against HIV-1. PLoS ONE 2011, 6, e26677. [Google Scholar] [CrossRef]
  232. Asada, Y.; Sukemori, A.; Watanabe, T.; Malla, K.J.; Yoshikawa, T.; Li, W.; Kuang, X.; Koike, K.; Chen, C.H.; Akiyama, T.; et al. Isolation, structure determination, and anti-HIV evaluation of tigliane-type diterpenes and biflavonoid from Stellera chamaejasme. J. Nat. Prod. 2013, 76, 852–857. [Google Scholar] [CrossRef]
  233. Lai, W.; Huang, L.; Zhu, L.; Ferrari, G.; Chan, C.; Li, W.; Lee, K.H.; Chen, C.H. Gnidimacrin, a Potent Anti-HIV Diterpene, Can Eliminate Latent HIV-1 Ex Vivo by Activation of Protein Kinase C beta. J. Med. Chem. 2015, 58, 8638–8646. [Google Scholar] [CrossRef] [PubMed]
  234. Xing, S.; Bullen, C.K.; Shroff, N.S.; Shan, L.; Yang, H.C.; Manucci, J.L.; Bhat, S.; Zhang, H.; Margolick, J.B.; Quinn, T.C.; et al. Disulfiram reactivates latent HIV-1 in a Bcl-2-transduced primary CD4+ T cell model without inducing global T cell activation. J. Virol. 2011, 85, 6060–6064. [Google Scholar] [CrossRef] [PubMed]
  235. Doyon, G.; Zerbato, J.; Mellors, J.W.; Sluis-Cremer, N. Disulfiram reactivates latent HIV-1 expression through depletion of the phosphatase and tensin homolog. AIDS 2013, 27, F7–F11. [Google Scholar] [CrossRef] [PubMed]
  236. Elliott, J.H.; McMahon, J.H.; Chang, C.C.; Lee, S.A.; Hartogensis, W.; Bumpus, N.; Savic, R.; Roney, J.; Hoh, R.; Solomon, A.; et al. Short-term administration of disulfiram for reversal of latent HIV infection: A phase 2 dose-escalation study. Lancet HIV 2015, 2, e520–e529. [Google Scholar] [CrossRef]
  237. Spivak, A.M.; Andrade, A.; Eisele, E.; Hoh, R.; Bacchetti, P.; Bumpus, N.N.; Emad, F.; Buckheit, R., 3rd; McCance-Katz, E.F.; Lai, J.; et al. A pilot study assessing the safety and latency-reversing activity of disulfiram in HIV-1-infected adults on antiretroviral therapy. Clin. Infect. Dis. 2014, 58, 883–890. [Google Scholar] [CrossRef] [PubMed]
  238. Laird, G.M.; Bullen, C.K.; Rosenbloom, D.I.; Martin, A.R.; Hill, A.L.; Durand, C.M.; Siliciano, J.D.; Siliciano, R.F. Ex vivo analysis identifies effective HIV-1 latency-reversing drug combinations. J. Clin. Investig. 2015, 125, 1901–1912. [Google Scholar] [CrossRef] [PubMed]
  239. Pache, L.; Dutra, M.S.; Spivak, A.M.; Marlett, J.M.; Murry, J.P.; Hwang, Y.; Maestre, A.M.; Manganaro, L.; Vamos, M.; Teriete, P.; et al. BIRC2/cIAP1 Is a Negative Regulator of HIV-1 Transcription and Can Be Targeted by Smac Mimetics to Promote Reversal of Viral Latency. Cell Host Microbe 2015, 18, 345–353. [Google Scholar] [CrossRef]
  240. Sun, S.C. The noncanonical NF-kappaB pathway. Immunol. Rev. 2012, 246, 125–140. [Google Scholar] [CrossRef]
  241. Nixon, C.C.; Mavigner, M.; Sampey, G.C.; Brooks, A.D.; Spagnuolo, R.A.; Irlbeck, D.M.; Mattingly, C.; Ho, P.T.; Schoof, N.; Cammon, C.G.; et al. Systemic HIV and SIV latency reversal via non-canonical NF-kappaB signalling in vivo. Nature 2020, 578, 160–165. [Google Scholar] [CrossRef]
  242. Pache, L.; Marsden, M.D.; Teriete, P.; Portillo, A.J.; Heimann, D.; Kim, J.T.; Soliman, M.S.A.; Dimapasoc, M.; Carmona, C.; Celeridad, M.; et al. Pharmacological Activation of Non-canonical NF-kappaB Signaling Activates Latent HIV-1 Reservoirs In Vivo. Cell Rep. Med. 2020, 1, 100037. [Google Scholar] [CrossRef]
  243. Berro, R.; de la Fuente, C.; Klase, Z.; Kehn, K.; Parvin, L.; Pumfery, A.; Agbottah, E.; Vertes, A.; Nekhai, S.; Kashanchi, F. Identifying the membrane proteome of HIV-1 latently infected cells. J. Biol. Chem. 2007, 282, 8207–8218. [Google Scholar] [CrossRef] [PubMed]
  244. Eckelman, B.P.; Salvesen, G.S. The human anti-apoptotic proteins cIAP1 and cIAP2 bind but do not inhibit caspases. J. Biol. Chem. 2006, 281, 3254–3260. [Google Scholar] [CrossRef] [PubMed]
  245. Bobardt, M.; Kuo, J.; Chatterji, U.; Chanda, S.; Little, S.J.; Wiedemann, N.; Vuagniaux, G.; Gallay, P.A. The inhibitor apoptosis protein antagonist Debio 1143 Is an attractive HIV-1 latency reversal candidate. PLoS ONE 2019, 14, e0211746. [Google Scholar] [CrossRef] [PubMed]
  246. Dashti, A.; Sukkestad, S.; Horner, A.M.; Neja, M.; Siddiqi, Z.; Waller, C.; Goldy, J.; Monroe, D.; Lin, A.; Schoof, N.; et al. AZD5582 plus SIV-specific antibodies reduce lymph node viral reservoirs in antiretroviral therapy-suppressed macaques. Nat. Med. 2023, 29, 2535–2546. [Google Scholar] [CrossRef]
  247. Lewin, S.R.; Attoye, T.; Bansbach, C.; Doehle, B.; Dube, K.; Dybul, M.; SenGupta, D.; Jiang, A.; Johnston, R.; Lamplough, R.; et al. Multi-stakeholder consensus on a target product profile for an HIV cure. Lancet HIV 2021, 8, e42–e50. [Google Scholar] [CrossRef] [PubMed]
  248. Martinez-Bonet, M.; Clemente, M.I.; Serramia, M.J.; Munoz, E.; Moreno, S.; Munoz-Fernandez, M.A. Synergistic Activation of Latent HIV-1 Expression by Novel Histone Deacetylase Inhibitors and Bryostatin-1. Sci. Rep. 2015, 5, 16445. [Google Scholar] [CrossRef] [PubMed]
  249. Reuse, S.; Calao, M.; Kabeya, K.; Guiguen, A.; Gatot, J.S.; Quivy, V.; Vanhulle, C.; Lamine, A.; Vaira, D.; Demonte, D.; et al. Synergistic activation of HIV-1 expression by deacetylase inhibitors and prostratin: Implications for treatment of latent infection. PLoS ONE 2009, 4, e6093. [Google Scholar] [CrossRef]
  250. Pardons, M.; Fromentin, R.; Pagliuzza, A.; Routy, J.P.; Chomont, N. Latency-Reversing Agents Induce Differential Responses in Distinct Memory CD4 T Cell Subsets in Individuals on Antiretroviral Therapy. Cell Rep. 2019, 29, 2783–2795. [Google Scholar] [CrossRef]
  251. Darcis, G.; Kula, A.; Bouchat, S.; Fujinaga, K.; Corazza, F.; Ait-Ammar, A.; Delacourt, N.; Melard, A.; Kabeya, K.; Vanhulle, C.; et al. An In-Depth Comparison of Latency-Reversing Agent Combinations in Various In Vitro and Ex Vivo HIV-1 Latency Models Identified Bryostatin-1+JQ1 and Ingenol-B+JQ1 to Potently Reactivate Viral Gene Expression. PLoS Pathog. 2015, 11, e1005063. [Google Scholar] [CrossRef]
  252. Matsuda, K.; Kobayakawa, T.; Kariya, R.; Tsuchiya, K.; Ryu, S.; Tsuji, K.; Ishii, T.; Gatanaga, H.; Yoshimura, K.; Okada, S.; et al. A Therapeutic Strategy to Combat HIV-1 Latently Infected Cells With a Combination of Latency-Reversing Agents Containing DAG-Lactone PKC Activators. Front. Microbiol. 2021, 12, 636276. [Google Scholar] [CrossRef]
  253. Rodari, A.; Darcis, G.; Van Lint, C.M. The Current Status of Latency Reversing Agents for HIV-1 Remission. Annu. Rev. Virol. 2021, 8, 491–514. [Google Scholar] [CrossRef]
  254. Huang, H.; Liu, S.; Jean, M.; Simpson, S.; Huang, H.; Merkley, M.; Hayashi, T.; Kong, W.; Rodriguez-Sanchez, I.; Zhang, X.; et al. A Novel Bromodomain Inhibitor Reverses HIV-1 Latency through Specific Binding with BRD4 to Promote Tat and P-TEFb Association. Front. Microbiol. 2017, 8, 1035. [Google Scholar] [CrossRef] [PubMed]
  255. Zhang, X.X.; Lin, J.; Liang, T.Z.; Duan, H.; Tan, X.H.; Xi, B.M.; Li, L.; Liu, S.W. The BET bromodomain inhibitor apabetalone induces apoptosis of latent HIV-1 reservoir cells following viral reactivation. Acta Pharmacol. Sin. 2019, 40, 98–110. [Google Scholar] [CrossRef] [PubMed]
  256. Washizaki, A.; Murata, M.; Seki, Y.; Kikumori, M.; Tang, Y.; Tan, W.; Wardani, N.P.; Irie, K.; Akari, H. The Novel PKC Activator 10-Methyl-Aplog-1 Combined with JQ1 Induced Strong and Synergistic HIV Reactivation with Tolerable Global T Cell Activation. Viruses 2021, 13, 2037. [Google Scholar] [CrossRef] [PubMed]
  257. Shibuya, Y.; Kudo, K.; Zeligs, K.P.; Anderson, D.; Hernandez, L.; Ning, F.; Cole, C.B.; Fergusson, M.; Kedei, N.; Lyons, J.; et al. SMAC Mimetics Synergistically Cooperate with HDAC Inhibitors Enhancing TNF-alpha Autocrine Signaling. Cancers 2023, 15, 1315. [Google Scholar] [CrossRef] [PubMed]
  258. Falcinelli, S.D.; Peterson, J.J.; Turner, A.W.; Irlbeck, D.; Read, J.; Raines, S.L.; James, K.S.; Sutton, C.; Sanchez, A.; Emery, A.; et al. Combined noncanonical NF-kappaB agonism and targeted BET bromodomain inhibition reverse HIV latency ex vivo. J. Clin. Investig. 2022, 132, e157281. [Google Scholar] [CrossRef]
  259. Hattori, S.I.; Matsuda, K.; Tsuchiya, K.; Gatanaga, H.; Oka, S.; Yoshimura, K.; Mitsuya, H.; Maeda, K. Combination of a Latency-Reversing Agent With a Smac Mimetic Minimizes Secondary HIV-1 Infection in vitro. Front. Microbiol. 2018, 9, 2022. [Google Scholar] [CrossRef] [PubMed]
  260. Molyer, B.; Kumar, A.; Angel, J.B. SMAC Mimetics as Therapeutic Agents in HIV Infection. Front. Immunol. 2021, 12, 780400. [Google Scholar] [CrossRef]
  261. Howard, J.N.; Bosque, A. IL-15 and N-803 for HIV Cure Approaches. Viruses 2023, 15, 1912. [Google Scholar] [CrossRef]
  262. Miller, J.S.; Davis, Z.B.; Helgeson, E.; Reilly, C.; Thorkelson, A.; Anderson, J.; Lima, N.S.; Jorstad, S.; Hart, G.T.; Lee, J.H.; et al. Safety and virologic impact of the IL-15 superagonist N-803 in people living with HIV: A phase 1 trial. Nat. Med. 2022, 28, 392–400. [Google Scholar] [CrossRef]
  263. Copertino, D.C., Jr.; Holmberg, C.S.; Weiler, J.; Ward, A.R.; Howard, J.N.; Levinger, C.; Pang, A.P.; Corley, M.J.; Dundar, F.; Zumbo, P.; et al. The latency-reversing agent HODHBt synergizes with IL-15 to enhance cytotoxic function of HIV-specific T cells. JCI Insight 2023, 8, e169028. [Google Scholar] [CrossRef] [PubMed]
  264. Bosque, A.; Nilson, K.A.; Macedo, A.B.; Spivak, A.M.; Archin, N.M.; Van Wagoner, R.M.; Martins, L.J.; Novis, C.L.; Szaniawski, M.A.; Ireland, C.M.; et al. Benzotriazoles Reactivate Latent HIV-1 through Inactivation of STAT5 SUMOylation. Cell Rep. 2017, 18, 1324–1334. [Google Scholar] [CrossRef] [PubMed]
  265. Macedo, A.B.; Levinger, C.; Nguyen, B.N.; Richard, J.; Gupta, M.; Cruz, C.R.Y.; Finzi, A.; Chiappinelli, K.B.; Crandall, K.A.; Bosque, A. The HIV Latency Reversal Agent HODHBt Enhances NK Cell Effector and Memory-Like Functions by Increasing Interleukin-15-Mediated STAT Activation. J. Virol. 2022, 96, e0037222. [Google Scholar] [CrossRef] [PubMed]
  266. Sorensen, E.S.; Macedo, A.B.; Resop, R.S.; Howard, J.N.; Nell, R.; Sarabia, I.; Newman, D.; Ren, Y.; Jones, R.B.; Planelles, V.; et al. Structure-Activity Relationship Analysis of Benzotriazine Analogues as HIV-1 Latency-Reversing Agents. Antimicrob. Agents Chemother. 2020, 64, e00888-20. [Google Scholar] [CrossRef]
  267. Soliman, S.H.A.; Cisneros, W.J.; Iwanaszko, M.; Aoi, Y.; Ganesan, S.; Walter, M.; Zeidner, J.M.; Mishra, R.K.; Kim, E.Y.; Wolinsky, S.M.; et al. Enhancing HIV-1 latency reversal through regulating the elongating RNA Pol II pause-release by a small-molecule disruptor of PAF1C. Sci. Adv. 2023, 9, eadf2468. [Google Scholar] [CrossRef] [PubMed]
  268. Clutton, G.T.; Jones, R.B. Diverse Impacts of HIV Latency-Reversing Agents on CD8+ T-Cell Function: Implications for HIV Cure. Front. Immunol. 2018, 9, 1452. [Google Scholar] [CrossRef]
  269. Kim, Y.; Anderson, J.L.; Lewin, S.R. Getting the “Kill” into “Shock and Kill”: Strategies to Eliminate Latent HIV. Cell Host Microbe 2018, 23, 14–26. [Google Scholar] [CrossRef]
  270. Ward, A.R.; Mota, T.M.; Jones, R.B. Immunological approaches to HIV cure. Semin. Immunol. 2021, 51, 101412. [Google Scholar] [CrossRef]
  271. Gupta, R.K.; Abdul-Jawad, S.; McCoy, L.E.; Mok, H.P.; Peppa, D.; Salgado, M.; Martinez-Picado, J.; Nijhuis, M.; Wensing, A.M.J.; Lee, H.; et al. HIV-1 remission following CCR5Delta32/Delta32 haematopoietic stem-cell transplantation. Nature 2019, 568, 244–248. [Google Scholar] [CrossRef]
  272. Hutter, G.; Nowak, D.; Mossner, M.; Ganepola, S.; Mussig, A.; Allers, K.; Schneider, T.; Hofmann, J.; Kucherer, C.; Blau, O.; et al. Long-term control of HIV by CCR5 Delta32/Delta32 stem-cell transplantation. N. Engl. J. Med. 2009, 360, 692–698. [Google Scholar] [CrossRef]
  273. Wang, C.X.; Cannon, P.M. The clinical applications of genome editing in HIV. Blood 2016, 127, 2546–2552. [Google Scholar] [CrossRef] [PubMed]
  274. Tebas, P.; Stein, D.; Tang, W.W.; Frank, I.; Wang, S.Q.; Lee, G.; Spratt, S.K.; Surosky, R.T.; Giedlin, M.A.; Nichol, G.; et al. Gene editing of CCR5 in autologous CD4 T cells of persons infected with HIV. N. Engl. J. Med. 2014, 370, 901–910. [Google Scholar] [CrossRef] [PubMed]
  275. Tebas, P.; Jadlowsky, J.K.; Shaw, P.A.; Tian, L.; Esparza, E.; Brennan, A.L.; Kim, S.; Naing, S.Y.; Richardson, M.W.; Vogel, A.N.; et al. CCR5-edited CD4+ T cells augment HIV-specific immunity to enable post-rebound control of HIV replication. J. Clin. Investig. 2021, 131. [Google Scholar] [CrossRef]
  276. Zeidan, J.; Sharma, A.A.; Lee, G.; Raad, A.; Fromentin, R.; Fourati, S.; Ghneim, K.; Sanchez, G.P.; Benne, C.; Canderan, G.; et al. Infusion of CCR5 Gene-Edited T Cells Allows Immune Reconstitution, HIV Reservoir Decay, and Long-Term Virological Control. bioRxiv 2021. [Google Scholar] [CrossRef]
  277. Ellwanger, J.H.; Kulmann-Leal, B.; Kaminski, V.L.; Rodrigues, A.G.; Bragatte, M.A.S.; Chies, J.A.B. Beyond HIV infection: Neglected and varied impacts of CCR5 and CCR5Delta32 on viral diseases. Virus Res. 2020, 286, 198040. [Google Scholar] [CrossRef]
  278. Gross, G.; Waks, T.; Eshhar, Z. Expression of immunoglobulin-T-cell receptor chimeric molecules as functional receptors with antibody-type specificity. Proc. Natl. Acad. Sci. USA 1989, 86, 10024–10028. [Google Scholar] [CrossRef]
  279. Mitsuyasu, R.T.; Anton, P.A.; Deeks, S.G.; Scadden, D.T.; Connick, E.; Downs, M.T.; Bakker, A.; Roberts, M.R.; June, C.H.; Jalali, S.; et al. Prolonged survival and tissue trafficking following adoptive transfer of CD4zeta gene-modified autologous CD4(+) and CD8(+) T cells in human immunodeficiency virus-infected subjects. Blood 2000, 96, 785–793. [Google Scholar] [CrossRef] [PubMed]
  280. Seif, M.; Einsele, H.; Loffler, J. CAR T Cells Beyond Cancer: Hope for Immunomodulatory Therapy of Infectious Diseases. Front. Immunol. 2019, 10, 2711. [Google Scholar] [CrossRef]
  281. Chmielewski, M.; Hombach, A.A.; Abken, H. Of CARs and TRUCKs: Chimeric antigen receptor (CAR) T cells engineered with an inducible cytokine to modulate the tumor stroma. Immunol. Rev. 2014, 257, 83–90. [Google Scholar] [CrossRef] [PubMed]
  282. Scholler, J.; Brady, T.L.; Binder-Scholl, G.; Hwang, W.T.; Plesa, G.; Hege, K.M.; Vogel, A.N.; Kalos, M.; Riley, J.L.; Deeks, S.G.; et al. Decade-long safety and function of retroviral-modified chimeric antigen receptor T cells. Sci. Transl. Med. 2012, 4, 132ra53. [Google Scholar] [CrossRef]
  283. Carvalho, T. First two patients receive CAR T cell therapy for HIV. Nat. Med. 2023, 29, 1290–1291. [Google Scholar] [CrossRef]
  284. Campos-Gonzalez, G.; Martinez-Picado, J.; Velasco-Hernandez, T.; Salgado, M. Opportunities for CAR-T Cell Immunotherapy in HIV Cure. Viruses 2023, 15, 789. [Google Scholar] [CrossRef] [PubMed]
  285. Nordstrom, J.L.; Ferrari, G.; Margolis, D.M. Bispecific antibody-derived molecules to target persistent HIV infection. J. Virus Erad. 2022, 8, 100083. [Google Scholar] [CrossRef] [PubMed]
  286. Evans, V.A.; van der Sluis, R.M.; Solomon, A.; Dantanarayana, A.; McNeil, C.; Garsia, R.; Palmer, S.; Fromentin, R.; Chomont, N.; Sekaly, R.P.; et al. Programmed cell death-1 contributes to the establishment and maintenance of HIV-1 latency. AIDS 2018, 32, 1491–1497. [Google Scholar] [CrossRef]
  287. Uldrick, T.S.; Adams, S.V.; Fromentin, R.; Roche, M.; Fling, S.P.; Goncalves, P.H.; Lurain, K.; Ramaswami, R.; Wang, C.J.; Gorelick, R.J.; et al. Pembrolizumab induces HIV latency reversal in people living with HIV and cancer on antiretroviral therapy. Sci. Transl. Med. 2022, 14, eabl3836. [Google Scholar] [CrossRef]
  288. Guihot, A.; Marcelin, A.G.; Massiani, M.A.; Samri, A.; Soulie, C.; Autran, B.; Spano, J.P. Drastic decrease of the HIV reservoir in a patient treated with nivolumab for lung cancer. Ann. Oncol. 2018, 29, 517–518. [Google Scholar] [CrossRef] [PubMed]
  289. Rasmussen, T.A.; Rajdev, L.; Rhodes, A.; Dantanarayana, A.; Tennakoon, S.; Chea, S.; Spelman, T.; Lensing, S.; Rutishauser, R.; Bakkour, S.; et al. Impact of Anti-PD-1 and Anti-CTLA-4 on the Human Immunodeficiency Virus (HIV) Reservoir in People Living With HIV With Cancer on Antiretroviral Therapy: The AIDS Malignancy Consortium 095 Study. Clin. Infect. Dis. 2021, 73, e1973–e1981. [Google Scholar] [CrossRef]
  290. Rust, B.J.; Kean, L.S.; Colonna, L.; Brandenstein, K.E.; Poole, N.H.; Obenza, W.; Enstrom, M.R.; Maldini, C.R.; Ellis, G.I.; Fennessey, C.M.; et al. Robust expansion of HIV CAR T cells following antigen boosting in ART-suppressed nonhuman primates. Blood 2020, 136, 1722–1734. [Google Scholar] [CrossRef]
  291. Wykes, M.N.; Lewin, S.R. Immune checkpoint blockade in infectious diseases. Nat. Rev. Immunol. 2018, 18, 91–104. [Google Scholar] [CrossRef]
  292. Kulpa, D.A.; Talla, A.; Brehm, J.H.; Ribeiro, S.P.; Yuan, S.; Bebin-Blackwell, A.G.; Miller, M.; Barnard, R.; Deeks, S.G.; Hazuda, D.; et al. Differentiation into an Effector Memory Phenotype Potentiates HIV-1 Latency Reversal in CD4(+) T Cells. J. Virol. 2019, 93, e00969-19. [Google Scholar] [CrossRef] [PubMed]
  293. Samer, S.; Thomas, Y.; Arainga, M.; Carter, C.; Shirreff, L.M.; Arif, M.S.; Avita, J.M.; Frank, I.; McRaven, M.D.; Thuruthiyil, C.T.; et al. Blockade of TGF-beta signaling reactivates HIV-1/SIV reservoirs and immune responses in vivo. JCI Insight 2022, 7, e162290. [Google Scholar] [CrossRef]
  294. Kime, J.; Bose, D.; Arainga, M.; Haque, M.R.; Fennessey, C.M.; Caddell, R.A.; Thomas, Y.; Ferrell, D.E.; Ali, S.; Grody, E.; et al. TGF-β blockade drives a transitional effector phenotype in T cells reversing SIV latency and decreasing SIV reservoirs in vivo. bioRxiv 2023. [Google Scholar] [CrossRef]
  295. Malim, M.H.; Emerman, M. HIV-1 accessory proteins--ensuring viral survival in a hostile environment. Cell Host Microbe 2008, 3, 388–398. [Google Scholar] [CrossRef] [PubMed]
  296. Diehl, N.; Schaal, H. Make yourself at home: Viral hijacking of the PI3K/Akt signaling pathway. Viruses 2013, 5, 3192–3212. [Google Scholar] [CrossRef] [PubMed]
  297. Juarez-Salcedo, L.M.; Desai, V.; Dalia, S. Venetoclax: Evidence to date and clinical potential. Drugs Context 2019, 8, 212574. [Google Scholar] [CrossRef]
  298. de Vos, S.; Leonard, J.P.; Friedberg, J.W.; Zain, J.; Dunleavy, K.; Humerickhouse, R.; Hayslip, J.; Pesko, J.; Wilson, W.H. Safety and efficacy of navitoclax, a BCL-2 and BCL-X(L) inhibitor, in patients with relapsed or refractory lymphoid malignancies: Results from a phase 2a study. Leuk. Lymphoma 2021, 62, 810–818. [Google Scholar] [CrossRef]
  299. Marconi, V.C.; Moser, C.; Gavegnano, C.; Deeks, S.G.; Lederman, M.M.; Overton, E.T.; Tsibris, A.; Hunt, P.W.; Kantor, A.; Sekaly, R.P.; et al. Randomized Trial of Ruxolitinib in Antiretroviral-Treated Adults With Human Immunodeficiency Virus. Clin. Infect. Dis. 2022, 74, 95–104. [Google Scholar] [CrossRef] [PubMed]
  300. Cummins, N.W.; Sainski, A.M.; Dai, H.; Natesampillai, S.; Pang, Y.P.; Bren, G.D.; de Araujo Correia, M.C.M.; Sampath, R.; Rizza, S.A.; O’Brien, D.; et al. Prime, Shock, and Kill: Priming CD4 T Cells from HIV Patients with a BCL-2 Antagonist before HIV Reactivation Reduces HIV Reservoir Size. J. Virol. 2016, 90, 4032–4048. [Google Scholar] [CrossRef]
  301. Gavegnano, C.; Brehm, J.H.; Dupuy, F.P.; Talla, A.; Ribeiro, S.P.; Kulpa, D.A.; Cameron, C.; Santos, S.; Hurwitz, S.J.; Marconi, V.C.; et al. Novel mechanisms to inhibit HIV reservoir seeding using Jak inhibitors. PLoS Pathog. 2017, 13, e1006740. [Google Scholar] [CrossRef]
  302. Li, P.; Kaiser, P.; Lampiris, H.W.; Kim, P.; Yukl, S.A.; Havlir, D.V.; Greene, W.C.; Wong, J.K. Stimulating the RIG-I pathway to kill cells in the latent HIV reservoir following viral reactivation. Nat. Med. 2016, 22, 807–811. [Google Scholar] [CrossRef]
  303. Garcia-Vidal, E.; Castellvi, M.; Pujantell, M.; Badia, R.; Jou, A.; Gomez, L.; Puig, T.; Clotet, B.; Ballana, E.; Riveira-Munoz, E.; et al. Evaluation of the Innate Immune Modulator Acitretin as a Strategy To Clear the HIV Reservoir. Antimicrob. Agents Chemother. 2017, 61, e01368-17. [Google Scholar] [CrossRef] [PubMed]
  304. 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. [Google Scholar] [CrossRef] [PubMed]
  305. Hussein, M.; Molina, M.A.; Berkhout, B.; Herrera-Carrillo, E. A CRISPR-Cas Cure for HIV/AIDS. Int. J. Mol. Sci. 2023, 24, 1563. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Cellular pathways targeted by LRAs. (A) Epigenetic modifiers: HDACi, which prevents histone deacetylation, and HMTi, which disrupts the deposition of repressive methylation marks, promote a relaxed chromatin structure, enhancing the accessibility of the viral promoter. BET proteins also impact the regulation of HIV-1 transcription. BRD4 competes with Tat for P-TEFb binding, thereby limiting Tat-mediated transcriptional activation. By displacing BRD4 from chromatin, BETi frees P-TEFb and enables its binding of Tat. (B) Pattern recognition receptor signaling: The stimulation of endosomal or cell surface TLRs leads to the activation of transcription factors NF-κB, AP-1, NFAT, and/or IRF family members. Since many TLRs are not expressed at significant levels in CD4+ T cells, TLR agonists identified as LRAs act mostly indirectly by inducing type I interferon production in plasmacytoid dendritic cells, which causes the downstream activation of CD4+ T cells and induces HIV-1 latency reversal. (C) Canonical NF-κB signaling: The recruitment of TAK1 by a range of transmembrane receptors, including TNFR, leads to activation of the IKK complex, the degradation of IκBα, and the translocation of the transcriptions factors RelA and p50 to the nucleus. This pathway can be activated by PKCa or by disulfiram to promote HIV-1 transcription. (D) Non-canonical NF-κB signaling is activated by a different set of receptors than the canonical pathway, including CD40 and LTβR. Non-canonical NF-κB signaling can be induced by Smac mimetics that antagonize cIAP proteins, leading to an accumulation of NIK, the cleavage of p100, and the translocation of the transcription factors RelB and p52 to the nucleus. Illustration created with BioRender.com.
Figure 1. Cellular pathways targeted by LRAs. (A) Epigenetic modifiers: HDACi, which prevents histone deacetylation, and HMTi, which disrupts the deposition of repressive methylation marks, promote a relaxed chromatin structure, enhancing the accessibility of the viral promoter. BET proteins also impact the regulation of HIV-1 transcription. BRD4 competes with Tat for P-TEFb binding, thereby limiting Tat-mediated transcriptional activation. By displacing BRD4 from chromatin, BETi frees P-TEFb and enables its binding of Tat. (B) Pattern recognition receptor signaling: The stimulation of endosomal or cell surface TLRs leads to the activation of transcription factors NF-κB, AP-1, NFAT, and/or IRF family members. Since many TLRs are not expressed at significant levels in CD4+ T cells, TLR agonists identified as LRAs act mostly indirectly by inducing type I interferon production in plasmacytoid dendritic cells, which causes the downstream activation of CD4+ T cells and induces HIV-1 latency reversal. (C) Canonical NF-κB signaling: The recruitment of TAK1 by a range of transmembrane receptors, including TNFR, leads to activation of the IKK complex, the degradation of IκBα, and the translocation of the transcriptions factors RelA and p50 to the nucleus. This pathway can be activated by PKCa or by disulfiram to promote HIV-1 transcription. (D) Non-canonical NF-κB signaling is activated by a different set of receptors than the canonical pathway, including CD40 and LTβR. Non-canonical NF-κB signaling can be induced by Smac mimetics that antagonize cIAP proteins, leading to an accumulation of NIK, the cleavage of p100, and the translocation of the transcription factors RelB and p52 to the nucleus. Illustration created with BioRender.com.
Viruses 15 02435 g001
Table 1. Overview of latency-promoting agents (LPA) and latency-reversing agents (LRA).
Table 1. Overview of latency-promoting agents (LPA) and latency-reversing agents (LRA).
ApproachClassMechanismExamplesSection
Latency-Promoting AgentsRNA-induced silencingsi/shRNARNA-induced silencingPromA; LTR-362; S4-siRNASection 3.1
lncRNANEAT1; NRON; PVT1; NKILA; AK130181Section 3.1
Tat inhibitiontrans-dominant Tat mutant Inhibition of Tat functionNullbasicSection 3.2
Tat inhibitorDidehydro-cortistatin A (dCA) Section 3.2
Latency-Reversing AgentsEpigenetic modifierslncRNARNA-induced gene expressionHEAL; MALAT1Section 4.1.1
Histone deacetylase inhibitors (HDACi)Inhibition of histone deacetylasesValproic acid; Vorinostat (SAHA); Panobinostat; Romidepsin; Givinosat; Belinostat; Entinostat; CC-4aSection 4.1.2
Histone methyltransferases inhibitors (HMTi)Modification of histone methylationChaetocin; DZNep; BIX-01294Section 4.1.3
Bromodomain and Extra-Terminal Domain Inhibitors (BETi)P-TEFb releaseJQ1; I-BET151; OTX015; CPI-203; MMQO; RVX-208Section 4.1.4
PAF1C inhibitorRelease of promoter-proximal paused RNA Pol II iPAF1CSection 4.3
Activators/inhibitors of Inducible Host FactorsTLR agonistsActivation of the NF-κB, NFAT, and AP-1 pathwaysPoly-ICLC; MGN1703; GS-986; GS-9620 (Vesatolimod) Section 4.2.1
PKC agonistsCanonical NF-κB activationPMA; Prostratin; Bryostratin; IngenolSection 4.2.2
PTEN inhibitorDisulfiramSection 4.2.2
Smac mimetic/ IAP antagonistNon-canonical NF-κB activationSBI-0637142; AZD5582; Ciapavir; Debio1143 (Xevinapant)Section 4.2.3
IL-15 stimulationJAK/STAT activationIL-15; N-803Section 4.3
BenzotriazolesSTAT5 activationHODHBtSection 4.3
TGF-β inhibitorsTGF-β inhibitionGalunisertib Section 4.4
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

Duggan, N.N.; Dragic, T.; Chanda, S.K.; Pache, L. Breaking the Silence: Regulation of HIV Transcription and Latency on the Road to a Cure. Viruses 2023, 15, 2435. https://doi.org/10.3390/v15122435

AMA Style

Duggan NN, Dragic T, Chanda SK, Pache L. Breaking the Silence: Regulation of HIV Transcription and Latency on the Road to a Cure. Viruses. 2023; 15(12):2435. https://doi.org/10.3390/v15122435

Chicago/Turabian Style

Duggan, Natasha N., Tatjana Dragic, Sumit K. Chanda, and Lars Pache. 2023. "Breaking the Silence: Regulation of HIV Transcription and Latency on the Road to a Cure" Viruses 15, no. 12: 2435. https://doi.org/10.3390/v15122435

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