Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                
Next Article in Journal
Description of Sheep Pox Outbreak in Spain in 2022–2023: Challenges Found and Lessons Learnt in Relation with Control and Eradication of This Disease
Previous Article in Journal
Development and Optimization of Oligonucleotide Ligation Assay (OLA) Probes for Detection of HIV-1 Resistance to Dolutegravir
Previous Article in Special Issue
Quantitative Prediction of Human Immunodeficiency Virus Drug Resistance
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

HIV Persistence, Latency, and Cure Approaches: Where Are We Now?

by
Tessa C. Chou
1,
Nishad S. Maggirwar
1 and
Matthew D. Marsden
1,2,*
1
Department of Microbiology and Molecular Genetics, School of Medicine, University of California, Irvine, CA 92617, USA
2
Department of Medicine, Division of Infectious Disease, School of Medicine, University of California, Irvine, CA 92617, USA
*
Author to whom correspondence should be addressed.
Viruses 2024, 16(7), 1163; https://doi.org/10.3390/v16071163
Submission received: 25 June 2024 / Revised: 13 July 2024 / Accepted: 17 July 2024 / Published: 19 July 2024
(This article belongs to the Special Issue HIV Reservoirs, Latency, and the Factors Responsible)

Abstract

:
The latent reservoir remains a major roadblock to curing human immunodeficiency virus (HIV) infection. Currently available antiretroviral therapy (ART) can suppress active HIV replication, reduce viral loads to undetectable levels, and halt disease progression. However, antiretroviral drugs are unable to target cells that are latently infected with HIV, which can seed viral rebound if ART is stopped. Consequently, a major focus of the field is to study the latent viral reservoir and develop safe and effective methods to eliminate it. Here, we provide an overview of the major mechanisms governing the establishment and maintenance of HIV latency, the key challenges posed by latent reservoirs, small animal models utilized to study HIV latency, and contemporary cure approaches. We also discuss ongoing efforts to apply these approaches in combination, with the goal of achieving a safe, effective, and scalable cure for HIV that can be extended to the tens of millions of people with HIV worldwide.

1. Background

Human immunodeficiency virus (HIV) remains a leading global health concern. In 2022, 39 million people across the world were living with HIV, 1.3 million people were newly infected that year, and over 40 million people had died of acquired immunodeficiency syndrome (AIDS) since the start of the HIV epidemic [1]. Combination antiretroviral therapy (ART) medications are used to prevent HIV acquisition and treat HIV infection [2]. These medications target different steps of the HIV life cycle to prevent virus replication. An impressive array of different antiretroviral drug classes have advanced to clinical use for HIV treatment, including pre-attachment, post-attachment, and fusion inhibitors, capsid inhibitors, non-nucleoside reverse-transcriptase inhibitors, nucleoside and nucleotide reverse-transcriptase inhibitors, integrase strand transfer inhibitors, and HIV protease inhibitors (Figure 1). During optimal treatment, several different antiretroviral drugs (typically three) belonging to at least two different classes are provided in combination to effectively suppress virus replication and reduce the probability of drug-resistant virus emerging in the treated individual [2]. Ideally, this treatment will reduce plasma viral loads to undetectable levels and prevent HIV infection from advancing to AIDS. ART is therefore critical in extending the lifespan of people with HIV (PWH) and has saved millions of lives. However, ART is not a cure for HIV infection. Though it successfully inhibits the replicating virus, ART is unable to target stable reservoirs of HIV, including within latently infected CD4+ T cells [3,4,5,6], which persist in the host despite long-term ART.
Replication-competent latent HIV reservoirs are formed when an infected cell harbors an intact, integrated HIV provirus that is expressing little or no RNA and no viral proteins but can be induced to produce new infectious virions with appropriate cellular stimuli [3,5,6,7]. Due to the long-lived nature of memory CD4+ T cells, integrated HIV can lie dormant as a provirus for decades, undetected by the immune system until reactivation [8,9,10]. Consequently, if ART treatment is stopped then viral outgrowth from reservoirs followed by rapid replication (virologic rebound) invariably occurs in PWH (Figure 2). ART must therefore be taken for life to continually suppress viral replication. ART has other important limitations, including a lack of drug availability in some parts of the world, and requirements for a robust medical infrastructure to deliver therapy and monitor patients for continued virus suppression, which is not available in many resource-limited settings [1]. Additional problems with current therapies include, but are not limited to, the development of antiviral drug resistance [11], side effects associated with long-term drug treatment and ART toxicity, ongoing immune activation and immunological dysfunction in ART-treated individuals [12,13], ART regiment adherence, and continued social stigma faced by PWH. Hence, there remains a compelling need to develop a cure for HIV to permanently prevent rebound from occurring and remove the requirement for long-term ART. To achieve this important goal, further characterization of the latent HIV reservoir is needed, along with the development of new methods to eliminate persistent viral reservoirs.
In this review, we describe the establishment of latent HIV reservoirs and the major challenges associated with eliminating them, highlight current small-animal models that are used to study HIV persistence, and illustrate promising contemporary strategies that are being advanced to eradicate the latent reservoir.

2. Establishment and Maintenance of HIV Latency

Lifelong ART is critical for preventing disease progression in PWH. In the event ART is stopped, even many years after therapy initiation, the virus rapidly re-emerges from rare reservoir cells and replicates to high levels. Understanding the nature of the viral “reservoir” that maintains infectious HIV genomes through long-term ART and eliminating sources of rebounding virus is, therefore, a major focus of the HIV field. HIV infects several different cell types, including CD4+ T helper cells and macrophages [14]. As macrophages are more resistant to the cytopathic effects of HIV infection, they may harbor low levels of HIV expression, thus potentially supporting chronic infection and persistence in immunoprivileged sites. However, there is increasing evidence that macrophage-lineage cells can support latent infection with HIV and may serve as viral reservoirs [15,16,17,18]. Yet, the best-defined and probably the largest persistent HIV reservoir consists of latently infected CD4+ T cells [3,4,5], which contain latent HIV genomes in the form of stable, [3,10] replication-competent proviruses integrated in host chromosomes present for the duration of the host cell’s life. As these cells are not expressing viral proteins, they are not recognized as infected by the host immune system and can thus evade clearance by immune effector mechanisms.
A crucial factor that influences HIV reservoir characteristics is the distinctive physiological properties of the preferred host cells for HIV in vivo. T cells, including CD4+ T cells, can exist in multiple different metabolic or activation states and, importantly, can transition between these states during the lifespan of the cell. While quiescent and waiting to encounter their cognate antigen, T cells are metabolically inactive, rarely divide, and are highly refractory to HIV infection [19]. Once stimulated by cognate antigen recognition, they enter a metabolically active state and proliferate, readily supporting HIV replication and producing abundant progeny virions. Though highly activated CD4+ T cells can harbor latent HIV genomes [20], latently infected cells are typically quiescent. Multiple elements make activated CD4+ T cells the preferred hosts for efficient HIV replication over their resting counterparts. Prior to integration, these include comparatively low CCR5 co-receptor expression on resting cells [21] (although unstimulated cells do also express CXCR4) and inefficient reverse transcription in non-activated cells [19], perturbed through abundantly expressed SAMHD1, which hydrolyzes the dNTPs necessary for HIV cDNA synthesis mediated by reverse transcriptase (RT) [22]. Hence, it is likely that most latently infected cells originally became infected by HIV when the host cell was partially stimulated with cytokines [23], other stimuli, or fully activated through their T-cell receptor and then reverted to a resting state before they could be killed by virus cytopathic effects or immune effector mechanisms. This could occur, for example, when an infected CD4+ T cell transitions to a resting memory cell as part of the normal process of immunological memory formation after antigen exposure. Indeed, CD4+ T cells infected during the process of transitioning from effector to memory cells may be more likely to form post-integration latency due to increases in CCR5 expression coupled with downregulation of cellular transcription [24]. If the resting latently infected host cell is subsequently activated, viral expression can be induced, resulting in the production of new infectious virions capable of seeding rebound.
Several different molecular mechanisms have been identified that can reversibly inhibit expression of HIV and thereby contribute to latency. These include transcription factor availability, the provirus integration site, epigenetic modifications, and the efficiency of transcriptional elongation [25,26,27]. These mechanisms can act individually or in concert to inhibit virus expression, making HIV latency regulation complex, multifactorial, and challenging to reliably disrupt.

2.1. Transcription Factor Availability

Following integration, the limited presence of host transcription factors in the nucleus of resting CD4+ T cells significantly hampers HIV RNA expression, as they are critical for efficient HIV long terminal repeat (LTR)-mediated transcription. The 5′ LTR promoter possesses multiple transcription factor binding sites that positively regulate virus expression, notably nuclear factor-kappa B (NF-kB) and nuclear factor of activated T cells (NFAT) [28,29]. NF-kB and NFAT are sequestered in the cytoplasm in resting cells, but translocate into the nucleus upon cellular activation, where they can promote HIV transcription. NF-kB is present in the cytoplasm of resting cells as a p50/RelA heterodimer, which is associated with I-kB and thereby held in an inactive form. Meanwhile, p50/p50 homodimers of NF-kB, which lack the transactivation domain of the active NF-kB p50/RelA heterodimer, are bound to the proviral HIV LTR in the nucleus of these resting cells. Upon cellular activation, I-kB is phosphorylated, ubiquitylated, and degraded by the 26S proteasome, which allows the p50/RelA heterodimer to translocate to the nucleus and bind to the 5′ LTR promoter, displacing the inactive p50/p50 homodimers. NFAT is similarly present in the cytoplasm of resting cells, and is dephosphorylated upon T cell activation, which induces its nuclear localization [27]. NFAT or NF-kB p50/RelA heterodimers bound to the 5′LTR induce transcriptional activation by recruiting histone acetyltransferases, including p300/CBP, which cause acetylation in histone tails [25,27,30,31]. Expression of the provirus by NF-kB is also dependent on protein–protein interaction between NF-kB and Sp1, which is constitutively bound to an adjacent site on the LTR promoter [32].

2.2. Integration Sites

While HIV is capable of integrating throughout the human genome, transcriptional profiling of CD4+ T cells indicates that the process is not random [33,34]. HIV preferentially integrates into sites in the host cell genome near active transcriptional units [34,35,36] and regional hotspots of enriched active genes near the nuclear envelope, with infected resting CD4+ T cells actively expressing the majority of genes containing integration sites [33]. This bias towards integrating in active genes improves the chances of successful expression, though integration-site frequency may vary in different populations of PWH. Notably, elite controllers have been observed to have increased proviral integration in repressed heterochromatin [37], which primarily consists of satellite DNA. However, the HIV provirus can also be influenced by host genes surrounding the integration site. Transcriptional interference can occur when transcription from the upstream promoter is ongoing as RNA polymerase II (RNAPII) reaches a downstream HIV provirus in the same orientation, leading to HIV-1 promoter occlusion and silencing of transcription [38] in some model systems and, conversely, enhancement of expression in others [39]. Alternatively, in the event HIV integrates in an opposing orientation of an actively transcribed gene, the opposing action of the convergent promotors has an inhibitory effect on HIV transcription and expression [39]. Notably, the overall contribution of transcriptional interference to HIV transcript expression is complex and has been observed to vary widely in different cell line models of HIV latency and primary cell assays [40].

2.3. Epigenetic Modifications

Following integration, HIV DNA is incorporated into the host genome, and, like all genes, is condensed into chromatin. RNA polymerase-mediated transcription and production of full-length HIV transcripts requires accessible chromatin, dictated by whether it is in an active or silent state. As such, methylation, ubiquitylation, and epigenetic chromatin modifiers can affect gene expression. Epigenetic chromatin modifiers regulate chromatin blocks and the transition between “loose” euchromatin and “compact” heterochromatin, and play key roles in regulating latency [41,42,43,44]. More specifically, histone acetylation and deacetylation are well understood as epigenetic regulators of HIV latency, due to histone deacetylase (HDAC)-mediated histone acetylation and the subsequent chromatin “tightening” being correlated with transcriptional repression, whereas histone acetyltransferases (HATs) unwind chromatin and enable transcriptional activation via acetylation [42,45].
Histone modifications additionally have important implications for nucleosome positioning by affecting the association between nucleosomes and DNA, thus potentially restricting transcription [46]. In the context of HIV latency, the deacetylation of viral nucleosome nuc-1 has been shown to be associated with LTR repression, indicating chromatin structure may dictate whether HIV is expressed [42]. Nuc-1 thus is also subject to chromatin remodeling complexes such as BAF (Brg1/Brm-associated factor), which can alter nuc-1 positioning on the LTR into a less energetically favorable position, promoting transcription repression [44].

2.4. Transcriptional Elongation

Given its role as a transcriptional activator, the presence of HIV Tat and its associated cofactors is critical for efficient HIV expression. Before Tat is expressed, HIV transcription is inefficient and associated with promoter-proximal pausing of RNA transcription and deficits in nascent RNA elongation [47]. However, during productive HIV replication, a small amount of Tat protein is still expressed, which binds to the transactivation response element (TAR) RNA stem loop present at the 5′ end of all HIV transcripts [48,49,50,51,52,53,54]. Tat enhances HIV transcription through several mechanisms, including recruitment of elongation regulator P-TEFb to the nascent viral RNA, which relieves the transcriptional block and enables rapid efficient full-length HIV transcript expression [55]. Factors that contribute to HIV latency in resting CD4+ T cells include very low levels of P-TEFb in these cells, and a lack of sufficient Tat expression in the absence of T cell activation [25].

3. Challenges of HIV Latency: Why Is the Latent Reservoir Difficult to Eliminate?

Several characteristics of the latent reservoir make it particularly challenging to study and eliminate. Latently infected cells typically express little or no viral RNA and no viral proteins, and no reliable marker has been identified that externally distinguishes these cells from their non-infected counterparts. Resting memory CD4+ T cells harboring integrated HIV are also rare in vivo, with an occurrence of approximately one in every million resting CD4+ T cells [56], typically translating to around a million latently infected cells per patient. However, latently infected-cell frequencies and overall reservoir size also vary substantially from person to person [56]. Moreover, this estimate of reservoir size is based on outgrowth of infectious HIV following a single round of ex vivo stimulation, and the total size of the reservoir based on cells harboring apparently intact HIV proviral sequences can be substantially higher [20].
Latently infected cells are also widely distributed throughout the body. Around 2% of CD4+ T cells are in the peripheral blood, but most are present in tissues [57], making the majority of the reservoir difficult to access and study in PWH. Simultaneously, the reservoir decays slowly during ART (if at all), and has been documented to have an initial half-life of around 44 months in the first few years of therapy [10] with no long-term reduction observed even after over two decades of treatment [10]. Latency is also established very early in the course of infection, typically within the first days or weeks following exposure [58,59]. Hence, the latent reservoir shows remarkable stability, and is sufficient to maintain lifelong infection, even with continuous ART.

3.1. Clonal Expansion and Homeostatic Proliferation

Several mechanisms can contribute to the maintenance of the latent reservoir over extended periods of time. These include the naturally long lifespan of memory CD4+ T cells in vivo and their propensity to divide via homeostatic proliferation [60,61] or antigen-driven clonal expansion [60,61].
Immunological memory is maintained in part by the occasional homeostatic proliferation of memory T cells, driven by cytokines including interleukin-7 (IL-7) [62]. This cell division can, in some cases, lead to expansion of cell clones containing a latent HIV provirus, producing many cell clones containing genetically identical HIV proviruses with the same HIV integration site [61,63]. Alternatively, the reservoir can be maintained through antigen-dependent clonal expansion. Upon recognizing their cognate antigen, memory CD4+ T cells divide. If a cell possesses an integrated latent HIV provirus and undergoes antigen-dependent clonal expansion, it will generate identical clones which contain integrated latent proviruses, enabling the latent reservoir to persist and potentially expand. Latently infected T cells specific to antigens, including cytomegalovirus (CMV), have been identified to possess identical proviral sequences [62], suggesting responses to common antigens including those encoded by other viruses [64,65] may also induce clonal expansion of the latent reservoir.
In some cases, the specific integration site of the HIV provirus may also influence the host cell in a way that increases propagation of that infected clone [63]. In the event the provirus is integrated into certain genes in the same orientation as the host gene transcription, the resulting insertional mutagenesis may induce proliferation of HIV-infected CD4+ memory T cells while dysregulating cellular growth. For example, identical clones of proviruses in STAT5B, MKL1, and BACH2 have been detected in PWH on ART with a broad tissue distribution [63,66].
Recent estimates suggest that more than 50% of all latently infected cells after ART treatment result from some type of clonal expansion [67]. While host cell activation and proliferation can be associated with latent HIV reactivation and killing of the host cell, these studies illustrate that division of a cell harboring an integrated latent HIV provirus does not always lead to the death of that host cell. Instead, host cell proliferation can function as a mechanism for reservoir expansion, potentially accompanied by the selection of cellular clones bearing proviruses that do not reactivate or result in the killing of the host cell through viral cytopathic effects during cell division.

3.2. Proviral Genetic Diversity and Its Functional Consequences

Beyond the diverse sites of latent HIV integration within the human genome, the HIV provirus itself has high genetic diversity, in part due to the error-prone nature of the reverse-transcriptase (RT) enzyme. RT is a low-fidelity enzyme that lacks exonuclease proofreading activity, resulting in frequent mutations. Moreover, during chronic HIV infection, virus replication occurs simultaneously in many millions of cells each day. The HIV genome is also very tolerant to mutations and can continue producing infectious virus with mutations present at many different sites throughout the genome. Together, these factors allow HIV to evolve rapidly—one million times faster than mammalian DNA [68]. This large diversity provides an abundant pool of mutants for natural selection to act upon. Strong selection pressures may thus favor growth of viruses with resistance to ART drugs, or those bearing CTL escape mutations [69,70,71] or mutations conferring resistance to antibodies [72,73]. These mutant viruses can be deposited in the latent reservoir and maintained during ART, making the reservoir an important repository of clinically relevant mutant viruses.

3.3. Complications of Defective Proviruses

Errors during reverse transcription also may lead to the generation of integrated proviruses with mutations that prevent the production of fully infectious progeny virions. A significant subset of these “defective proviruses” may not be able to express at all, but others have been shown to encode and produce viral proteins such as Gag and/or Nef [74]. These proteins can be recognized by the host immune system, enabling CTL-mediated killing of the host cell [75], contributing to persistent immune activation and chronic inflammation in PWH in ART [76]. The presence of these defective proviruses also complicates molecular analysis of replication-competent latent virus using assays that detect short subgenomic HIV sequences, as the analyzed sequences may not be derived from intact proviruses [77]. Given the role defective proviruses may play in obscuring reservoir size and driving persistent immune activation in PWH on ART, it remains important to understand the factors dictating proviral transcription and protein production for improved therapeutics, as current ART treatments may be unable to target chronic inflammation resulting from HIV RNA produced by defective proviruses.

3.4. Other Reservoir Cells

Beyond CD4+ T cells, other cell types might also serve as viral reservoirs, including myeloid cells [15,16]. In particular, macrophages are readily infected by HIV [78] and tissue resident macrophages such as microglia in the brain can be long-lived [79,80], potentially providing a stable reservoir site. HIV has been detected in macrophages isolated from various tissues in PWH, including the brain and CNS [81,82]. In one recent study, brain myeloid cells were isolated from non-human primates and PWH on ART, and integrated SIV and HIV DNA was detectable in these cells [83]. Virus isolated following ex vivo HDACi stimulation was then able to successfully replicate, supporting the existence of macrophage HIV reservoirs in vivo [83].
In additional work, quantitative viral outgrowth assays with monocyte-derived macrophages [17] were used to show that intact proviruses in macrophages are present in some PWH and capable of producing replication-competent virus that can infect new cells. Moreover, additional cell types have been found in some studies to serve as latent HIV reservoirs and may play a role in pathogenesis [84,85]. Hence, understanding the mechanisms governing HIV latency across all reservoir cell types remains critical for effective cure approaches that target all reservoirs.
In summary, the latent reservoir is dynamic and continuously maintained over time through multiple mechanisms of clonal expansion. Infected cells are widely distributed throughout the body [57,86,87,88], with the majority being concentrated in difficult-to-sample lymphoid organs. Simultaneously, the high genetic diversity of HIV enables its rapid evolution, but errors in viral replication lead to defective proviruses that complicate analysis of the rebound-competent reservoir. As a result of these factors, it remains challenging to identify, isolate, and study latently infected cells.

4. Animal Models of HIV Latency

In vitro models of latent HIV infection using cell lines and primary cells have provided useful isolated environments for mechanistic studies of HIV biology and cure approaches, but do not fully recapitulate additional complex host processes in the whole host organism. For example, in vivo circulation, compartmentalization, metabolism, and physiologic processes coordinated by multiple cell and tissue types cannot be adequately modeled solely with in vitro models. Latent reservoirs are also rare in vivo, and primarily reside in difficult-to-access tissues, complicating their study using clinical samples. Furthermore, interventions to deplete the latent reservoir are often experimental in nature and require preclinical in vivo testing before evaluation in people with HIV. Together, these factors necessitate the use of in vivo animal models, which complement and validate in vitro studies while forming foundations for clinical research. Currently, the most commonly used in vivo models of HIV infection include non-human primates (NHPs) and humanized mice.

4.1. Non-Human Primate Models

NHP models historically include, but are not limited to, HIV-infected chimpanzees (discontinued for ethical reasons, including their endangered-species status) and rhesus, pigtail, or cynomolgus macaques infected with simian immunodeficiency virus (SIV) or recombinant simian–human immunodeficiency virus (SHIV) [89,90]. Because NHPs share a close phylogenetic relationship with humans and can be infected with related SIV and SHIV, their similar physiology and immune system provide key advantages in accurately modeling natural routes of transmission and HIV pathologies such as chronic immune activation and CD4+ T cell depletion. NHP models can be studied in a controlled setting, regulating factors including strain and dose of virus and experimental agents, exposure route, timing of infection, and ART duration, which are otherwise difficult or impossible to control in humans [89,90]. Simultaneously, upon necropsy, they can provide a complete set of viable tissue samples for characterization of the latent reservoir, which are typically unavailable in clinical human studies.
Some limitations of NHP models include their relatively high cost, logistical issues with large-animal housing and handling, species-specific differences between humans and non-human primates, and virologic differences between SIV and HIV. Nevertheless, NHP studies have provided insights into a broad array of topics including viral latency and reservoir analysis [91,92] protection studies [93,94,95], and cure approaches [96,97]. As such, they have proven to be versatile and important in vivo models for HIV persistence during ART.

4.2. Humanized Mouse Models

Though there are no known exogenous lentiviruses that can infect non-modified mice, mouse models have been created that allow infection with HIV while supporting the establishment of a latent reservoir, which serve as important in vivo models for use in HIV persistence studies [97,98,99,100,101,102,103,104]. “Humanized mice” are typically used as models for infection with HIV. These are generated by transplanting human cells and/or tissues into suitable recipient mice [105,106,107]. Many different humanized mouse models exist, which vary in the recipient mouse strain used, the human cell or tissue types transplanted, and the specifics of the experimental procedures (reviewed in [105,106,107]). However, the overall goal of these models is to provide a suitable in vivo environment for HIV replication and study how the virus interacts with cells and tissues in a living host.
To prevent transplanted human tissues from being eliminated by innate and adaptive murine immune responses, immunodeficient mouse strains are used as recipients for human cells. Original recipient mice included nude mice that lack T cells due to Foxn1mu mutations [108] and severe combined immunodeficiency (SCID) mice, which possess mutations in in the protein kinase, DNA-activated, catalytic subunit (Prkdc) gene, abolishing both B and T cells [109]. An alternative pathway for interfering with T cell and B cell development is to disrupt one of the recombination activating genes (Rag1 and Rag2), as they are critically important for B and T cell receptor rearrangements [110,111,112,113]. Rag mutations also do not induce the same higher susceptibility to radiation damage that is associated with the SCID mutation, due to its involvement in DNA repair.
Further reductions in immune competency have been achieved by eliminating receptors for key cytokines. The interleukin-2 (IL-2) common γ chain (IL2Rγ) is a component of the receptors for the cytokines IL-2, IL-4, IL-7, IL-9, IL-15, and IL-21 [114,115]. As such, recipient mouse strains with disrupted IL2Rγ do not respond appropriately to these cytokines. When combined with SCID or Rag mutations, this results in a profoundly immunodeficient animal, with the lack of IL-15 in particular contributing to an absence of NK cells in these animals [116]. Mouse strains bearing these common gamma chain mutations include the NOD.Cg-PrkdcscidIl2rgtm1Wjl/SzJ (NSG), NOD.Cg-Rag1tm1Mom Il2rgtm1Wjl/SzJ (NRG), NOD.Cg-PrkdcscidIl2rgtm1Sug/ShiJic (NOG), and C;129S4-Rag2tm1.1Flv Il2rgtm1.1Flv/J (BRG) mice.
The earliest humanized mouse models for HIV infection included human peripheral-blood leukocyte (hu-PBL) mice [117,118,119]. Hu-PBLs are typically generated through intraperitoneal, intravenous, or intrasplenic injection of mature PBMCs/specific immune cell types into SCID or NSG mice [117]. Despite being a simple, cost-effective model with circulating human immune cells, hu-PBLs are prone to GVHD-induced immune activation [120] occurring within weeks after transplantation, which limits their use in long-term HIV reservoir studies.
SCID-human thymus/liver (hu-Thy/Liv) and human hematopoietic stem cell (hu-HSC or hu-CD34) mice are models where stem cells undergo in vivo differentiation into mature immune cells that can support HIV replication. SCID-hu-Thy/Liv models are composed of a human fetal liver and thymus-derived implant, which is surgically placed under the kidney capsule and provides an environment for CD34+ HSCs present in the fetal liver tissue to differentiate into T cells [121]. This generates an organoid that is structurally and functionally similar to a human thymus, but the resultant cells are primarily immature or naïve T cells constrained to the implant tissue, with few circulating human cells. [106]. Hu-HSCs/hu-CD34s are created by making space in the bone marrow for transplanted stem cells (e.g., through irradiation), then injecting CD34+ HSCs, which in turn differentiate into a diverse population of human immune cells [122] that can circulate systemically.
Interestingly, NOD/SCID mice injected with CD34+ HSCs have been shown to lack peripheral T cells and can support replication of macrophage-tropic HIV, resulting in myeloid-only (MoM) humanized mouse HIV models [16,123], allowing the study of macrophage infection without the confounding presence of T cells. Improved T cell development and mucosal immune cell reconstitution has been achieved by the simultaneous transfusion with CD34+ cells and transplant of Thy/Liv tissue to create the bone marrow–liver–thymus (BLT) mouse [124,125,126]. BLT mice can produce a near-complete complement of human immune lineages in vivo, including T cells, B cells, NK cells, monocyte/macrophages and dendritic cells.
While these models allow multilineage immune reconstitutions with different human immune cell lineages, they do have limitations [107], including reduced cell numbers in certain immune lineages and adaptive immune responses that are not as robust as those occurring in humans. This is being addressed with further advances in recipient mouse strains, which are engineered to express human cytokines, chemokines, or leukocyte antigens (HLAs) to facilitate the differentiation of specific immune-cell lineages, or support the production of adaptive T cell and antibody responses (Figure 3).
Examples of such recipient mouse strains include mice expressing either human IL-15 (NSG-Tg(Hu-IL-15)) alone [127,128] or with human signal regulatory protein α (SIRPA) [129], which facilitate reconstitution with human NK cells, and NRG mice expressing HLA—DR4 alone (DRAG) [130] or in tandem with HLA-A2 (DRAGA), which support improved T and B cell function and development upon HSC engraftment [131,132]. NSG-Tg(Hu-IL-15)) mice have been used in recent HIV cure studies, including NK augmentation and bispecific antibody research [133], while HIV-specific antibody responses and viral replication have been characterized in DRAGA mice [134]. In summary, humanized mice are powerful tools for studying HIV infection in vivo, and recent improvements have further enhanced their physiologic relevancy, providing opportunities to more accurately model HIV infection and the HIV–human immune system interaction in vivo.

5. HIV Cure Approaches

Latently infected cells represent the major reservoir allowing persistence of the virus during ART. Therefore, most HIV cure efforts are directed towards eliminating this reservoir, with the goal of allowing PWH to stop ART without rebound occurring. Multiple cure strategies (Figure 4) have been explored within the field, including killing reservoir cells following virus reactivation, permanently silencing latent genomes, editing and inactivating the latent proviruses, and transplantation or gene-therapy approaches to create an HIV-resistant immune system.

5.1. Latency Reversal

The “activation–elimination” approach, otherwise known as “kick and kill” or “shock and kill”, is a cure strategy in which latently infected cells are induced to express viral proteins (kick) using latency reversal agents (LRAs), allowing the immune system or viral cytopathic effects to eliminate the infected cell (kill). Experimental approaches to reactivating the HIV provirus expression have included protein kinase C (PKC) modulators [101,135,136], second mitochondrial-derived activators of caspases (SMAC) mimetics [137], BET- bromodomain inhibitors [138], histone deacetylase inhibitors (HDACis) [139,140], and various other stimulation methods [7,141].
PKC modulators compete with endogenous ligand diacylglycerol (DAG) to bind to cytosolic PKC. Upon PKC modulator–PKC binding, a complex is formed and translocated into intracellular membranes which induces downstream signaling pathways that result in phosphorylation of IkB, enabling NF-kB-mediated transcription of the HIV provirus and triggering viral protein expression in latently infected cells [101,142]. Naturally-occurring PKC modulators, including bryostatin-1, and prostratin, can reverse HIV from latency in vitro and ex vivo [101,143,144]. Moreover, the design and synthesis of various analogs and slow-release prodrugs of naturally occurring PKC modulators have greatly improved the latency reversal capacity and in vivo tolerability of these compounds [100,101,135,145,146,147,148,149]. For example, a particularly potent bryostatin-1 analog was capable of inducing HIV expression in humanized BLT mice that were infected with HIV and ART-treated, leading to the death of latently infected cells and a reduction in the rebound-competent reservoir [100,101].
The non-canonical NF-kB pathway can also be exploited to reverse HIV latency, [150] and can be activated using SMAC mimetics [97,99,137], which block inhibitors of apoptosis proteins (IAPs) by antagonizing anti-apoptotic XIAP1, cIAP1, and cIAP2, promoting pro-apoptotic activity and enhancing tumor necrosis factor-dependent apoptosis [151]. As cIAP1 and cIAP2 degrade NF-kB-inducing kinase and block p100 processing into p52, SMAC mimetics allow NIK accumulation to eventually trigger the non-canonical pathway of NF-kB activation and the subsequent viral transcription [137].
Both BET-bromodomain inhibitors and HDACis represent LRA classes that target epigenetic modifications that silence expression, restricting HIV expression. BET-bromodomain inhibitors bind acetylated lysine residues (bromodomains) on BET family member BRD4, preventing competition with HIV Tat for downstream interaction with p-TEFB [138,152]. HDACis inhibit the deacetylation of chromatin, unwinding host DNA that may have integrated provirus and allowing access to transcription factors such as NF-kB. Historically, HDACis have been approved for T cell lymphoma treatment, but in vitro and in vivo studies have established that various HDACis, including vorinostat [139], romidepsin [153,154,155], and panobinostat [156], function as LRAs. HDACis have also been documented to synergize with other LRAs, including the PKC modulator bryostatin-1 [144] and SMAC mimetics [137], which may represent promising combinations for future in vivo studies.
Despite these important advances in our understanding of the pathways governing latency and approaches that can exploit those pathways, no LRA has yet been identified that can induce expression of all latent HIV in a safe and effective manner. Promising future efforts in this area are directed at creating improved individual or combination LRAs with better latency reversal properties and augmenting the killing of cells that have been reactivated to express viral proteins, to better eliminate these latently infected cells.

5.2. Kill Augmentation

To fully clear infected cells upon latency reversal, a few key factors must be considered when formulating an appropriate kill strategy. Depending on the immunological status of the individual prior to ART initiation and the length of time since ART was initiated, anti-HIV immune responses may be sub-optimal. This can be due to immunological damage that occurred during the chronic phase of HIV infection, or because anti-HIV immune responses have waned over time, due to insufficient HIV antigen exposure during ART. In untreated infection, CD8+ cytotoxic T lymphocytes (CTLs) can also eventually become exhausted, due to chronic exposure to HIV-1 antigens [157,158]. Moreover, the mutation-prone virus will eventually express epitopes that can escape CTL detection [71]. Finally, HIV-infected cells can downregulate HLA-A and HLA-B, while maintaining the expression of HLA-C and HLA-E, allowing for natural killer (NK) cell and CTL immune evasion [159]. Various methods have been employed to circumvent these barriers to killing HIV-infected cells that have been induced to express from latency.
NK cells have been investigated as a cell type that can be utilized to kill HIV-infected cells (including those reversed from latency) due to their ability to release cytotoxic granules, induce apoptosis via TNF-related apoptosis-inducing ligand (TRAIL) and via Fas/FasL, as well as promote antibody-dependent cellular cytotoxicity (ADCC) [160,161]. Activation of NK cells can be accelerated as a part of the “kick and kill” approach via administration of IL-15 [71,162]. NK cells can also be utilized to selectively target antigens via the attachment of the variable region of an antigen-specific antibody to the variable region of an anti-CD16 antibody (bispecific killer cell engagers (BiKEs)), which can attach to CD16 on the NK cell, thus allowing for NK cell-mediated killing of specific targets [163,164]. In the context of an HIV cure approach, BiKEs that target the gp41 stump of HIV env have been constructed [163]. These BiKEs were constructed by combining PWH-derived gp41 stump-specific antibodies, with the CD16-targeting variable region of a phage-derived antibody (NM3E2) [163]. The efficacy of gp41 stump-specific BiKEs were compared to gp41 stump-specific mAbs, and it was found that NK cell degranulation was 2.5–3.5-fold higher with the BiKEs approach than the mAb approach, quantified by CD107a expression [163]. Within a similar vein, chimeric antigen receptor (CAR) NK cells can be produced by modifying NK cells, for example to express CD4zeta, which can bind to HIV-1 gp120 and kill cells that are productively infected with HIV [165]. One limitation to these NK cell approaches, however, is that NK cells can be subject to immune exhaustion. Exhausted NK cells upregulate PD-1, which makes NK cells unable to degranulate, promote ADCC, and/or effectively produce cytokines [166,167].
Another proposed kill approach is through chimeric antigen receptor T cells (CAR-T Cells), which can specifically target infected cells expressing HIV-1 antigen. CAR T cells have proven effective in cancer treatment [168], and it has been speculated that arming CD8+ T cells with CARs directed towards HIV-1 proteins that are exposed on the surface of productively infected cells may improve their recognition and killing by the transgenic CD8+ cells. Example work involving HIV-specific CARs includes the creation of a second-generation D1D2CAR 4-1BB CAR T cell, which has been successfully used to transduce hematopoietic stem cells during the creation of BLT mice, where they can pass positive and negative selection during thymopoiesis [169]. The D1D2CAR 4-1BB CAR-T cell expresses a truncated CD4 molecule, which is missing the D3 and D4 subunits, allowing it to avoid IL-16 stimulation and engagement of its T cell receptor, and preventing it from serving as an entry receptor for HIV-1 [169]. In vivo data from D1D2CAR 4-1BB CAR-expressing BLT humanized mice have shown a delay in viral rebound after ART cessation, indicating that this kill approach has promise as a potential therapeutic. Future studies could combine the CAR-T cell approach with LRAs to evaluate their effect on reservoirs. Similar to NK cells, however, CAR-T cells also experience exhaustion after prolonged viral antigen exposure [169].
Immune checkpoint protein blockade (ICB) is an additional proposed method to improve the killing of cells productively infected with HIV. Upon prolonged immune activation resulting from persistent infection, CTLs display an exhausted phenotype, usually indicated by the upregulation of immune checkpoint proteins (ICPs) PD-1, LAG3, CTLA-4, TIGIT, CD160, CD244, and TIM3 [157,158]. ICBs can be used to counter the reduced immune response caused by ICPs, resuming the killing of infected cells. The blockade of PD-1 via PD ligand 1 (PD-L1)-specific antibodies has been shown to improve HIV-specific activity in exhausted CTLs [157]. Additionally, the ability for ICB to cause HIV-latency reversal in infected CD4+ T cells has been observed. Ex vivo stimulation of CD4+ T cells isolated from PWH with the PD-1 inhibitor pembrolizumab in conjunction with LRA showed an increase in viral RNA production [170]. Conversely, CD4+ T cells isolated from ART-unsuppressed and -suppressed individuals living with HIV, did not show an increase in viral production upon T cell receptor stimulation during PD-1 engagement [170]. SIV-infected ART-treated rhesus macaques given a CTLA-4 blockade showed a decrease in viral RNA in lymph nodes, as well as improved SIV-specific CD4+ and CD8+ T cell effector function [171].
Clinical trials with ICBs have shown varying levels of effectiveness, as treatment of PWH who also had cancer with anti-PD-1 did not show a consistent reduction in the viral reservoir, quantified by cell-associated HIV-DNA [157]. Additionally, transient HIV-1 latency reversal was observed in an HIV-infected, ART-suppressed patient who was given the immune checkpoint inhibitor of PD-1, nivolumab, while on ART [172]. This was observed in combination with an increase in HIV-specific IFNγ+ CD8+ cells, an increase in plasma IL-6, an increase in CD4+ and CD8+ cell counts, and a decrease in PD-1 in T cells. These findings suggest that nivolumab can improve the HIV-specific CD8-cell-mediated inflammatory response. However, despite these changes in the inflammatory state of the individual, there was little to no impact on reduction of the HIV reservoir. [172]. In an additional study, administration of ipilimumab (an inhibitor of CTLA-4) in a person with HIV, provided evidence of potential HIV latency reversal. Upon ipilimumab treatment, an increase in CD4+ T cells was observed, alongside an increase in cell-associated non-spliced HIV-1 RNA after each dose of ipilimumab. A subsequent decline in plasma HIV-1 RNA was later observed [173]. Additional evidence to support the potential efficacy of ICB on HIV latency reversal was observed in three patients who were given ICB infusions, and showed an increase in cell-associated non-spliced HIV-1 RNA. However, a decrease in HIV DNA was not observed [174]. Overall, these studies indicate that ICB may assist in HIV latency reversal and improve immune responses, but additional approaches will likely be needed to fully deplete the latent reservoir.
Broadly neutralizing antibodies (bNAbs) against HIV envelope proteins have been investigated as an additional kill approach to facilitate eliminating the viral reservoir. For example, 3BNC117, an experimental bNAb that is specific to the CD4 binding site on HIV Env, has been shown to delay viral rebound in a human analytical ART treatment interruption (ATI) study [175]. Additionally, VRC01 is another bNAb specific to the CD4 binding site on HIV-1 Env, but has shown only a modest delay in viral rebound after ART cessation in 24 human participants in two clinical trials, AIDS Clinical Trials Group A5340, and NIH 15-I-0140 [176]. Compared to historical controls, VRC01 caused a slight delay in viral rebound (median viral rebound time in the A5340 and NIH trials was 4 and 5.6 weeks, respectively) [176]. However, a combination approach of anti-HIV bNAbs has also been investigated with much more substantial effects. 3BNC117 and 10-1074, a bNAb that is specific to the base of the V3 loop of gp120, were both administered in humans two days before ART treatment interruption and resulted in a significant delay in viral rebound compared to bNAb monotherapy [177].
Another clinical trial investigated the use of 3BNC117 and 10-1074 co-therapy on ART-suppressed PWH, and assessed the time to viral rebound after ATI. Seven out of seven study participants that were given the bNAb treatment did not experience viral rebound before week 28 of the study (median time to viral rebound was 39.6 weeks), whereas six out of the seven study participants that were given the placebo treatment, experienced viral rebound before week 28 of the study (median time to viral rebound was 9.4 weeks) [178].
Further combinations of bNAbs have also been extended into the clinic, and have shown promising results. Five non-ART-suppressed PWH were given combinations of three bNAbs, PGDM1400 (HIV-1 V2-glycan-specific), PGT121 (V3-glycan-specific), and VRC07-523LS (CD4-binding-site specific) [179]. Individuals given a single infusion of a combination of all three bNAbs showed a sharp and rapid drop in viral load (median time to reach lowest level of viremia was 10 days). Viral rebound to pre-infusion levels occurred after a median of 20 days in all patients [179].
Similar to other cancers, chemotherapy-refractory chronic lymphocytic leukemia (CLL) involves an upregulation of factors involved in cell survival, allowing for the survival of cancerous cells [180,181]. To downregulate antiapoptotic pathways and facilitate killing of infected cells, BH3 (BCL-2 homology domain 3) mimetics, such as venetoclax, have been used with promising results and were approved for the treatment of chemotherapy-refractory chronic lymphocytic leukemia (CLL) [182,183]. In vitro characterization of HIV-infected cells has demonstrated a similar upregulation of antiapoptotic pathways, suggesting that the virus modulates these pathways to ensure the longevity of its host [184]. BH3 mimetics antagonize BCL-2 family pro-survival proteins, allowing cells to complete apoptotic pathways and more efficiently deplete HIV-infected cells [182]. In ex vivo CD4+ T cells from PWH, venetoclax successfully depleted intact and total HIV-1 DNA. Additionally, a prolonged 6-week treatment of venetoclax during ART treatment in humanized mice delayed viral rebound by up to 2 weeks after ART cessation [182]. It has been shown that a combination therapeutic approach of venetoclax with the myeloid cell leukemia sequence 1 (MCL-1) inhibitor S63845 can further delay viral rebound after ART cessation [182]. Inhibition of MCL-1, an anti-apoptotic protein and critical T cell regulator, by S63845, can inhibit antiapoptotic pathways that have been upregulated by HIV, and induce the depletion of infected cells [182,185].
Additional efforts to rejuvenate exhausted T cells have focused on the use of autophagy induction and the mTOR inhibitor rapamycin, which was found to reduce IFN-I signaling, a key driver of T cell exhaustion during an HIV infection [186]. Administration of rapamycin with ART resulted in decreased viral rebound upon ART cessation in BLT mice. Additionally, T cells analyzed from treated animals showed an improvement in functionality and downregulation of exhaustion markers, including PD-1 and TIM3 [186].
Taken together, these approaches (and other similar ones) may prove useful beyond augmenting the “kill” arm of “kick and kill” approaches to latency depletion. Strategies that provide long-term enhancements to HIV immunity, such as stem cell-based CAR-T gene therapies, or vectored production of bNAbs may also contribute to continued immune surveillance. This could help to rapidly eliminate HIV that emerges from rare latently infected cells that evaded cure efforts.

5.3. Stem Cell Gene Therapy

One potential approach for HIV cure is to replace the HIV target cells in vivo with those that are resistant to HIV infection. Without susceptible host cells, the virus would be unable to maintain a persistent infection, even in the absence of ART. This approach has proved successful in several individuals who were living with HIV and also required a bone marrow transplant for acute myeloid leukemia treatment. In these cases, the donors of the CD34+ hematopoietic stem cells used in the transplant each had a naturally occurring homozygous 32-bp deletion in the CCR5 gene (CCR5-∆32), resulting in a truncated CCR5 that is not expressed on the cell surface and thus unable to be used by R5-tropic HIV strains to enter cells [187]. These R5 strains are the most commonly transmitted variants and predominate throughout the presymptomatic phase of infection in most PWH. Several PWH including the “Berlin” [187,188], “Dusseldorf” [189], and “London” patients [190,191] received stem cell transplants from CCR5-∆32 donors. As a result of the treatment, the vast majority of HIV-infected cells present were cleared, while new immune cells were not susceptible to HIV infection, allowing HIV viral loads to remain undetectable even 13 years following transplant in one case [189].
While promising, chemotherapy and CCR5-∆32 bone marrow transplants are associated with many challenges, including high mortality risks and financial costs associated with the procedures, added complications that may result from graft-versus-host disease, and difficulties finding a matching CCR5-∆32 donor. Moreover, while a CCR5-∆32 transplantation may protect against the R5-tropic HIV, recipients remain vulnerable to reactivation of X4-tropic virus harbored prior to receiving their transplant [192]. However, these transplant successes represent important proofs of concept that HIV can be cured, and thus point the way to new approaches that could be more widely adopted. For example, a substantial research effort is currently directed towards developing gene therapy approaches that can disrupt CCR5 in an individual’s own stem cells before being transplanted back into the same person. Though allogeneic stem cell reinfusion has been demonstrated [193], developing autologous transplantation approaches may be necessary to avoid graft incompatibility and difficulties associated with finding appropriate HIV-resistant stem cell donors.

5.4. Block and Lock

A “block and lock” approach has also been considered for HIV cure, which, instead of inducing expression of reservoir cells, has the opposite aim of permanently silencing all latent proviruses. By interfering with HIV transcription, which occurs during reactivation of replication-competent reservoir cells and actively replicating viruses, both can be targeted with this strategy. This could be used to permanently silence latent proviruses and also remove the possibility of ongoing immune activation from HIV proteins produced by defective proviruses. HIV latency is highly dependent upon host transcription factor availability and is influenced by factors involved in chromatin and epigenetic conditions, transcriptional elongation, and integration site selection. Thus, there are multiple potential strategies that have been applied in block and lock approaches, including (but not limited to) didehydro-cortistatin A (dCA)-mediated Tat inhibition, Janus kinase (Jak)-STAT inhibitors, and bromodomain-containing protein 4 (BRD4) modulators [194].
An analog of naturally occurring steroidal alkaloid cortistatin A (CA), dCA binds to the TAR-binding domain of Tat, reducing P-TEFb-mediated transcriptional initiation and elongation [195]. These effects have been shown to be Tat-specific [196], and further observations of long-term dCA treatment have shown it may also block transcription through inducing repressive epigenetic changes such as increased nucleosome occupancy at the nuc-1 region of the 5′ LTR promoter and enhanced recruitment of BAF [196,197]. Importantly, dCA treatment in both BLT mice [197] and primary CD4+ T cells isolated from ART-suppressed patients [198] reduced and delayed viral rebound, establishing proof of concept for a potential functional HIV cure.
T cell homeostasis is, in part, regulated by cytokines that activate the Jak-STAT pathway, which is activated during HIV infection [199,200]. Originally approved for rheumatoid arthritis and myelofibrosis, Jak inhibitors ruxolitinib and tofacitinib have been assessed for their ability to inhibit HIV reactivation in cell lines and primary cells [201], as their anti-inflammatory properties may prevent HIV spread by reducing T cell activation. Both compounds demonstrated selective dose-dependent inhibition of HIV proliferation, indicating the potential for repurposing other similar anti-inflammatory drugs for blocking HIV reactivation.
BRD4 competes with Tat to bind to P-TEFb [152] and thus has been explored as an avenue for inhibiting HIV expression. Small molecule ZL0580, which binds to bromodomain 1 of BRD4, was identified as a compound for suppressing HIV transcription [202]. This is thought to occur through multiple mechanisms: inducing repressive chromatin structure at the LTR promoter and promoting BRD4 to CDK9 interactions, while reducing Tat-CDK9 interactions [202]. Taken together, this research has important implications for the epigenetic role of BRD4 in regulating HIV latency, and thus may represent another class of block and lock compounds.
One caveat to the “block and lock” approach is that replication-competent HIV genomes will remain in the host, as the goal of this strategy is not to eliminate the HIV genome, only silence it. It is difficult to predict with certainty whether these intact HIV proviruses could reactivate at some later date, for example, if transcriptional or epigenetic changes occur that relieve the block on expression. Currently, these agents are unable to induce long-term or permanent silencing of the HIV provirus unless continuously administered. However, they may be able to complement current ART by silencing intact or defective proviruses that contribute to chronic inflammation. Furthermore, the human genome has a large number of endogenous retroviruses that do not express under most circumstances, and driving latent HIV into complete dormancy in a similar fashion may be feasible, particularly if more readily reactivated latent HIV genomes are first purged by LRAs.

5.5. Direct Targeting of the HIV Provirus

Perhaps the most conceptually simple method for eliminating a latent virus that is encoded in host-cell chromosomes is direct editing of the viral sequence to remove or permanently inactivate it. There is a rapidly growing list of powerful molecular biology tools, which can directly edit genomic DNA in a sequence-specific manner, which have been exploited to achieve this goal. Early approaches to disrupting integrated HIV utilized evolved recombinases to excise the provirus [203]. More recent approaches have utilized zinc finger nucleases, transcription activator-like effector nucleases, or clustered regularly interspaced short palindromic repeat (CRISPR)/CRISPR- associated nuclease 9 (Cas9) for gene editing of the HIV provirus. In particular, CRISPR/Cas9 has risen to prominence due to advantages in limited off-target effects, the potential to be delivered by lentiviral and adeno-associated virus vectors [204], and reduced construction time in comparison to transcription activator-like nucleases and zinc finger nucleases. For example, CRISPR/Cas9 was used to genetically disrupt the HIV LTR, inhibiting active and latent provirus transcription [204]. Multiple additional studies have been conducted to assess whether removing integrated HIV or related retroviruses is possible through CRISPR/Cas9, with promising results in vitro [205,206,207] and in vivo [205,206,207].
Beyond direct targeting of the genome, CRISPR/Cas9 has also been used to reactivate latent HIV, where engineered single guide RNAs targeting various regions within the HIV LTR have been delivered in tandem with a deficient Cas9 combined with transcription activator domain-specific single guide RNAs [208,209,210]. Gene editing has also been used to modify host cells to make them refractory to HIV infection, such as CCR5 gene editing to inhibit the spread of R5-tropic HIV [211,212].

6. Conclusions

Developing a safe and scalable cure for HIV that could be extended to the 40 million PWH worldwide remains elusive. This is due to the complex mechanisms of latency establishment and maintenance, and the nature of the latent reservoir in vivo, which is composed of rare, long-lived, and systemically distributed cells. Though there are many complications associated with identifying and targeting reservoir cells, extensive work with isolated molecular biology systems, cell line models, primary human cells, and advanced in vivo animal models of HIV latency have identified several promising approaches for curing HIV. Importantly, these strategies are not mutually exclusive. An ultimate cure may be complex, for example, involving reactivation of the reservoir using LRAs, and the subsequent use of “block and lock” agents to permanently silence residual virus that is readily induced to express. Genetically augmented anti-HIV immune cells and molecules such as CAR-T cells and bNAbs might also be used to aid in the “kick and kill” process then assist long-term immune surveillance to rapidly eliminate any residual virus that evades initial cure efforts.
It should also be noted any HIV cure needs to provide significantly greater benefits and/or diminished side effects in comparison to lifelong ART, because with improved treatment, HIV has become a chronic but manageable condition for many. The kick and kill approach has the limitation that it must reactivate latent HIV in all reservoir cells to cure the infection, but might also be used to complement other cure strategies by depleting any residual reservoir cells that they fail to eliminate. Transplantations, which have proved efficacious in CCR5∆32 donor transplantation recipients, are limited by their complexity and by matching donor availability, and may require gene therapy to successfully scale. Block-and-lock approaches may also be useful as ART regimen supplements because they might prevent residual HIV protein expression that contributes to persistent immune activation and chronic inflammation in PWH on ART. However, they face the challenge of accomplishing permanent HIV provirus silencing in the absence of drugs. Finally, gene editing might offer a permanent solution through completely excising the HIV provirus or modifying susceptible cells to resist infection, but must successfully modify all reservoir cells while simultaneously avoiding off-target effects.
Though each of the approaches described here presents its own merits and challenges, a greater understanding of the mechanisms behind HIV latency maintenance and reservoir establishment, coupled with improvements in therapeutic strategies, show promise in eliminating persistent reservoirs and, ultimately, developing a cure for HIV.

Author Contributions

T.C.C., N.S.M. and M.D.M. all contributed to the writing and editing of this manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

The authors’ laboratory is funded by the National Institute of Health grant numbers AI172727 and AI172410.

Acknowledgments

Figures were created using Biorender.com.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. HIV/AIDS, J.U.N.P. UNAIDS Global AIDS/HIV Statistics Fact Sheet; Joint United Nations Programme on HIV/AIDS 2023; UNAIDS: Geneva, Switzerland, 2023. [Google Scholar]
  2. Department of Health and Human Services. Panel on Antiretroviral Guidelines for Adults and Adolescents; Guidelines for the Use of Antiretroviral Agents in Adults and Adolescents with HIV; Department of Health and Human Services: Washington, DC, USA, 2024. [Google Scholar]
  3. 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] [PubMed]
  4. 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]
  5. Wong, J.K.; Hezareh, M.; Günthard, H.F.; Havlir, D.V.; Ignacio, C.C.; Spina, C.A.; Richman, D.D. Recovery of Replication-Competent HIV despite Prolonged Suppression of Plasma Viremia. Science 1997, 278, 1291–1295. [Google Scholar] [CrossRef]
  6. Chun, T.-W.; Stuyver, L.; Mizell, S.B.; Ehler, L.A.; Mican, J.A.M.; 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]
  7. Marsden, M.D.; Zack, J.A. Experimental Approaches for Eliminating Latent HIV. For. Immunopathol. Dis. Therap. 2015, 6, 91–99. [Google Scholar] [CrossRef] [PubMed]
  8. Chun, T.W.; Davey, R.T.; Engel, D.; Lane, H.C.; Fauci, A.S. Re-Emergence of HIV after Stopping Therapy. Nature 1999, 401, 874–875. [Google Scholar] [CrossRef] [PubMed]
  9. Davey, R.T.; Bhat, N.; Yoder, C.; Chun, T.W.; Metcalf, J.A.; Dewar, R.; Natarajan, V.; Lempicki, R.A.; Adelsberger, J.W.; Miller, K.D.; et al. HIV-1 and T Cell Dynamics after Interruption of Highly Active Antiretroviral Therapy (HAART) in Patients with a History of Sustained Viral Suppression. Proc. Natl. Acad. Sci. USA 1999, 96, 15109–15114. [Google Scholar] [CrossRef] [PubMed]
  10. McMyn, N.F.; Varriale, J.; Fray, E.J.; Zitzmann, C.; MacLeod, H.; Lai, J.; Singhal, A.; Moskovljevic, M.; Garcia, M.A.; Lopez, B.M.; et al. The Latent Reservoir of Inducible, Infectious HIV-1 Does Not Decrease despite Decades of Antiretroviral Therapy. J. Clin. Investig. 2023, 133, e171554. [Google Scholar] [CrossRef] [PubMed]
  11. Gregson, J.; Tang, M.; Ndembi, N.; Hamers, R.L.; Rhee, S.-Y.; Marconi, V.C.; Diero, L.; Brooks, K.A.; Theys, K.; Rinke De Wit, T.; et al. Global Epidemiology of Drug Resistance after Failure of WHO Recommended First-Line Regimens for Adult HIV-1 Infection: A Multicentre Retrospective Cohort Study. Lancet Infect. Dis. 2016, 16, 565–575. [Google Scholar] [CrossRef]
  12. Klatt, N.R.; Chomont, N.; Douek, D.C.; Deeks, S.G. Immune Activation and HIV Persistence: Implications for Curative Approaches to HIV Infection. Immunol. Rev. 2013, 254, 326–342. [Google Scholar] [CrossRef]
  13. Cai, C.W.; Sereti, I. Residual Immune Dysfunction under Antiretroviral Therapy. Semin. Immunol. 2021, 51, 101471. [Google Scholar] [CrossRef] [PubMed]
  14. Moir, S.; Chun, T.-W.; Fauci, A.S. Pathogenic Mechanisms of HIV Disease. Annu. Rev. Pathol. Mech. Dis. 2011, 6, 223–248. [Google Scholar] [CrossRef] [PubMed]
  15. Pham, V.; Marsden, M.D. Can Macrophages Form a Latent Reservoir of HIV? Future Virol. 2021, 16, 75–77. [Google Scholar] [CrossRef] [PubMed]
  16. Honeycutt, J.B.; Thayer, W.O.; Baker, C.E.; Ribeiro, R.M.; Lada, S.M.; Cao, Y.; Cleary, R.A.; Hudgens, M.G.; Richman, D.D.; Garcia, J.V. HIV Persistence in Tissue Macrophages of Humanized Myeloid-Only Mice during Antiretroviral Therapy. Nat. Med. 2017, 23, 638–643. [Google Scholar] [CrossRef] [PubMed]
  17. Veenhuis, R.T.; Abreu, C.M.; Costa, P.A.G.; Ferreira, E.A.; Ratliff, J.; Pohlenz, L.; Shirk, E.N.; Rubin, L.H.; Blankson, J.N.; Gama, L.; et al. Monocyte-Derived Macrophages Contain Persistent Latent HIV Reservoirs. Nat. Microbiol. 2023, 8, 833–844. [Google Scholar] [CrossRef] [PubMed]
  18. Andrade, V.M.; Mavian, C.; Babic, D.; Cordeiro, T.; Sharkey, M.; Barrios, L.; Brander, C.; Martinez-Picado, J.; Dalmau, J.; Llano, A.; et al. A Minor Population of Macrophage-Tropic HIV-1 Variants Is Identified in Recrudescing Viremia Following Analytic Treatment Interruption. Proc. Natl. Acad. Sci. USA 2020, 117, 9981–9990. [Google Scholar] [CrossRef] [PubMed]
  19. Zack, J.A.; Arrigo, S.J.; Weitsman, S.R.; Go, A.S.; Haislip, A.; Chen, I.S.Y. HIV-1 Entry into Quiescent Primary Lymphocytes: Molecular Analysis Reveals a Labile, Latent Viral Structure. Cell 1990, 61, 213–222. [Google Scholar] [CrossRef] [PubMed]
  20. Ho, Y.-C.; Shan, L.; Hosmane, N.N.; Wang, J.; Laskey, S.B.; Rosenbloom, D.I.S.; Lai, J.; Blankson, J.N.; Siliciano, J.D.; Siliciano, R.F. Replication-Competent Noninduced Proviruses in the Latent Reservoir Increase Barrier to HIV-1 Cure. Cell 2013, 155, 540–551. [Google Scholar] [CrossRef] [PubMed]
  21. Bleul, C.C.; Wu, L.; Hoxie, J.A.; Springer, T.A.; Mackay, C.R. The HIV Coreceptors CXCR4 and CCR5 Are Differentially Expressed and Regulated on Human T Lymphocytes. Proc. Natl. Acad. Sci. USA 1997, 94, 1925–1930. [Google Scholar] [CrossRef]
  22. Baldauf, H.-M.; Pan, X.; Erikson, E.; Schmidt, S.; Daddacha, W.; Burggraf, M.; Schenkova, K.; Ambiel, I.; Wabnitz, G.; Gramberg, T.; et al. SAMHD1 Restricts HIV-1 Infection in Resting CD4+ T Cells. Nat. Med. 2012, 18, 1682–1688. [Google Scholar] [CrossRef]
  23. Cameron, P.U.; Saleh, S.; Sallmann, G.; Solomon, A.; Wightman, F.; Evans, V.A.; Boucher, G.; Haddad, E.K.; Sekaly, R.-P.; Harman, A.N.; et al. Establishment of HIV-1 Latency in Resting CD4+ T Cells Depends on Chemokine-Induced Changes in the Actin Cytoskeleton. Proc. Natl. Acad. Sci. USA 2010, 107, 16934–16939. [Google Scholar] [CrossRef]
  24. Shan, L.; Deng, K.; Gao, H.; Xing, S.; Capoferri, A.A.; Durand, C.M.; Rabi, S.A.; Laird, G.M.; Kim, M.; Hosmane, N.N.; et al. Transcriptional Reprogramming during Effector-to-Memory Transition Renders CD4+ T Cells Permissive for Latent HIV-1 Infection. Immunity 2017, 47, 766–775.e3. [Google Scholar] [CrossRef] [PubMed]
  25. Mbonye, U.; Karn, J. The Cell Biology of HIV-1 Latency and Rebound. Retrovirology 2024, 21, 6. [Google Scholar] [CrossRef]
  26. 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. [Google Scholar] [CrossRef] [PubMed]
  27. Ruelas, D.S.; Greene, W.C. An Integrated Overview of HIV-1 Latency. Cell 2013, 155, 519–529. [Google Scholar] [CrossRef]
  28. Nabel, G.; Baltimore, D. An Inducible Transcription Factor Activates Expression of Human Immunodeficiency Virus in T Cells. Nature 1987, 326, 711–713. [Google Scholar] [CrossRef]
  29. Duverger, A.; Wolschendorf, F.; Zhang, M.; Wagner, F.; Hatcher, B.; Jones, J.; Cron, R.Q.; Van Der Sluis, R.M.; Jeeninga, R.E.; Berkhout, B.; et al. An AP-1 Binding Site in the Enhancer/Core Element of the HIV-1 Promoter Controls the Ability of HIV-1 to Establish Latent Infection. J. Virol. 2013, 87, 2264–2277. [Google Scholar] [CrossRef]
  30. Perkins, N.D.; Felzien, L.K.; Betts, J.C.; Leung, K.; Beach, D.H.; Nabel, G.J. Regulation of NF-kappaB by Cyclin-Dependent Kinases Associated with the P300 Coactivator. Science 1997, 275, 523–527. [Google Scholar] [CrossRef]
  31. García-Rodríguez, C.; Rao, A. Nuclear Factor of Activated T Cells (NFAT)-Dependent Transactivation Regulated by the Coactivators P300/CREB-Binding Protein (CBP). J. Exp. Med. 1998, 187, 2031–2036. [Google Scholar] [CrossRef] [PubMed]
  32. 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]
  33. Serrao, E.; Krishnan, L.; Shun, M.-C.; Li, X.; Cherepanov, P.; Engelman, A.; Maertens, G.N. Integrase Residues That Determine Nucleotide Preferences at Sites of HIV-1 Integration: Implications for the Mechanism of Target DNA Binding. Nucleic Acids Res. 2014, 42, 5164–5176. [Google Scholar] [CrossRef] [PubMed]
  34. Schröder, A.R.W.; 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] [PubMed]
  35. Han, Y.; Lassen, K.; Monie, D.; Sedaghat, A.R.; Shimoji, S.; Liu, X.; Pierson, T.C.; Margolick, J.B.; Siliciano, R.F.; Siliciano, J.D. Resting CD4 + T Cells from Human Immunodeficiency Virus Type 1 (HIV-1)-Infected Individuals Carry Integrated HIV-1 Genomes within Actively Transcribed Host Genes. J. Virol. 2004, 78, 6122–6133. [Google Scholar] [CrossRef] [PubMed]
  36. Liu, H.; Dow, E.C.; Arora, R.; Kimata, J.T.; Bull, L.M.; Arduino, R.C.; Rice, A.P. Integration of Human Immunodeficiency Virus Type 1 in Untreated Infection Occurs Preferentially within Genes. J. Virol. 2006, 80, 7765–7768. [Google Scholar] [CrossRef]
  37. Woldemeskel, B.A.; Kwaa, A.K.; Blankson, J.N. Viral Reservoirs in Elite Controllers of HIV-1 Infection: Implications for HIV Cure Strategies. eBioMedicine 2020, 62, 103118. [Google Scholar] [CrossRef] [PubMed]
  38. Lenasi, T.; Contreras, X.; Peterlin, B.M. Transcriptional Interference Antagonizes Proviral Gene Expression to Promote HIV Latency. Cell Host Microbe 2008, 4, 123–133. [Google Scholar] [CrossRef] [PubMed]
  39. Han, Y.; Lin, Y.B.; An, W.; Xu, J.; Yang, H.-C.; O’Connell, K.; Dordai, D.; Boeke, J.D.; Siliciano, J.D.; Siliciano, R.F. Orientation-Dependent Regulation of Integrated HIV-1 Expression by Host Gene Transcriptional Readthrough. Cell Host Microbe 2008, 4, 134–146. [Google Scholar] [CrossRef] [PubMed]
  40. Telwatte, S.; Morón-López, S.; Aran, D.; Kim, P.; Hsieh, C.; Joshi, S.; Montano, M.; Greene, W.C.; Butte, A.J.; Wong, J.K.; et al. Heterogeneity in HIV and Cellular Transcription Profiles in Cell Line Models of Latent and Productive Infection: Implications for HIV Latency. Retrovirology 2019, 16, 32. [Google Scholar] [CrossRef]
  41. Tyagi, M.; Pearson, R.J.; Karn, J. Establishment of HIV Latency in Primary CD4+ Cells Is Due to Epigenetic Transcriptional Silencing and P-TEFb Restriction. J. Virol. 2010, 84, 6425–6437. [Google Scholar] [CrossRef]
  42. 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]
  43. Peterson, J.J.; Lewis, C.A.; Burgos, S.D.; Manickam, A.; Xu, Y.; Rowley, A.A.; Clutton, G.; Richardson, B.; Zou, F.; Simon, J.M.; et al. A Histone Deacetylase Network Regulates Epigenetic Reprogramming and Viral Silencing in HIV-Infected Cells. Cell Chem. Biol. 2023, 30, 1617–1633.e9. [Google Scholar] [CrossRef] [PubMed]
  44. 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] [PubMed]
  45. Jiang, G.; Espeseth, A.; Hazuda, D.J.; Margolis, D.M. C-Myc and Sp1 Contribute to Proviral Latency by Recruiting Histone Deacetylase 1 to the Human Immunodeficiency Virus Type 1 Promoter. J. Virol. 2007, 81, 10914–10923. [Google Scholar] [CrossRef] [PubMed]
  46. Venkatesh, S.; Workman, J.L. Histone Exchange, Chromatin Structure and the Regulation of Transcription. Nat. Rev. Mol. Cell Biol. 2015, 16, 178–189. [Google Scholar] [CrossRef] [PubMed]
  47. Herrmann, C.H.; Rice, A.P. Specific Interaction of the Human Immunodeficiency Virus Tat Proteins with a Cellular Protein Kinase. Virology 1993, 197, 601–608. [Google Scholar] [CrossRef] [PubMed]
  48. Feng, S.; Holland, E.C. HIV-1 Tat Trans-Activation Requires the Loop Sequence within Tar. Nature 1988, 334, 165–167. [Google Scholar] [CrossRef] [PubMed]
  49. Madore, S.J.; Cullen, B.R. Genetic Analysis of the Cofactor Requirement for Human Immunodeficiency Virus Type 1 Tat Function. J. Virol. 1993, 67, 3703–3711. [Google Scholar] [CrossRef] [PubMed]
  50. Ashe, M.P.; Griffin, P.; James, W.; Proudfoot, N.J. Poly(A) Site Selection in the HIV-1 Provirus: Inhibition of Promoter-Proximal Polyadenylation by the Downstream Major Splice Donor Site. Genes Dev. 1995, 9, 3008–3025. [Google Scholar] [CrossRef]
  51. Kaiser, P.; Joshi, S.K.; Kim, P.; Li, P.; Liu, H.; Rice, A.P.; Wong, J.K.; Yukl, S.A. Assays for Precise Quantification of Total (Including Short) and Elongated HIV-1 Transcripts. J. Virol. Methods 2017, 242, 1–8. [Google Scholar] [CrossRef]
  52. Yukl, S.A.; Kaiser, P.; Kim, P.; Telwatte, S.; Joshi, S.K.; Vu, M.; Lampiris, H.; Wong, J.K. HIV Latency in Isolated Patient CD4+ T Cells May Be Due to Blocks in HIV Transcriptional Elongation, Completion, and Splicing. Sci. Transl. Med. 2018, 10, eaap9927. [Google Scholar] [CrossRef]
  53. Feinberg, M.B.; Baltimore, D.; Frankel, A.D. The Role of Tat in the Human Immunodeficiency Virus Life Cycle Indicates a Primary Effect on Transcriptional Elongation. Proc. Natl. Acad. Sci. USA 1991, 88, 4045–4049. [Google Scholar] [CrossRef] [PubMed]
  54. Laspia, M.F.; Rice, A.P.; Mathews, M.B. HIV-1 Tat Protein Increases Transcriptional Initiation and Stabilizes Elongation. Cell 1989, 59, 283–292. [Google Scholar] [CrossRef] [PubMed]
  55. Kessler, M.; Mathews, M.B. Premature Termination and Processing of Human Immunodeficiency Virus Type 1-Promoted Transcripts. J. Virol. 1992, 66, 4488–4496. [Google Scholar] [CrossRef]
  56. Chun, T.-W.; Carruth, L.; Finzi, D.; Shen, X.; DiGiuseppe, J.A.; Taylor, H.; Hermankova, M.; Chadwick, K.; Margolick, J.; Quinn, T.C.; et al. Quantification of Latent Tissue Reservoirs and Total Body Viral Load in HIV-1 Infection. Nature 1997, 387, 183–188. [Google Scholar] [CrossRef] [PubMed]
  57. Wong, J.K.; Yukl, S.A. Tissue Reservoirs of HIV. Curr. Opin. HIV AIDS 2016, 11, 362–370. [Google Scholar] [CrossRef] [PubMed]
  58. Moran, J.A.; Marsden, M.D. HIV Establishes an Early Foothold. Cell Host Microbe 2023, 31, 571–573. [Google Scholar] [CrossRef] [PubMed]
  59. Gantner, P.; Buranapraditkun, S.; Pagliuzza, A.; Dufour, C.; Pardons, M.; Mitchell, J.L.; Kroon, E.; Sacdalan, C.; Tulmethakaan, N.; Pinyakorn, S.; et al. HIV Rapidly Targets a Diverse Pool of CD4+ T Cells to Establish Productive and Latent Infections. Immunity 2023, 56, 653–668.e5. [Google Scholar] [CrossRef] [PubMed]
  60. Wang, Z.; Gurule, E.E.; Brennan, T.P.; Gerold, J.M.; Kwon, K.J.; Hosmane, N.N.; Kumar, M.R.; Beg, S.A.; Capoferri, A.A.; Ray, S.C.; et al. Expanded Cellular Clones Carrying Replication-Competent HIV-1 Persist, Wax, and Wane. Proc. Natl. Acad. Sci. USA 2018, 115, E2575–E2584. [Google Scholar] [CrossRef] [PubMed]
  61. Chomont, N.; El-Far, M.; Ancuta, P.; Trautmann, L.; Procopio, F.A.; Yassine-Diab, B.; Boucher, G.; Boulassel, M.-R.; Ghattas, G.; Brenchley, J.M.; et al. HIV Reservoir Size and Persistence Are Driven by T Cell Survival and Homeostatic Proliferation. Nat. Med. 2009, 15, 893–900. [Google Scholar] [CrossRef]
  62. Simonetti, F.R.; Zhang, H.; Soroosh, G.P.; Duan, J.; Rhodehouse, K.; Hill, A.L.; Beg, S.A.; McCormick, K.; Raymond, H.E.; Nobles, C.L.; et al. Antigen-Driven Clonal Selection Shapes the Persistence of HIV-1–Infected CD4+ T Cells In Vivo. J. Clin. Investig. 2021, 131, e145254. [Google Scholar] [CrossRef]
  63. Maldarelli, F.; Wu, X.; Su, L.; Simonetti, F.R.; Shao, W.; Hill, S.; Spindler, J.; Ferris, A.L.; Mellors, J.W.; Kearney, M.F.; et al. Specific HIV Integration Sites Are Linked to Clonal Expansion and Persistence of Infected Cells. Science 2014, 345, 179–183. [Google Scholar] [CrossRef] [PubMed]
  64. Gianella, S.; Anderson, C.M.; Var, S.R.; Oliveira, M.F.; Lada, S.M.; Vargas, M.V.; Massanella, M.; Little, S.J.; Richman, D.D.; Strain, M.C.; et al. Replication of Human Herpesviruses Is Associated with Higher HIV DNA Levels during Antiretroviral Therapy Started at Early Phases of HIV Infection. J. Virol. 2016, 90, 3944–3952. [Google Scholar] [CrossRef] [PubMed]
  65. Chaillon, A.; Nakazawa, M.; Rawlings, S.A.; Curtin, G.; Caballero, G.; Scott, B.; Anderson, C.; Gianella, S. Subclinical Cytomegalovirus and Epstein-Barr Virus Shedding Is Associated with Increasing HIV DNA Molecular Diversity in Peripheral Blood during Suppressive Antiretroviral Therapy. J. Virol. 2020, 94, e00927-20. [Google Scholar] [CrossRef] [PubMed]
  66. McManus, W.R.; Bale, M.J.; Spindler, J.; Wiegand, A.; Musick, A.; Patro, S.C.; Sobolewski, M.D.; Musick, V.K.; Anderson, E.M.; Cyktor, J.C.; et al. HIV-1 in Lymph Nodes Is Maintained by Cellular Proliferation during Antiretroviral Therapy. J. Clin. Investig. 2019, 129, 4629–4642. [Google Scholar] [CrossRef] [PubMed]
  67. Yeh, Y.-H.J.; Yang, K.; Razmi, A.; Ho, Y.-C. The Clonal Expansion Dynamics of the HIV-1 Reservoir: Mechanisms of Integration Site-Dependent Proliferation and HIV-1 Persistence. Viruses 2021, 13, 1858. [Google Scholar] [CrossRef] [PubMed]
  68. Li, W.H.; Tanimura, M.; Sharp, P.M. Rates and Dates of Divergence between AIDS Virus Nucleotide Sequences. Mol. Biol. Evol. 1988, 5, 313–330. [Google Scholar] [CrossRef] [PubMed]
  69. Phillips, R.E.; Rowland-Jones, S.; Nixon, D.F.; Gotch, F.M.; Edwards, J.P.; Ogunlesi, A.O.; Elvin, J.G.; Rothbard, J.A.; Bangham, C.R.M.; Rizza, C.R.; et al. Human Immunodeficiency Virus Genetic Variation That Can Escape Cytotoxic T Cell Recognition. Nature 1991, 354, 453–459. [Google Scholar] [CrossRef] [PubMed]
  70. Collins, K.L.; Chen, B.K.; Kalams, S.A.; Walker, B.D.; Baltimore, D. HIV-1 Nef Protein Protects Infected Primary Cells against Killing by Cytotoxic T Lymphocytes. Nature 1998, 391, 397–401. [Google Scholar] [CrossRef]
  71. Deng, K.; Pertea, M.; Rongvaux, A.; Wang, L.; Durand, C.M.; Ghiaur, G.; Lai, J.; McHugh, H.L.; Hao, H.; Zhang, H.; et al. Broad CTL Response Is Required to Clear Latent HIV-1 Due to Dominance of Escape Mutations. Nature 2015, 517, 381–385. [Google Scholar] [CrossRef]
  72. Dacheux, L.; Moreau, A.; Ataman-Önal, Y.; Biron, F.; Verrier, B.; Barin, F. Evolutionary Dynamics of the Glycan Shield of theHuman Immunodeficiency Virus Envelope during Natural Infection andImplications for Exposure of the 2G12Epitope. J. Virol. 2004, 78, 12625–12637. [Google Scholar] [CrossRef]
  73. Moore, P.L.; Ranchobe, N.; Lambson, B.E.; Gray, E.S.; Cave, E.; Abrahams, M.-R.; Bandawe, G.; Mlisana, K.; Abdool Karim, S.S.; Williamson, C.; et al. Limited Neutralizing Antibody Specificities Drive Neutralization Escape in Early HIV-1 Subtype C Infection. PLoS Pathog. 2009, 5, e1000598. [Google Scholar] [CrossRef] [PubMed]
  74. Imamichi, H.; Smith, M.; Adelsberger, J.W.; Izumi, T.; Scrimieri, F.; Sherman, B.T.; Rehm, C.A.; Imamichi, T.; Pau, A.; Catalfamo, M.; et al. Defective HIV-1 Proviruses Produce Viral Proteins. Proc. Natl. Acad. Sci. USA 2020, 117, 3704–3710. [Google Scholar] [CrossRef]
  75. Pollack, R.A.; Jones, R.B.; Pertea, M.; Bruner, K.M.; Martin, A.R.; Thomas, A.S.; Capoferri, A.A.; Beg, S.A.; Huang, S.-H.; Karandish, S.; et al. Defective HIV-1 Proviruses Are Expressed and Can Be Recognized by Cytotoxic T Lymphocytes, Which Shape the Proviral Landscape. Cell Host Microbe 2017, 21, 494–506.e4. [Google Scholar] [CrossRef] [PubMed]
  76. Singh, K.; Natarajan, V.; Dewar, R.; Rupert, A.; Badralmaa, Y.; Zhai, T.; Winchester, N.; Scrimieri, F.; Smith, M.; Davis, I.; et al. Long-Term Persistence of Transcriptionally Active ‘Defective’ HIV-1 Proviruses: Implications for Persistent Immune Activation during Antiretroviral Therapy. AIDS 2023, 37, 2119–2130. [Google Scholar] [CrossRef] [PubMed]
  77. Reeves, D.B.; Gaebler, C.; Oliveira, T.Y.; Peluso, M.J.; Schiffer, J.T.; Cohn, L.B.; Deeks, S.G.; Nussenzweig, M.C. Impact of Misclassified Defective Proviruses on HIV Reservoir Measurements. Nat. Commun. 2023, 14, 4186. [Google Scholar] [CrossRef] [PubMed]
  78. Aquaro, S.; Bagnarelli, P.; Guenci, T.; De Luca, A.; Clementi, M.; Balestra, E.; Caliò, R.; Perno, C. Long-term Survival and Virus Production in Human Primary Macrophages Infected by Human Immunodeficiency Virus. J. Med. Virol. 2002, 68, 479–488. [Google Scholar] [CrossRef] [PubMed]
  79. Lassmann, H.; Schmied, M.; Vass, K.; Hickey, W.F. Bone Marrow Derived Elements and Resident Microglia in Brain Inflammation. Glia 1993, 7, 19–24. [Google Scholar] [CrossRef] [PubMed]
  80. Garaci, E.; Caroleo, M.C.; Aloe, L.; Aquaro, S.; Piacentini, M.; Costa, N.; Amendola, A.; Micera, A.; Caliò, R.; Perno, C.-F.; et al. Nerve Growth Factor Is an Autocrine Factor Essential for the Survival of Macrophages Infected with HIV. Proc. Natl. Acad. Sci. USA 1999, 96, 14013–14018. [Google Scholar] [CrossRef] [PubMed]
  81. Ko, A.; Kang, G.; Hattler, J.B.; Galadima, H.I.; Zhang, J.; Li, Q.; Kim, W.-K. Macrophages but Not Astrocytes Harbor HIV DNA in the Brains of HIV-1-Infected Aviremic Individuals on Suppressive Antiretroviral Therapy. J. Neuroimmune Pharmacol. 2019, 14, 110–119. [Google Scholar] [CrossRef]
  82. Fischer-Smith, T.; Croul, S.; Sverstiuk, A.E.; Capini, C.; L’Heureux, D.; Régulier, E.G.; Richardson, M.W.; Amini, S.; Morgello, S.; Khalili, K.; et al. CNS Invasion by CD14+/CD16+ Peripheral Blood-Derived Monocytes in HIV Dementia: Perivascular Accumulation and Reservoir of HIV Infection. J. Neurovirol. 2001, 7, 528–541. [Google Scholar] [CrossRef]
  83. 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] [PubMed]
  84. Sebastian, N.T.; Zaikos, T.D.; Terry, V.; Taschuk, F.; McNamara, L.A.; Onafuwa-Nuga, A.; Yucha, R.; Signer, R.A.J.; Riddell Iv, J.; Bixby, D.; et al. CD4 Is Expressed on a Heterogeneous Subset of Hematopoietic Progenitors, Which Persistently Harbor CXCR4 and CCR5-Tropic HIV Proviral Genomes in Vivo. PLoS Pathog. 2017, 13, e1006509. [Google Scholar] [CrossRef]
  85. Sacha, J.B.; Ndhlovu, L.C. Strategies to Target Non-T-Cell HIV Reservoirs. Curr. Opin. HIV AIDS 2016, 11, 376–382. [Google Scholar] [CrossRef] [PubMed]
  86. Chun, T.; Nickle, D.C.; Justement, J.S.; Meyers, J.H.; Roby, G.; Hallahan, C.W.; Kottilil, S.; Moir, S.; Mican, J.M.; Mullins, J.I.; et al. Persistence of HIV in Gut-Associated Lymphoid Tissue despite Long-Term Antiretroviral Therapy. J. Infect. Dis. 2008, 197, 714–720. [Google Scholar] [CrossRef] [PubMed]
  87. Fletcher, C.V.; Staskus, K.; Wietgrefe, S.W.; Rothenberger, M.; Reilly, C.; Chipman, J.G.; Beilman, G.J.; Khoruts, A.; Thorkelson, A.; Schmidt, T.E.; et al. Persistent HIV-1 Replication Is Associated with Lower Antiretroviral Drug Concentrations in Lymphatic Tissues. Proc. Natl. Acad. Sci. USA 2014, 111, 2307–2312. [Google Scholar] [CrossRef] [PubMed]
  88. Jenabian, M.-A.; Costiniuk, C.T.; Mehraj, V.; Ghazawi, F.M.; Fromentin, R.; Brousseau, J.; Brassard, P.; Bélanger, M.; Ancuta, P.; Bendayan, R.; et al. Immune Tolerance Properties of the Testicular Tissue as a Viral Sanctuary Site in ART-Treated HIV-Infected Adults. AIDS 2016, 30, 2777–2786. [Google Scholar] [CrossRef] [PubMed]
  89. Evans, D.T.; Silvestri, G. Nonhuman Primate Models in AIDS Research. Curr. Opin. HIV AIDS 2013, 8, 255–261. [Google Scholar] [CrossRef] [PubMed]
  90. Estes, J.D.; Wong, S.W.; Brenchley, J.M. Nonhuman Primate Models of Human Viral Infections. Nat. Rev. Immunol. 2018, 18, 390–404. [Google Scholar] [CrossRef] [PubMed]
  91. Estes, J.D.; Kityo, C.; Ssali, F.; Swainson, L.; Makamdop, K.N.; Del Prete, G.Q.; Deeks, S.G.; Luciw, P.A.; Chipman, J.G.; Beilman, G.J.; et al. Defining Total-Body AIDS-Virus Burden with Implications for Curative Strategies. Nat. Med. 2017, 23, 1271–1276. [Google Scholar] [CrossRef]
  92. Dinoso, J.B.; Rabi, S.A.; Blankson, J.N.; Gama, L.; Mankowski, J.L.; Siliciano, R.F.; Zink, M.C.; Clements, J.E. A Simian Immunodeficiency Virus-Infected Macaque Model To Study Viral Reservoirs That Persist during Highly Active Antiretroviral Therapy. J. Virol. 2009, 83, 9247–9257. [Google Scholar] [CrossRef]
  93. Desrosiers, R.C.; Wyand, M.S.; Kodama, T.; Ringler, D.J.; Arthur, L.O.; Sehgal, P.K.; Letvin, N.L.; King, N.W.; Daniel, M.D. Vaccine Protection against Simian Immunodeficiency Virus Infection. Proc. Natl. Acad. Sci. USA 1989, 86, 6353–6357. [Google Scholar] [CrossRef] [PubMed]
  94. 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] [PubMed]
  95. Johnson, P.R.; Schnepp, B.C.; Zhang, J.; Connell, M.J.; Greene, S.M.; Yuste, E.; Desrosiers, R.C.; Reed Clark, K. Vector-Mediated Gene Transfer Engenders Long-Lived Neutralizing Activity and Protection against SIV Infection in Monkeys. Nat. Med. 2009, 15, 901–906. [Google Scholar] [CrossRef] [PubMed]
  96. 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] [PubMed]
  97. 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-κB Signalling in Vivo. Nature 2020, 578, 160–165. [Google Scholar] [CrossRef] [PubMed]
  98. Kim, J.T.; Zhang, T.-H.; Carmona, C.; Lee, B.; Seet, C.S.; Kostelny, M.; Shah, N.; Chen, H.; Farrell, K.; Soliman, M.S.A.; et al. Latency Reversal plus Natural Killer Cells Diminish HIV Reservoir in Vivo. Nat. Commun. 2022, 13, 121. [Google Scholar] [CrossRef]
  99. 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-κB Signaling Activates Latent HIV-1 Reservoirs In Vivo. Cell Rep. Med. 2020, 1, 100037. [Google Scholar] [CrossRef] [PubMed]
  100. Marsden, M.D.; Zhang, T.; Du, Y.; Dimapasoc, M.; Soliman, M.S.A.; Wu, X.; Kim, J.T.; Shimizu, A.; Schrier, A.; Wender, P.A.; et al. Tracking HIV Rebound Following Latency Reversal Using Barcoded HIV. Cell Rep. Med. 2020, 1, 100162. [Google Scholar] [CrossRef]
  101. 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]
  102. Marsden, M.D.; Kovochich, M.; Suree, N.; Shimizu, S.; Mehta, R.; Cortado, R.; Bristol, G.; An, D.S.; Zack, J.A. HIV Latency in the Humanized BLT Mouse. J. Virol. 2012, 86, 339–347. [Google Scholar] [CrossRef]
  103. Brooks, D.G.; Hamer, D.H.; Arlen, P.A.; Gao, L.; Bristol, G.; Kitchen, C.M.R.; Berger, E.A.; Zack, J.A. Molecular Characterization, Reactivation, and Depletion of Latent HIV. Immunity 2003, 19, 413–423. [Google Scholar] [CrossRef] [PubMed]
  104. Choudhary, S.K.; Archin, N.M.; Cheema, M.; Dahl, N.P.; Garcia, J.V.; Margolis, D.M. Latent HIV-1 Infection of Resting CD4+ T Cells in the Humanized Rag2−/− γc−/− Mouse. J. Virol. 2012, 86, 114–120. [Google Scholar] [CrossRef] [PubMed]
  105. Marsden, M.D.; Zack, J.A. Studies of Retroviral Infection in Humanized Mice. Virology 2015, 479–480, 297–309. [Google Scholar] [CrossRef]
  106. Marsden, M.D.; Zack, J.A. Humanized Mouse Models for Human Immunodeficiency Virus Infection. Annu. Rev. Virol. 2017, 4, 393–412. [Google Scholar] [CrossRef]
  107. Marsden, M.D. Benefits and Limitations of Humanized Mice in HIV Persistence Studies. Retrovirology 2020, 17, 7. [Google Scholar] [CrossRef] [PubMed]
  108. Flanagan, S.P. ‘Nude’, a New Hairless Gene with Pleiotropic Effects in the Mouse. Genet. Res. 1966, 8, 295–309. [Google Scholar] [CrossRef]
  109. Bosma, G.C.; Custer, R.P.; Bosma, M.J. A Severe Combined Immunodeficiency Mutation in the Mouse. Nature 1983, 301, 527–530. [Google Scholar] [CrossRef] [PubMed]
  110. Pearson, T.; Shultz, L.D.; Miller, D.; King, M.; Laning, J.; Fodor, W.; Cuthbert, A.; Burzenski, L.; Gott, B.; Lyons, B.; et al. Non-Obese Diabetic–Recombination Activating Gene-1 (NOD– Rag 1 Null) Interleukin (IL)-2 Receptor Common Gamma Chain (IL 2 RγNull) Null Mice: A Radioresistant Model for Human Lymphohaematopoietic Engraftment. Clin. Exp. Immunol. 2008, 154, 270–284. [Google Scholar] [CrossRef]
  111. Shultz, L.D.; Lyons, B.L.; Burzenski, L.M.; Gott, B.; Chen, X.; Chaleff, S.; Kotb, M.; Gillies, S.D.; King, M.; Mangada, J.; et al. Human Lymphoid and Myeloid Cell Development in NOD/LtSz-Scid IL2RγNull Mice Engrafted with Mobilized Human Hemopoietic Stem Cells. J. Immunol. 2005, 174, 6477–6489. [Google Scholar] [CrossRef]
  112. Mombaerts, P.; Iacomini, J.; Johnson, R.S.; Herrup, K.; Tonegawa, S.; Papaioannou, V.E. RAG-1-Deficient Mice Have No Mature B and T Lymphocytes. Cell 1992, 68, 869–877. [Google Scholar] [CrossRef]
  113. Shinkai, Y. RAG-2-Deficient Mice Lack Mature Lymphocytes Owing to Inability to Initiate V(D)J Rearrangement. Cell 1992, 68, 855–867. [Google Scholar] [CrossRef] [PubMed]
  114. Cao, X.; Shores, E.W.; Hu-Li, J.; Anver, M.R.; Kelsail, B.L.; Russell, S.M.; Drago, J.; Noguchi, M.; Grinberg, A.; Bloom, E.T.; et al. Defective Lymphoid Development in Mice Lacking Expression of the Common Cytokine Receptor γ Chain. Immunity 1995, 2, 223–238. [Google Scholar] [CrossRef] [PubMed]
  115. DiSanto, J.P.; Müller, W.; Guy-Grand, D.; Fischer, A.; Rajewsky, K. Lymphoid Development in Mice with a Targeted Deletion of the Interleukin 2 Receptor Gamma Chain. Proc. Natl. Acad. Sci. USA 1995, 92, 377–381. [Google Scholar] [CrossRef] [PubMed]
  116. Ranson, T.; Vosshenrich, C.A.J.; Corcuff, E.; Richard, O.; Müller, W.; Di Santo, J.P. IL-15 Is an Essential Mediator of Peripheral NK-Cell Homeostasis. Blood 2003, 101, 4887–4893. [Google Scholar] [CrossRef] [PubMed]
  117. Mosier, D.E.; Gulizia, R.J.; Baird, S.M.; Wilson, D.B. Transfer of a Functional Human Immune System to Mice with Severe Combined Immunodeficiency. Nature 1988, 335, 256–259. [Google Scholar] [CrossRef] [PubMed]
  118. King, M.A.; Covassin, L.; Brehm, M.A.; Racki, W.; Pearson, T.; Leif, J.; Laning, J.; Fodor, W.; Foreman, O.; Burzenski, L.; et al. Human Peripheral Blood Leucocyte Non-Obese Diabetic-Severe Combined Immunodeficiency Interleukin-2 Receptor Gamma Chain Gene Mouse Model of Xenogeneic Graft- versus -Host-like Disease and the Role of Host Major Histocompatibility Complex. Clin. Exp. Immunol. 2009, 157, 104–118. [Google Scholar] [CrossRef] [PubMed]
  119. Harui, A.; Kiertscher, S.M.; Roth, M.D. Reconstitution of huPBL-NSG Mice with Donor-Matched Dendritic Cells Enables Antigen-Specific T-Cell Activation. J. Neuroimmune Pharmacol. 2011, 6, 148–157. [Google Scholar] [CrossRef] [PubMed]
  120. Kim, J.T.; Bresson-Tan, G.; Zack, J.A. Current Advances in Humanized Mouse Models for Studying NK Cells and HIV Infection. Microorganisms 2023, 11, 1984. [Google Scholar] [CrossRef]
  121. McCune, J.; Namikawa, R.; Kaneshima, H.; Shultz, L.; Lieberman, M.; Weissman, I. The SCID-Hu Mouse: Murine Model for the Analysis of Human Hematolymphoid Differentiation and Function. Science 1988, 241, 1632–1639. [Google Scholar] [CrossRef]
  122. Traggiai, E.; Chicha, L.; Mazzucchelli, L.; Bronz, L.; Piffaretti, J.-C.; Lanzavecchia, A.; Manz, M.G. Development of a Human Adaptive Immune System in Cord Blood Cell-Transplanted Mice. Science 2004, 304, 104–107. [Google Scholar] [CrossRef]
  123. Honeycutt, J.B.; Wahl, A.; Baker, C.; Spagnuolo, R.A.; Foster, J.; Zakharova, O.; Wietgrefe, S.; Caro-Vegas, C.; Madden, V.; Sharpe, G.; et al. Macrophages Sustain HIV Replication in Vivo Independently of T Cells. J. Clin. Investig. 2016, 126, 1353–1366. [Google Scholar] [CrossRef] [PubMed]
  124. Melkus, M.W.; Estes, J.D.; Padgett-Thomas, A.; Gatlin, J.; Denton, P.W.; Othieno, F.A.; Wege, A.K.; Haase, A.T.; Garcia, J.V. Humanized Mice Mount Specific Adaptive and Innate Immune Responses to EBV and TSST-1. Nat. Med. 2006, 12, 1316–1322. [Google Scholar] [CrossRef] [PubMed]
  125. Tonomura, N.; Habiro, K.; Shimizu, A.; Sykes, M.; Yang, Y.-G. Antigen-Specific Human T-Cell Responses and T Cell-Dependent Production of Human Antibodies in a Humanized Mouse Model. Blood 2008, 111, 4293–4296. [Google Scholar] [CrossRef]
  126. Lan, P.; Tonomura, N.; Shimizu, A.; Wang, S.; Yang, Y.-G. Reconstitution of a Functional Human Immune System in Immunodeficient Mice through Combined Human Fetal Thymus/Liver and CD34+ Cell Transplantation. Blood 2006, 108, 487–492. [Google Scholar] [CrossRef]
  127. Abeynaike, S.A.; Huynh, T.R.; Mehmood, A.; Kim, T.; Frank, K.; Gao, K.; Zalfa, C.; Gandarilla, A.; Shultz, L.; Paust, S. Human Hematopoietic Stem Cell Engrafted IL-15 Transgenic NSG Mice Support Robust NK Cell Responses and Sustained HIV-1 Infection. Viruses 2023, 15, 365. [Google Scholar] [CrossRef]
  128. Aryee, K.; Burzenski, L.M.; Yao, L.; Keck, J.G.; Greiner, D.L.; Shultz, L.D.; Brehm, M.A. Enhanced Development of Functional Human NK Cells in NOD-scid-IL2rgnull Mice Expressing Human IL15. FASEB J. 2022, 36, e22476. [Google Scholar] [CrossRef]
  129. Herndler-Brandstetter, D.; Shan, L.; Yao, Y.; Stecher, C.; Plajer, V.; Lietzenmayer, M.; Strowig, T.; De Zoete, M.R.; Palm, N.W.; Chen, J.; et al. Humanized Mouse Model Supports Development, Function, and Tissue Residency of Human Natural Killer Cells. Proc. Natl. Acad. Sci. USA 2017, 114, E9626–E9634. [Google Scholar] [CrossRef]
  130. Danner, R.; Chaudhari, S.N.; Rosenberger, J.; Surls, J.; Richie, T.L.; Brumeanu, T.-D.; Casares, S. Expression of HLA Class II Molecules in Humanized NOD.Rag1KO.IL2RgcKO Mice Is Critical for Development and Function of Human T and B Cells. PLoS ONE 2011, 6, e19826. [Google Scholar] [CrossRef] [PubMed]
  131. Majji, S.; Wijayalath, W.; Shashikumar, S.; Pow-Sang, L.; Villasante, E.; Brumeanu, T.D.; Casares, S. Differential Effect of HLA Class-I versus Class-II Transgenes on Human T and B Cell Reconstitution and Function in NRG Mice. Sci. Rep. 2016, 6, 28093. [Google Scholar] [CrossRef]
  132. Majji, S.; Wijayalath, W.; Shashikumar, S.; Brumeanu, T.D.; Casares, S. Humanized DRAGA Mice Immunized with Plasmodium Falciparum Sporozoites and Chloroquine Elicit Protective Pre-Erythrocytic Immunity. Malar. J. 2018, 17, 114. [Google Scholar] [CrossRef]
  133. Board, N.L.; Yuan, Z.; Wu, F.; Moskovljevic, M.; Ravi, M.; Sengupta, S.; Mun, S.S.; Simonetti, F.R.; Lai, J.; Tebas, P.; et al. Bispecific Antibodies Promote Natural Killer Cell-Mediated Elimination of HIV-1 Reservoir Cells. Nat. Immunol. 2024, 25, 462–470. [Google Scholar] [CrossRef] [PubMed]
  134. Ollerton, M.T.; Folkvord, J.M.; Peachman, K.K.; Shashikumar, S.; Morrison, E.B.; Jagodzinski, L.L.; Peel, S.A.; Khreiss, M.; D’Aquila, R.T.; Casares, S.; et al. HIV-1 Infected Humanized DRAGA Mice Develop HIV-Specific Antibodies despite Lack of Canonical Germinal Centers in Secondary Lymphoid Tissues. Front. Immunol. 2022, 13, 1047277. [Google Scholar] [CrossRef] [PubMed]
  135. Sloane, J.L.; Benner, N.L.; Keenan, K.N.; Zang, X.; Soliman, M.S.A.; Wu, X.; Dimapasoc, M.; Chun, T.-W.; Marsden, M.D.; Zack, J.A.; et al. Prodrugs of PKC Modulators Show Enhanced HIV Latency Reversal and an Expanded Therapeutic Window. Proc. Natl. Acad. Sci. USA 2020, 117, 10688–10698. [Google Scholar] [CrossRef] [PubMed]
  136. Albert, B.J.; Niu, A.; Ramani, R.; Marshall, G.R.; Wender, P.A.; Williams, R.M.; Ratner, L.; Barnes, A.B.; Kyei, G.B. Combinations of Isoform-Targeted Histone Deacetylase Inhibitors and Bryostatin Analogues Display Remarkable Potency to Activate Latent HIV without Global T-Cell Activation. Sci. Rep. 2017, 7, 7456. [Google Scholar] [CrossRef] [PubMed]
  137. 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] [PubMed]
  138. 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] [PubMed]
  139. 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]
  140. Wu, G.; Swanson, M.; Talla, A.; Graham, D.; Strizki, J.; Gorman, D.; Barnard, R.J.O.; Blair, W.; Søgaard, O.S.; Tolstrup, M.; et al. HDAC Inhibition Induces HIV-1 Protein and Enables Immune-Based Clearance Following Latency Reversal. JCI Insight 2017, 2, e92901. [Google Scholar] [CrossRef] [PubMed]
  141. Marsden, M.D.; Zack, J.A. HIV Cure Strategies: A Complex Approach for a Complicated Viral Reservoir? Future Virol. 2019, 14, 5–8. [Google Scholar] [CrossRef]
  142. Kulkosky, J.; Culnan, D.M.; Roman, J.; Dornadula, G.; Schnell, M.; Boyd, M.R.; Pomerantz, R.J. Prostratin: Activation of Latent HIV-1 Expression Suggests a Potential Inductive Adjuvant Therapy for HAART. Blood 2001, 98, 3006–3015. [Google Scholar] [CrossRef]
  143. Gutiérrez, C.; Serrano-Villar, S.; Madrid-Elena, N.; Pérez-Elías, M.J.; Martín, M.E.; Barbas, C.; Ruipérez, J.; Muñoz, E.; Muñoz-Fernández, 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] [PubMed]
  144. Perez, M.; De Vinuesa, A.; Sanchez-Duffhues, G.; Marquez, N.; Bellido, M.; Munoz-Fernandez, M.; Moreno, S.; Castor, T.; Calzado, M.; Munoz, E. Bryostatin-1 Synergizes with Histone Deacetylase Inhibitors to Reactivate HIV-1 from Latency. Curr. HIV Res. 2010, 8, 418–429. [Google Scholar] [CrossRef] [PubMed]
  145. Wender, P.A.; Hardman, C.T.; Ho, S.; Jeffreys, M.S.; Maclaren, J.K.; Quiroz, R.V.; Ryckbosch, S.M.; Shimizu, A.J.; Sloane, J.L.; Stevens, M.C. Scalable Synthesis of Bryostatin 1 and Analogs, Adjuvant Leads against Latent HIV. Science 2017, 358, 218–223. [Google Scholar] [CrossRef] [PubMed]
  146. Hardman, C.; Ho, S.; Shimizu, A.; Luu-Nguyen, Q.; Sloane, J.L.; Soliman, M.S.A.; Marsden, M.D.; Zack, J.A.; Wender, P.A. Synthesis and Evaluation of Designed PKC Modulators for Enhanced Cancer Immunotherapy. Nat. Commun. 2020, 11, 1879. [Google Scholar] [CrossRef] [PubMed]
  147. Marsden, M.D.; Wu, X.; Navab, S.M.; Loy, B.A.; Schrier, A.J.; DeChristopher, B.A.; Shimizu, A.J.; Hardman, C.T.; Ho, S.; Ramirez, C.M.; et al. Characterization of Designed, Synthetically Accessible Bryostatin Analog HIV Latency Reversing Agents. Virology 2018, 520, 83–93. [Google Scholar] [CrossRef] [PubMed]
  148. Beans, E.J.; Fournogerakis, D.; Gauntlett, C.; Heumann, L.V.; Kramer, R.; Marsden, M.D.; Murray, D.; Chun, T.-W.; Zack, J.A.; Wender, P.A. Highly Potent, Synthetically Accessible Prostratin Analogs Induce Latent HIV Expression in Vitro and Ex Vivo. Proc. Natl. Acad. Sci. USA 2013, 110, 11698–11703. [Google Scholar] [CrossRef] [PubMed]
  149. DeChristopher, B.A.; Loy, B.A.; Marsden, M.D.; Schrier, A.J.; Zack, J.A.; Wender, P.A. Designed, Synthetically Accessible Bryostatin Analogues Potently Induce Activation of Latent HIV Reservoirs In Vitro. Nat. Chem. 2012, 4, 705–710. [Google Scholar] [CrossRef] [PubMed]
  150. Sun, S.-C. Non-Canonical NF-κB Signaling Pathway. Cell Res. 2011, 21, 71–85. [Google Scholar] [CrossRef]
  151. 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]
  152. 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]
  153. 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]
  154. 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]
  155. Søgaard, 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]
  156. 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]
  157. Baron, M.; Soulié, C.; Lavolé, A.; Assoumou, L.; Abbar, B.; Fouquet, B.; Rousseau, A.; Veyri, M.; Samri, A.; Makinson, A.; et al. Impact of Anti PD-1 Immunotherapy on HIV Reservoir and Anti-Viral Immune Responses in People Living with HIV and Cancer. Cells 2022, 11, 1015. [Google Scholar] [CrossRef]
  158. Ward, A.R.; Mota, T.M.; Jones, R.B. Immunological Approaches to HIV Cure. Semin. Immunol. 2021, 51, 101412. [Google Scholar] [CrossRef]
  159. Cohen, G.B.; Gandhi, R.T.; Davis, D.M.; Mandelboim, O.; Chen, B.K.; Strominger, J.L.; Baltimore, D. The Selective Downregulation of Class I Major Histocompatibility Complex Proteins by HIV-1 Protects HIV-Infected Cells from NK Cells. Immunity 1999, 10, 661–671. [Google Scholar] [CrossRef]
  160. Bournazos, S.; Wang, T.T.; Dahan, R.; Maamary, J.; Ravetch, J.V. Signaling by Antibodies: Recent Progress. Annu. Rev. Immunol. 2017, 35, 285–311. [Google Scholar] [CrossRef]
  161. Prager, I.; Liesche, C.; Van Ooijen, H.; Urlaub, D.; Verron, Q.; Sandström, N.; Fasbender, F.; Claus, M.; Eils, R.; Beaudouin, J.; et al. NK Cells Switch from Granzyme B to Death Receptor–Mediated Cytotoxicity during Serial Killing. J. Exp. Med. 2019, 216, 2113–2127. [Google Scholar] [CrossRef]
  162. Flórez-Álvarez, L.; Hernandez, J.C.; Zapata, W. NK Cells in HIV-1 Infection: From Basic Science to Vaccine Strategies. Front. Immunol. 2018, 9, 2290. [Google Scholar] [CrossRef]
  163. Ramadoss, N.S.; Zhao, N.Q.; Richardson, B.A.; Grant, P.M.; Kim, P.S.; Blish, C.A. Enhancing Natural Killer Cell Function with Gp41-Targeting Bispecific Antibodies to Combat HIV Infection. AIDS 2020, 34, 1313–1323. [Google Scholar] [CrossRef] [PubMed]
  164. Duan, S.; Liu, S. Targeting NK Cells for HIV-1 Treatment and Reservoir Clearance. Front. Immunol. 2022, 13, 842746. [Google Scholar] [CrossRef] [PubMed]
  165. Ni, Z.; Knorr, D.A.; Bendzick, L.; Allred, J.; Kaufman, D.S. Expression of Chimeric Receptor CD4ζ by Natural Killer Cells Derived from Human Pluripotent Stem Cells Improves In Vitro Activity but Does Not Enhance Suppression of HIV Infection In Vivo. Stem Cells 2014, 32, 1021–1031. [Google Scholar] [CrossRef] [PubMed]
  166. Norris, S.; Coleman, A.; Kuri-Cervantes, L.; Bower, M.; Nelson, M.; Goodier, M.R. PD-1 Expression on Natural Killer Cells and CD8+ T Cells During Chronic HIV-1 Infection. Viral Immunol. 2012, 25, 329–332. [Google Scholar] [CrossRef] [PubMed]
  167. Schafer, J.L.; Li, H.; Evans, T.I.; Estes, J.D.; Reeves, R.K. Accumulation of Cytotoxic CD16+ NK Cells in Simian Immunodeficiency Virus-Infected Lymph Nodes Associated with In Situ Differentiation and Functional Anergy. J. Virol. 2015, 89, 6887–6894. [Google Scholar] [CrossRef] [PubMed]
  168. Feins, S.; Kong, W.; Williams, E.F.; Milone, M.C.; Fraietta, J.A. An Introduction to Chimeric Antigen Receptor (CAR) T-cell Immunotherapy for Human Cancer. Am. J. Hematol. 2019, 94, S3–S9. [Google Scholar] [CrossRef] [PubMed]
  169. Zhen, A.; Carrillo, M.A.; Mu, W.; Rezek, V.; Martin, H.; Hamid, P.; Chen, I.S.Y.; Yang, O.O.; Zack, J.A.; Kitchen, S.G. Robust CAR-T Memory Formation and Function via Hematopoietic Stem Cell Delivery. PLoS Pathog. 2021, 17, e1009404. [Google Scholar] [CrossRef] [PubMed]
  170. Fromentin, R.; DaFonseca, S.; Costiniuk, C.T.; El-Far, M.; Procopio, F.A.; Hecht, F.M.; Hoh, R.; Deeks, S.G.; Hazuda, D.J.; Lewin, S.R.; et al. PD-1 Blockade Potentiates HIV Latency Reversal Ex Vivo in CD4+ T Cells from ART-Suppressed Individuals. Nat. Commun. 2019, 10, 814. [Google Scholar] [CrossRef] [PubMed]
  171. Hryniewicz, A.; Boasso, A.; Edghill-Smith, Y.; Vaccari, M.; Fuchs, D.; Venzon, D.; Nacsa, J.; Betts, M.R.; Tsai, W.-P.; Heraud, J.-M.; et al. CTLA-4 Blockade Decreases TGF-Beta, IDO, and Viral RNA Expression in Tissues of SIVmac251-Infected Macaques. Blood 2006, 108, 3834–3842. [Google Scholar] [CrossRef]
  172. Le Garff, G.; Samri, A.; Lambert-Niclot, S.; Even, S.; Lavolé, A.; Cadranel, J.; Spano, J.-P.; Autran, B.; Marcelin, A.-G.; Guihot, A. Transient HIV-Specific T Cells Increase and Inflammation in an HIV-Infected Patient Treated with Nivolumab. AIDS 2017, 31, 1048–1051. [Google Scholar] [CrossRef]
  173. Wightman, F.; Solomon, A.; Kumar, S.S.; Urriola, N.; Gallagher, K.; Hiener, B.; Palmer, S.; Mcneil, C.; Garsia, R.; Lewin, S.R. Effect of Ipilimumab on the HIV Reservoir in an HIV-Infected Individual with Metastatic Melanoma. AIDS 2015, 29, 504–506. [Google Scholar] [CrossRef] [PubMed]
  174. Lau, J.S.Y.; McMahon, J.H.; Gubser, C.; Solomon, A.; Chiu, C.Y.H.; Dantanarayana, A.; Chea, S.; Tennakoon, S.; Zerbato, J.M.; Garlick, J.; et al. The Impact of Immune Checkpoint Therapy on the Latent Reservoir in HIV-Infected Individuals with Cancer on Antiretroviral Therapy. AIDS 2021, 35, 1631–1636. [Google Scholar] [CrossRef] [PubMed]
  175. Scheid, J.F.; Horwitz, J.A.; Bar-On, Y.; Kreider, E.F.; Lu, C.-L.; Lorenzi, J.C.C.; Feldmann, A.; Braunschweig, M.; Nogueira, L.; Oliveira, T.; et al. HIV-1 Antibody 3BNC117 Suppresses Viral Rebound in Humans during Treatment Interruption. Nature 2016, 535, 556–560. [Google Scholar] [CrossRef] [PubMed]
  176. Bar, K.J.; Sneller, M.C.; Harrison, L.J.; Justement, J.S.; Overton, E.T.; Petrone, M.E.; Salantes, D.B.; Seamon, C.A.; Scheinfeld, B.; Kwan, R.W.; et al. Effect of HIV Antibody VRC01 on Viral Rebound after Treatment Interruption. N. Engl. J. Med. 2016, 375, 2037–2050. [Google Scholar] [CrossRef] [PubMed]
  177. Mendoza, P.; Gruell, H.; Nogueira, L.; Pai, J.A.; Butler, A.L.; Millard, K.; Lehmann, C.; Suárez, I.; Oliveira, T.Y.; Lorenzi, J.C.C.; et al. Combination Therapy with Anti-HIV-1 Antibodies Maintains Viral Suppression. Nature 2018, 561, 479–484. [Google Scholar] [CrossRef] [PubMed]
  178. Sneller, M.C.; Blazkova, J.; Justement, J.S.; Shi, V.; Kennedy, B.D.; Gittens, K.; Tolstenko, J.; McCormack, G.; Whitehead, E.J.; Schneck, R.F.; et al. Combination Anti-HIV Antibodies Provide Sustained Virological Suppression. Nature 2022, 606, 375–381. [Google Scholar] [CrossRef]
  179. Julg, B.; Stephenson, K.E.; Wagh, K.; Tan, S.C.; Zash, R.; Walsh, S.; Ansel, J.; Kanjilal, D.; Nkolola, J.; Walker-Sperling, V.E.K.; et al. Safety and Antiviral Activity of Triple Combination Broadly Neutralizing Monoclonal Antibody Therapy against HIV-1: A Phase 1 Clinical Trial. Nat. Med. 2022, 28, 1288–1296. [Google Scholar] [CrossRef] [PubMed]
  180. Kern, C.; Cornuel, J.-F.; Billard, C.; Tang, R.; Rouillard, D.; Stenou, V.; Defrance, T.; Ajchenbaum-Cymbalista, F.; Simonin, P.-Y.; Feldblum, S.; et al. Involvement of BAFF and APRIL in the Resistance to Apoptosis of B-CLL through an Autocrine Pathway. Blood 2004, 103, 679–688. [Google Scholar] [CrossRef] [PubMed]
  181. Cuní, S.; Pérez-Aciego, P.; Pérez-Chacón, G.; Vargas, J.A.; Sánchez, A.; Martín-Saavedra, F.M.; Ballester, S.; García-Marco, J.; Jordá, J.; Durántez, A. A Sustained Activation of PI3K/NF-κB Pathway Is Critical for the Survival of Chronic Lymphocytic Leukemia B Cells. Leukemia 2004, 18, 1391–1400. [Google Scholar] [CrossRef]
  182. Arandjelovic, P.; Kim, Y.; Cooney, J.P.; Preston, S.P.; Doerflinger, M.; McMahon, J.H.; Garner, S.E.; Zerbato, J.M.; Roche, M.; Tumpach, C.; et al. Venetoclax, Alone and in Combination with the BH3 Mimetic S63845, Depletes HIV-1 Latently Infected Cells and Delays Rebound in Humanized Mice. Cell Rep. Med. 2023, 4, 101178. [Google Scholar] [CrossRef]
  183. Roberts, A.W.; Davids, M.S.; Pagel, J.M.; Kahl, B.S.; Puvvada, S.D.; Gerecitano, J.F.; Kipps, T.J.; Anderson, M.A.; Brown, J.R.; Gressick, L.; et al. Targeting BCL2 with Venetoclax in Relapsed Chronic Lymphocytic Leukemia. N. Engl. J. Med. 2016, 374, 311–322. [Google Scholar] [CrossRef] [PubMed]
  184. Kuo, H.-H.; Ahmad, R.; Lee, G.Q.; Gao, C.; Chen, H.-R.; Ouyang, Z.; Szucs, M.J.; Kim, D.; Tsibris, A.; Chun, T.-W.; et al. Anti-Apoptotic Protein BIRC5 Maintains Survival of HIV-1-Infected CD4+ T Cells. Immunity 2018, 48, 1183–1194.e5. [Google Scholar] [CrossRef] [PubMed]
  185. Opferman, J.T.; Letai, A.; Beard, C.; Sorcinelli, M.D.; Ong, C.C.; Korsmeyer, S.J. Development and Maintenance of B and T Lymphocytes Requires Antiapoptotic MCL-1. Nature 2003, 426, 671–676. [Google Scholar] [CrossRef] [PubMed]
  186. Mu, W.; Rezek, V.; Martin, H.; Carrillo, M.A.; Tomer, S.; Hamid, P.; Lizarraga, M.A.; Tibbe, T.D.; Yang, O.O.; Jamieson, B.D.; et al. Autophagy Inducer Rapamycin Treatment Reduces IFN-I–Mediated Inflammation and Improves Anti–HIV-1 T Cell Response In Vivo. JCI Insight 2022, 7, e159136. [Google Scholar] [CrossRef] [PubMed]
  187. Allers, K.; Schneider, T. CCR5Δ32 Mutation and HIV Infection: Basis for Curative HIV Therapy. Curr. Opin. Virol. 2015, 14, 24–29. [Google Scholar] [CrossRef] [PubMed]
  188. Allers, K.; Hütter, G.; Hofmann, J.; Loddenkemper, C.; Rieger, K.; Thiel, E.; Schneider, T. Evidence for the Cure of HIV Infection by CCR5Δ32/Δ32 Stem Cell Transplantation. Blood 2011, 117, 2791–2799. [Google Scholar] [CrossRef] [PubMed]
  189. Jensen, B.-E.O.; Knops, E.; Cords, L.; Lübke, N.; Salgado, M.; Busman-Sahay, K.; Estes, J.D.; Huyveneers, L.E.P.; Perdomo-Celis, F.; Wittner, M.; et al. In-Depth Virological and Immunological Characterization of HIV-1 Cure after CCR5Δ32/Δ32 Allogeneic Hematopoietic Stem Cell Transplantation. Nat. Med. 2023, 29, 583–587. [Google Scholar] [CrossRef] [PubMed]
  190. Gupta, R.K.; Peppa, D.; Hill, A.L.; Gálvez, C.; Salgado, M.; Pace, M.; McCoy, L.E.; Griffith, S.A.; Thornhill, J.; Alrubayyi, A.; et al. Evidence for HIV-1 Cure after CCR5Δ32/Δ32 Allogeneic Haemopoietic Stem-Cell Transplantation 30 Months Post Analytical Treatment Interruption: A Case Report. Lancet HIV 2020, 7, e340–e347. [Google Scholar] [CrossRef] [PubMed]
  191. Hütter, G.; Nowak, D.; Mossner, M.; Ganepola, S.; Müßig, A.; Allers, K.; Schneider, T.; Hofmann, J.; Kücherer, 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]
  192. Abbate, A.; Gold, K.J.; Goldman, E.B.; Moseley, K.L. More on Shift of HIV Tropism in Stem-Cell Transplantation with CCR5 Delta32/Delta32 Mutation. N. Engl. J. Med. 2014, 371, 2437–2438. [Google Scholar] [CrossRef]
  193. Xu, L.; Wang, J.; Liu, Y.; Xie, L.; Su, B.; Mou, D.; Wang, L.; Liu, T.; Wang, X.; Zhang, B.; et al. CRISPR-Edited Stem Cells in a Patient with HIV and Acute Lymphocytic Leukemia. N. Engl. J. Med. 2019, 381, 1240–1247. [Google Scholar] [CrossRef] [PubMed]
  194. Vansant, G.; Bruggemans, A.; Janssens, J.; Debyser, Z. Block-And-Lock Strategies to Cure HIV Infection. Viruses 2020, 12, 84. [Google Scholar] [CrossRef] [PubMed]
  195. Mousseau, G.; Clementz, M.A.; Bakeman, W.N.; Nagarsheth, N.; Cameron, M.; Shi, J.; Baran, P.; Fromentin, R.; Chomont, N.; Valente, S.T. An Analog of the Natural Steroidal Alkaloid Cortistatin A Potently Suppresses Tat-Dependent HIV Transcription. Cell Host Microbe 2012, 12, 97–108. [Google Scholar] [CrossRef] [PubMed]
  196. 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] [PubMed]
  197. 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]
  198. 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, e00465-15. [Google Scholar] [CrossRef] [PubMed]
  199. Kohler, J.J.; Tuttle, D.L.; Coberley, C.R.; Sleasman, J.W.; Goodenow, M.M. Human Immunodeficiency Virus Type 1 (HIV-1) Induces Activation of Multiple STATs in CD4+ Cells of Lymphocyte or Monocyte/Macrophage Lineages. J. Leukoc. Biol. 2003, 73, 407–416. [Google Scholar] [CrossRef] [PubMed]
  200. Alhetheel, A.; Yakubtsov, Y.; Abdkader, K.; Sant, N.; Diaz-Mitoma, F.; Kumar, A.; Kryworuchko, M. Amplification of the Signal Transducer and Activator of Transcription I Signaling Pathway and Its Association with Apoptosis in Monocytes from HIV-Infected Patients. AIDS 2008, 22, 1137–1144. [Google Scholar] [CrossRef] [PubMed]
  201. Gavegnano, C.; Detorio, M.; Montero, C.; Bosque, A.; Planelles, V.; Schinazi, R.F. Ruxolitinib and Tofacitinib Are Potent and Selective Inhibitors of HIV-1 Replication and Virus Reactivation In Vitro. Antimicrob. Agents Chemother. 2014, 58, 1977–1986. [Google Scholar] [CrossRef]
  202. Niu, Q.; Liu, Z.; Alamer, E.; Fan, X.; Chen, H.; Endsley, J.; Gelman, B.B.; Tian, B.; Kim, J.H.; Michael, N.L.; et al. Structure-Guided Drug Design Identifies a BRD4-Selective Small Molecule That Suppresses HIV. J. Clin. Investig. 2019, 129, 3361–3373. [Google Scholar] [CrossRef]
  203. Sarkar, I.; Hauber, I.; Hauber, J.; Buchholz, F. HIV-1 Proviral DNA Excision Using an Evolved Recombinase. Science 2007, 316, 1912–1915. [Google Scholar] [CrossRef] [PubMed]
  204. Ebina, H.; Misawa, N.; Kanemura, Y.; Koyanagi, Y. Harnessing the CRISPR/Cas9 System to Disrupt Latent HIV-1 Provirus. Sci. Rep. 2013, 3, 2510. [Google Scholar] [CrossRef] [PubMed]
  205. Hu, W.; Kaminski, R.; Yang, F.; Zhang, Y.; Cosentino, L.; Li, F.; Luo, B.; Alvarez-Carbonell, D.; Garcia-Mesa, Y.; Karn, J.; et al. RNA-Directed Gene Editing Specifically Eradicates Latent and Prevents New HIV-1 Infection. Proc. Natl. Acad. Sci. USA 2014, 111, 11461–11466. [Google Scholar] [CrossRef] [PubMed]
  206. Liao, H.-K.; Gu, Y.; Diaz, A.; Marlett, J.; Takahashi, Y.; Li, M.; Suzuki, K.; Xu, R.; Hishida, T.; Chang, C.-J.; et al. Use of the CRISPR/Cas9 System as an Intracellular Defense against HIV-1 Infection in Human Cells. Nat. Commun. 2015, 6, 6413. [Google Scholar] [CrossRef] [PubMed]
  207. Wang, G.; Zhao, N.; Berkhout, B.; Das, A.T. A Combinatorial CRISPR-Cas9 Attack on HIV-1 DNA Extinguishes All Infectious Provirus in Infected T Cell Cultures. Cell Rep. 2016, 17, 2819–2826. [Google Scholar] [CrossRef] [PubMed]
  208. Zhang, Y.; Yin, C.; Zhang, T.; Li, F.; Yang, W.; Kaminski, R.; Fagan, P.R.; Putatunda, R.; Young, W.-B.; Khalili, K.; et al. CRISPR/gRNA-Directed Synergistic Activation Mediator (SAM) Induces Specific, Persistent and Robust Reactivation of the HIV-1 Latent Reservoirs. Sci. Rep. 2015, 5, 16277. [Google Scholar] [CrossRef] [PubMed]
  209. Saayman, S.M.; Lazar, D.C.; Scott, T.A.; Hart, J.R.; Takahashi, M.; Burnett, J.C.; Planelles, V.; Morris, K.V.; Weinberg, M.S. Potent and Targeted Activation of Latent HIV-1 Using the CRISPR/dCas9 Activator Complex. Mol. Ther. 2016, 24, 488–498. [Google Scholar] [CrossRef] [PubMed]
  210. Klinnert, S.; Schenkel, C.D.; Freitag, P.C.; Günthard, H.F.; Plückthun, A.; Metzner, K.J. Targeted Shock-and-Kill HIV-1 Gene Therapy Approach Combining CRISPR Activation, Suicide Gene tBid and Retargeted Adenovirus Delivery. Gene Ther. 2024, 31, 74–84. [Google Scholar] [CrossRef]
  211. Li, C.; Guan, X.; Du, T.; Jin, W.; Wu, B.; Liu, Y.; Wang, P.; Hu, B.; Griffin, G.E.; Shattock, R.J.; et al. Inhibition of HIV-1 Infection of Primary CD4+ T-Cells by Gene Editing of CCR5 Using Adenovirus-Delivered CRISPR/Cas9. J. Gen. Virol. 2015, 96, 2381–2393. [Google Scholar] [CrossRef]
  212. Ye, L.; Wang, J.; Beyer, A.I.; Teque, F.; Cradick, T.J.; Qi, Z.; Chang, J.C.; Bao, G.; Muench, M.O.; Yu, J.; et al. Seamless Modification of Wild-Type Induced Pluripotent Stem Cells to the Natural CCR5Δ32 Mutation Confers Resistance to HIV Infection. Proc. Natl. Acad. Sci. USA 2014, 111, 9591–9596. [Google Scholar] [CrossRef]
Figure 1. Steps of the HIV life cycle targeted by currently available antiretroviral drugs. HIV binds to the cellular receptor CD4 and a co-receptor (typically CCR5 or CXCR4) to enter a target cell through interaction with the viral gp120 envelope protein. Upon binding, the envelope transmembrane protein gp41 facilitates virus–host cell membrane fusion, triggering release of the viral core into the cytoplasm. A double-stranded DNA copy of the viral RNA is then reverse transcribed around the time it is shuttled into the nucleus, then the HIV integrase enzyme permanently inserts the HIV genome into the host-cell chromosomal DNA. Now stably integrated, the HIV provirus is transcribed into RNA, which is exported to the cytoplasm (in some cases after splicing). After the RNA is translated into proteins, immature virions are assembled and bud through the plasma membrane. The HIV protease enzyme catalyzes virion maturation by cleaving polypeptides within the new virion. Currently approved classes of antiretroviral drugs are shown in blue boxes. HIV latency occurs when the provirus pauses after integration, during which time it expresses little or no RNA and no viral proteins. As there are no currently approved antiretroviral drugs that target an integrated HIV provirus, latently infected cells can persist despite ART (red box).
Figure 1. Steps of the HIV life cycle targeted by currently available antiretroviral drugs. HIV binds to the cellular receptor CD4 and a co-receptor (typically CCR5 or CXCR4) to enter a target cell through interaction with the viral gp120 envelope protein. Upon binding, the envelope transmembrane protein gp41 facilitates virus–host cell membrane fusion, triggering release of the viral core into the cytoplasm. A double-stranded DNA copy of the viral RNA is then reverse transcribed around the time it is shuttled into the nucleus, then the HIV integrase enzyme permanently inserts the HIV genome into the host-cell chromosomal DNA. Now stably integrated, the HIV provirus is transcribed into RNA, which is exported to the cytoplasm (in some cases after splicing). After the RNA is translated into proteins, immature virions are assembled and bud through the plasma membrane. The HIV protease enzyme catalyzes virion maturation by cleaving polypeptides within the new virion. Currently approved classes of antiretroviral drugs are shown in blue boxes. HIV latency occurs when the provirus pauses after integration, during which time it expresses little or no RNA and no viral proteins. As there are no currently approved antiretroviral drugs that target an integrated HIV provirus, latently infected cells can persist despite ART (red box).
Viruses 16 01163 g001
Figure 2. HIV plasma viral loads at different phases of HIV infection. Upon acquisition of HIV, viral loads rapidly increase. After initial viremia is reduced by the adaptive immune response (including HIV-specific CD8+ T cells), infection proceeds to a long and typically asymptomatic period that often lasts 10 years. In this stage, very high levels of virus are produced each day leading to high plasma viral loads, but this is offset by the immune response and regeneration of immune cells by the hematopoietic system. If left untreated, infection will eventually progress to AIDS. This occurs when CD4+ T cell numbers are significantly reduced (<200 cells/microliter blood), and the virus has caused sufficient immunological damage so that efficient immune responses can no longer be mounted. The immunodeficient individual is then vulnerable to many opportunistic infections and diseases. Upon initiating ART treatment, HIV viral loads in plasma are ideally reduced to undetectable levels. For HIV to remain suppressed, ART must be continually taken to prevent viral rebound from latent reservoirs that are not cleared by currently available ART.
Figure 2. HIV plasma viral loads at different phases of HIV infection. Upon acquisition of HIV, viral loads rapidly increase. After initial viremia is reduced by the adaptive immune response (including HIV-specific CD8+ T cells), infection proceeds to a long and typically asymptomatic period that often lasts 10 years. In this stage, very high levels of virus are produced each day leading to high plasma viral loads, but this is offset by the immune response and regeneration of immune cells by the hematopoietic system. If left untreated, infection will eventually progress to AIDS. This occurs when CD4+ T cell numbers are significantly reduced (<200 cells/microliter blood), and the virus has caused sufficient immunological damage so that efficient immune responses can no longer be mounted. The immunodeficient individual is then vulnerable to many opportunistic infections and diseases. Upon initiating ART treatment, HIV viral loads in plasma are ideally reduced to undetectable levels. For HIV to remain suppressed, ART must be continually taken to prevent viral rebound from latent reservoirs that are not cleared by currently available ART.
Viruses 16 01163 g002
Figure 3. Schematic of common humanized mouse models for HIV infection and persistence. HIV does not infect non-modified mice, so common murine models for studying HIV replication, pathogenesis, and cure approaches rely on the use of immunodeficient mice transplanted with human cells and/or tissues (humanized mice). For production of mice with near-complete immune systems, this has historically involved highly immunodeficient recipient mouse strains including the NOD-SCID-common gamma (NSG) or NOD-rag-gamma (NRG), which are either irradiated or treated with busulfan to clear space in the bone marrow, followed by infusion with hematopoietic stem cells. This may be performed alone (hu-CD34) or, in the case of bone marrow/liver/thymus (BLT), performed in conjunction with the transplant of fetal liver and thymus tissue under the kidney capsule to generate a human thymus organoid where T cells can develop on human thymic stroma. More recent advances include the use of recipient mice bearing various important human immune genes that further improve the recapitulation of a working human immune system. For example, DRAGA mice (base strain NRG) expressing human HLAs (allowing optimized antigen presentation) or mice that augment human immune cell repopulation, such as NSG-Tg(IL-15) mice that produce human IL-15, allowing the improved development of NK cells.
Figure 3. Schematic of common humanized mouse models for HIV infection and persistence. HIV does not infect non-modified mice, so common murine models for studying HIV replication, pathogenesis, and cure approaches rely on the use of immunodeficient mice transplanted with human cells and/or tissues (humanized mice). For production of mice with near-complete immune systems, this has historically involved highly immunodeficient recipient mouse strains including the NOD-SCID-common gamma (NSG) or NOD-rag-gamma (NRG), which are either irradiated or treated with busulfan to clear space in the bone marrow, followed by infusion with hematopoietic stem cells. This may be performed alone (hu-CD34) or, in the case of bone marrow/liver/thymus (BLT), performed in conjunction with the transplant of fetal liver and thymus tissue under the kidney capsule to generate a human thymus organoid where T cells can develop on human thymic stroma. More recent advances include the use of recipient mice bearing various important human immune genes that further improve the recapitulation of a working human immune system. For example, DRAGA mice (base strain NRG) expressing human HLAs (allowing optimized antigen presentation) or mice that augment human immune cell repopulation, such as NSG-Tg(IL-15) mice that produce human IL-15, allowing the improved development of NK cells.
Viruses 16 01163 g003
Figure 4. Current HIV cure approaches. General approaches to developing a cure for HIV that are currently being explored within the field include block and lock, provirus editing/silencing, latency reversal and kill augmentation (kick and kill), and stem cell transplantation/gene therapy. Block and lock agents include didehydro-cortistatin A (dCA), JAK-STAT inhibitors, and BRD4 modulators. Common provirus editing/silencing approaches have utilized zinc nuclease fingers (ZNFs), transcription activator-like effector nucleases (TALENs), and clustered regularly interspaced short palindromic repeats (CRISPR). Latency reversal cure approaches have included the use of protein kinase C (PKC) modulators, histone deacetylase inhibitors (HDACis), bromodomain extra-terminal motif (BET) bromodomain inhibitors, and second mitochondrial-derived activator of caspases (SMAC) mimetics. Kill augmentation has been explored with broadly neutralizing antibodies (bnAbs), programmed cell death protein (PD-1) boosted cytotoxic lymphocytes (CTLs), chimeric antigen receptor T and NK cells (CAR-T/CAR-NKs), and bispecific/trispecific antibodies, while modified stem cells with CCR5-∆32 bone marrow transplants have been used to apparently cure HIV in a select few individuals.
Figure 4. Current HIV cure approaches. General approaches to developing a cure for HIV that are currently being explored within the field include block and lock, provirus editing/silencing, latency reversal and kill augmentation (kick and kill), and stem cell transplantation/gene therapy. Block and lock agents include didehydro-cortistatin A (dCA), JAK-STAT inhibitors, and BRD4 modulators. Common provirus editing/silencing approaches have utilized zinc nuclease fingers (ZNFs), transcription activator-like effector nucleases (TALENs), and clustered regularly interspaced short palindromic repeats (CRISPR). Latency reversal cure approaches have included the use of protein kinase C (PKC) modulators, histone deacetylase inhibitors (HDACis), bromodomain extra-terminal motif (BET) bromodomain inhibitors, and second mitochondrial-derived activator of caspases (SMAC) mimetics. Kill augmentation has been explored with broadly neutralizing antibodies (bnAbs), programmed cell death protein (PD-1) boosted cytotoxic lymphocytes (CTLs), chimeric antigen receptor T and NK cells (CAR-T/CAR-NKs), and bispecific/trispecific antibodies, while modified stem cells with CCR5-∆32 bone marrow transplants have been used to apparently cure HIV in a select few individuals.
Viruses 16 01163 g004
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

Chou, T.C.; Maggirwar, N.S.; Marsden, M.D. HIV Persistence, Latency, and Cure Approaches: Where Are We Now? Viruses 2024, 16, 1163. https://doi.org/10.3390/v16071163

AMA Style

Chou TC, Maggirwar NS, Marsden MD. HIV Persistence, Latency, and Cure Approaches: Where Are We Now? Viruses. 2024; 16(7):1163. https://doi.org/10.3390/v16071163

Chicago/Turabian Style

Chou, Tessa C., Nishad S. Maggirwar, and Matthew D. Marsden. 2024. "HIV Persistence, Latency, and Cure Approaches: Where Are We Now?" Viruses 16, no. 7: 1163. https://doi.org/10.3390/v16071163

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