Next Article in Journal
Coat Proteins of the Novel Victoriviruses FaVV1 and FaVV2 Suppress Sexual Reproduction and Virulence in the Pathogen of Fusarium Head Blight
Previous Article in Journal
Comparison of Different HIV-1 Resistance Interpretation Tools for Next-Generation Sequencing in Italy
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Exploring HIV-1 Maturation: A New Frontier in Antiviral Development

1
Department Biology, College of Arts and Sciences, American University, Washington, DC 20016, USA
2
Virus-Cell Interaction Section, HIV Dynamics and Replication Program, Center for Cancer Research, National Cancer Institute, Frederick, MS 21702, USA
3
ViiV Healthcare, 36 E. Industrial Road, Branford, CT 06405, USA
4
District of Columbia Center for AIDS Research, Washington, DC 20052, USA
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Viruses 2024, 16(9), 1423; https://doi.org/10.3390/v16091423
Submission received: 14 August 2024 / Revised: 1 September 2024 / Accepted: 3 September 2024 / Published: 6 September 2024
(This article belongs to the Section Human Virology and Viral Diseases)

Abstract

:
HIV-1 virion maturation is an essential step in the viral replication cycle to produce infectious virus particles. Gag and Gag-Pol polyproteins are assembled at the plasma membrane of the virus-producer cells and bud from it to the extracellular compartment. The newly released progeny virions are initially immature and noninfectious. However, once the Gag polyprotein is cleaved by the viral protease in progeny virions, the mature capsid proteins assemble to form the fullerene core. This core, harboring two copies of viral genomic RNA, transforms the virion morphology into infectious virus particles. This morphological transformation is referred to as maturation. Virion maturation influences the distribution of the Env glycoprotein on the virion surface and induces conformational changes necessary for the subsequent interaction with the CD4 receptor. Several host factors, including proteins like cyclophilin A, metabolites such as IP6, and lipid rafts containing sphingomyelins, have been demonstrated to have an influence on virion maturation. This review article delves into the processes of virus maturation and Env glycoprotein recruitment, with an emphasis on the role of host cell factors and environmental conditions. Additionally, we discuss microscopic technologies for assessing virion maturation and the development of current antivirals specifically targeting this critical step in viral replication, offering long-acting therapeutic options.

1. Introduction

The human immunodeficiency virus (HIV) is in the genus Lentiviridae of the Retroviridae family. Currently, over 39.9 million people live with HIV worldwide, with 1.3 million new infections reported in 2023 (UNAIDS). HIV predominantly infects CD4 T-lymphocytes, inducing a rapid decline in their numbers that, if not treated, eventually leads to acquired immunodeficiency syndrome (AIDS). Effective combination antiretroviral therapy (cART) successfully halts disease progression, reducing AIDS-related deaths by 69% since the peak in 2004 and by 51% since 2010 (UNAIDS). This advancement has allowed many people living with HIV (PLWH) to achieve a normal life expectancy. While cART effectively suppresses virus replication, it does not eradicate infected cells, and HIV persists in latent reservoirs [1,2,3]. As a result, an HIV functional cure has not been achieved with the current cART, necessitating lifelong treatment for PLWH. Ongoing research aims to optimize cART regimens and develop new drugs with improved safety profiles and longer dosing intervals. The current cART regimens for HIV treatment typically consist of different classes targeting the viral enzymes: reverse transcriptase (RT), protease (PR), and integrase (IN). These classes include nucleoside and non-nucleoside reverse transcriptase inhibitors (NRTIs and NNRTIs, respectively), protease inhibitors (PIs), and integrase strand transfer inhibitors (INSTIs), often formulated into single-tablet regimens containing two or three drugs to enhance efficacy, minimize resistance, and improve patient adherence (Figure 1C,E). In addition to antiretroviral drugs targeting viral enzymes, current cART regimens may also include entry inhibitors and fusion inhibitors, which block the virion from binding to and fusing with target cells [4] (Figure 1A,B). Fusion inhibitors, in particular, belong to a class of antiretrovirals that could be used in HIV pre-exposure prophylaxis (PrEP) [5], enabling a comprehensive approach to HIV treatment and prevention. Although most current cART regimens are designed for once-daily oral dosing, recently developed long-acting antivirals are designed to be administered by injection once every few months [6]. The first and currently only long-acting regimen, Cabenuva, a combination of an INSTI (cabotegravir: CAB) and an NNRTI (rilpivirine: RPV), is administered as a once-monthly or bi-monthly HIV treatment and was approved by the FDA in 2021 [7]. Following Cabenuva, Lenacapavir (LEN) was approved by the FDA in December 2022 as a first-in-class antiviral and has the potential for applications in long-acting regimens. LEN is a capsid inhibitor (CAI), a new class of antivirals that target multiple stages of the viral replication cycle, including viral maturation (Figure 1D,H,K and Table 1) [8].
HIV-1 Gag is an essential protein involved in the formation of virus particles. It is composed of four major domains: matrix (MA), capsid (CA), nucleocapsid (NC), and p6, along with the short spacer peptides SP1 and SP2 (Figure 2A). The production of infectious virus particles is mediated by the enzymatic activity of the viral PR released from the Gag-Pol polyprotein in the progeny virion [9]. Upon the budding of progeny virions from virus-producer cells, the proteolytic processing of the Gag polyprotein by PR is initiated. This enzymatic process results in important structural rearrangements and the formation of each mature Gag protein (Figure 1K) [9]. MA plays a crucial role in the incorporation of envelope glycoproteins (Env) into the budding virions [10]. The N-terminal domain of MA in the Gag polyprotein is myristoylated and binds to the plasma membrane, which is anchored by electrostatic interactions in its highly basic region (HBR) to facilitate progeny virion assembly [11,12]. CA assembles into approximately 250 hexameric and exactly 12 pentameric units, forming the conical fullerene core that encapsulates the viral genomic RNA (gRNA). This structure is vital for reverse transcription, nuclear import, and uncoating post-infection [13]. NC is responsible for packaging the viral gRNA into the progeny virions [14,15]. NC also facilitates reverse transcription in the core upon its release to the cytoplasm post-fusion by inducing the annealing of the transfer RNA (tRNA) primer to the viral gRNA to initiate the synthesis of double-stranded viral DNA [15,16]. Lastly, p6 recruits the endosomal sorting complex required for transport (ESCRT) machinery, which is necessary for virion budding and subsequent release, and directs the incorporation of the Vpr protein into the virion [17].
In this review article, we discuss the importance of virion maturation in the viral replication cycle and how host factors affect the HIV maturation process. We also provide an overview of several classes of inhibitors targeting virion maturation, their mechanisms of action, the current state of their clinical development, and the next generation of technologies to further understand these mechanisms.

2. Viral Replication Cycle

HIV primarily targets CD4 T-cells by utilizing its surface protein gp120 to attach to the host cell [18]. The virus binds to the CD4 receptor and a co-receptor, either the C-C chemokine receptor type 5 (CCR5) or C-X-C chemokine receptor type 4 (CXCR4), on the T-cell surface (Figure 1A) [18]. Following virion attachment, the viral lipid envelope fuses with the target cell plasma membrane, releasing the core harboring the viral gRNA into the host cell (Figure 1B). Inside the core, RT initiates the synthesis of double-stranded viral DNA (Figure 1C) [19]. The core travels into the nucleus through the nuclear pore and uncoats upon the completion of viral DNA synthesis within the nucleus (Figure 1D,E). Eventually, the newly synthesized viral DNA integrates into the host genome, a process catalyzed by the IN complex (Figure 1E) [18,20,21]. The viral genome integrated into the DNA of a host cell is referred to as a provirus. The host RNA polymerase II transcribes the proviral DNA. This transcription is significantly enhanced by the HIV-1 Tat protein through its binding to the transactivation response (TAR) region of the viral messenger RNA (mRNA), eventually activating HIV-1 transcriptional elongation (Figure 1F) [22,23]. Tat also recruits the host transcription elongation factor b (p-TEFb) complex to the HIV-1 long terminal repeat (LTR) promoter region, increasing the efficiency of transcription and thereby promoting the production of viral RNA (vRNA) [22]. Full-length unspliced vRNA transcripts initially undergo splicing to generate fully spliced RNA species that encode solely Tat, Rev, and Nef [24,25,26,27]. These transcripts enhance the subsequent singly or partially spliced RNA species encoding Vpu: Env, and the other accessory proteins, Vif and Vpr. Some vRNA transcripts remain unspliced and serve as the viral mRNA template for the synthesis of the Gag and Gag-Pol polyproteins or as the viral genome that is packaged into nascent progeny particles (Figure 1I) [18,26,28,29]. Gag and Gag-Pol polyproteins assemble at the plasma membrane, where the immature HIV-1 particle buds from the host cell while acquiring a portion of the host cell membrane and incorporating the viral envelope (Env) glycoproteins (Figure 1G–J) [18,30]. The virus becomes infectious after maturation, a process that involves the PR-mediated cleavage of Gag and Gag-Pol polyproteins to form a mature HIV-1 core (Figure 1K) [18,31,32]. Virion maturation is the final stage in the viral replication cycle when various biochemical and morphological changes occur to ensure the transition of immature, noninfectious particles into mature, infectious viruses (Figure 1K) [32].

3. Molecular Mechanisms of HIV-1 Maturation

3.1. Fundamentals in HIV-1 Maturation

The translational ratio of Gag to Gag-Pol polyproteins from the same full-length viral mRNA is determined by a ribosomal frameshift during translation [33,34]. Approximately 5–10% of the ribosomes that initiate translation of the full-length vRNA undergo a −1 frameshift, leading to the translation of the Gag-Pol polyprotein instead of the Gag polyprotein [34,35]. While the Gag polyprotein encompasses the MA, CA, NC, and p6 domains, the Gag-Pol polyprotein lacks the p6 domain due to the ribosomal frameshift. However, the Gag-Pol polyprotein additionally includes the Pol domains, PR, RT, and IN, which are crucial for viral replication and integration into the host genome [36,37]. The Gag polyprotein undergoes cleavage in a stepwise manner, starting between SP1 and NC, followed by MA-CA and SP2-p6 cleavages, and finally removing SP1 and SP2 short peptides from CA and NC, respectively (Figure 2A) [38]. The Gag-Pol polyprotein is produced through a frameshifting event at the NC/SP1 boundary, translating the p6 domain in an alternative reading frame (Transframe/TF) [39]. The initial cleavage events separate the Gag and Pol regions by cleaving between TF and PR, a process mediated by the poorly active PR precursor (Figure 2B) [40,41]. This step releases the protease from the precursor, allowing the ordered processing of the Gag and Pol polyprotein. This proteolytic processing of the Gag and Gag-Pol polyproteins induces significant architectural changes, transforming the immature virion with a dense outer protein layer (the immature Gag lattice) into a mature virion with a defined conical core (Figure 1K and Figure 3) [42]. The mature CA assembles into a closed lattice of hexameric and pentameric units, which is essential for protecting the viral gRNA and ensuring the productive infection of new host cells (Figure 1A,B and Figure 3) [36,37].

3.2. Host Proteins in HIV-1 Maturation and Core Association

Several host factors play various roles in facilitating or inhibiting HIV-1 virion maturation and core formation. Cyclophilin A (CypA) is a peptidyl-prolyl isomerase that binds to the HIV-1 CA protein via a proline-rich loop, comprising CA residues 85 to 93, with glycine and proline at positions 89 and 90, respectively, serving as key binding motifs for CypA [50,51,52]. While CypA is packaged into the budding virion to support the generation of infectious viruses [17,52,53,54], its primary function is mediated by the CypA expressed in target cells, interacting with the invading HIV-1 CA core. This interaction, occurring with multiple loops on the capsid exterior in the target cell cytoplasm, influences HIV infection in a cell-type-specific manner [55,56,57,58], likely by affecting capsid stability and uncoating [59,60]. In myeloid cells, inhibiting this interaction via cyclosporin A (CsA) or mutating the CypA binding loop on CA leads to HIV-1 inducing immune responses, possibly due to premature uncoating and viral DNA release in the cytoplasm [61]. The exact molecular mechanisms remain unclear but may involve dynamic allostery [62]. Additionally, humans encode a large family of tripartite motif-containing (TRIM) proteins that recognize the CA lattice once the core is released into the cytoplasm of the target cell [63]. This recognition stimulates restriction factors targeting the viral core [64]. Although human TRIM5α has some activity against HIV-1, it is not as potent as TRIM5α from other primates, such as rhesus macaques [65]. The human TRIM5α protein has a limited ability to recognize and bind to the HIV-1 core, which is one reason why HIV-1 infection can successfully be achieved in human cells [66]. In contrast, TRIM5α from rhesus macaques can strongly restrict HIV-1 by recognizing the CA core more effectively, leading to a more robust antiviral response [67,68]. An alternative splicing form of the chimeric transcript encoding the TRIM motif fused to a C-terminal CypA domain (TRIMCyp) also shows antiviral activity by binding to the proline-rich CypA-binding loop on the HIV-1 CA [59,69,70]. However, TRIMCyp is not naturally expressed in humans [71,72]. In addition, the restriction activity facilitated by rhesus TRIM5α along with TRIMCyp also improves CD8 T-cell-mediated inhibition of HIV-1, indicating that there is a direct link between rhesus TRIM5α and CD8 T-cell response to infection [73]. The human TRIM5α may have lost these activities through evolution, or HIV-1 may have evolved to escape human TRIM5α-mediated restriction [58,73,74,75].
Cleavage and polyadenylation specificity factor 6 (CPSF6) is a cellular factor that plays a role in the HIV-1 replication cycle. CPSF6 contains an RNA-binding domain that enables its interaction with RNA substrates and other protein components of the CPSF complex, thereby coordinating RNA processing events [13]. CPSF6 shuttles between the nucleus and cytoplasm, displaying dynamic functions in RNA metabolism and gene regulation. In response to viral infections, CPSF6 binds to the HIV-1 CA hexamer of the core to facilitate its nuclear import, which is crucial for infection (Figure 1D and Figure 3) [46,47]. Recent studies have shown that CPSF6 not only impacts the efficiency of viral DNA integration but also influences the selection of integration sites within the host genome [76]. This can affect the stability of the provirus and its susceptibility to transcriptional silencing or activation.
The HIV capsid core is trafficked toward the nucleus (retrograde transport) through the host cell’s microtubule network via the coordinated action of dynein and kinesin motor proteins, which facilitate the directional movement of the capsid core toward the nuclear pore complex (NPC) (Figure 1C) [77,78,79,80]. Microtubule-associated proteins, MAP1A and MAP1S, interact with the CA, playing a role in linking the CA protein to the host cell’s microtubule network [81]. Protein bicaudal D2 (BICD2), a dynein adaptor protein, binds to HIV-1 CA and recruits dynein and dynactin to form a protein complex, facilitating retrograde transport [82,83]. Fasciculation and elongation factor zeta 1 (FEZ1), a kinesin-1 adaptor protein, promotes retrograde transport by bridging kinesin-1 and the CA core through its interaction with the CA hexamer [84,85].
Nup358, located on the cytoplasmic filaments of the NPC, interacts with the HIV-1 CA to facilitate the docking of the viral core at the NPC (Figure 1D). This process is aided by kinesin-1 motor proteins such as KIF5B, which transport the viral core towards the nucleus [86]. Nup358 binds to the CypA binding loop of the HIV-1 CA via its C-terminal cyclophilin domain [87,88]. CPSF6 has been reported to enhance the interaction between Nup358 and the HIV-1 core [86]. Nup153, an essential component of the nuclear basket of the NPC, also interacts with CA hexamers, specifically with a hydrophobic pocket between adjacent CA monomers through its phenylalanine-glycine (FG) motifs (Figure 3) [47,48,89], while a unique triple-arginine (RRR) motif targets the tri-hexamer interface [90]. These interactions are essential for stabilizing the HIV-1 core and facilitating its passage through the NPC [91]. At the NPC, CPSF6 and Nup153 collaborate to facilitate the translocation of the core into the nucleus. CPSF6 assists in releasing the viral core from the NPC following the initial docking mediated by Nup153 (Figure 1D) [92].
Human Myxovirus resistance protein 2 (Mx2/MxB), a dynamin-like large GTPase of the dynamin superfamily, was initially identified as a regulator of cell-cycle progression and cytoplasmic-nuclear transport [93]. MxB also inhibits HIV-1 infection by blocking several stages of the virus replication cycle, most notably nuclear import [91,94,95,96,97,98,99,100], through binding to Nup358, masking its interaction with the HIV-1 core and, thus, impeding the nuclear import of the viral core [101].

3.3. Host Metabolites in HIV-1 Maturation and Core Association

HIV-1 can also exploit host cell metabolites for its replication [102,103], including inositol hexakisphosphate (IP6), a cellular metabolite derived from inositol, through a series of phosphorylation reactions, which is involved in both HIV-1 assembly and maturation [104]. Inositol undergoes phosphorylation by various kinases to produce inositol monophosphate (IP1), inositol bisphosphate (IP2), and inositol triphosphate (IP3). Inositol polyphosphate multikinase (IPMK) then catalyzes the conversion of IP3 to inositol tetrakisphosphate (IP4) and IP4 to pentakisphosphate (IP5). Ultimately, IP5 2-Kinase (IPPK) phosphorylates IP5 to produce IP6, which promotes the assembly of both the immature virus particle and the formation of the fullerene core during virus maturation [46,105]. In addition, IP6 promotes the assembly of the immature Gag lattice by binding to two lysine rings in a six-helix bundle that forms at the CA-SP1 junction during particle assembly, thereby driving the incorporation of IP6 into the mature particle [105,106,107,108]. This interaction aids in the cleavage of the Gag polyprotein by the PR and stabilizes the CA structure, thus influencing the maturation of the virus and its ability to infect new cells [109,110,111,112]. During virus maturation, IP6 binds the electropositive pore at the center of the mature CA hexamer and pentamer formed by arginine at position 18 (Arg18) and lysine at position 25 (Lys25) (Figure 3) [112,113]. In the absence of IP6, IP5 also interacts with the Gag polyprotein and is packaged into the budding virion at similar levels as IP6 [108,111,114]. IP5 is coordinated at the center of the HIV-1 hexamer through a ring of Arg18 during maturation (Figure 3) [46,105]. IP5 is located at the center of the six-fold axis, resulting in the averaging of six equivalent binding positions due to symmetry. Unlike IP6, which is less planar because of its axial phosphate, the density of IP5 is much clearer, allowing it to be distinctly positioned in a parallel stacking arrangement with the Arg18 ring (Figure 3). The additional axial phosphate in IP6 would be oriented away from the Arg18 ring, leading to the substitution of IP5 for IP6 without reducing infectivity, which is likely because the interaction is primarily driven by the equatorial phosphates, particularly when the ligand is in a planar conformation [105,108]. FEZ1, as discussed in Section 3.2, competes with IP6 through its interaction with the ARG18 ring of the CA hexamer (ref.). Viral particles produced without sufficient IP6 are noninfectious because CA assembly is inefficient, resulting in deficient core formation [105,115]. Furthermore, IP6 stabilizes the CA core in the target cell, thereby promoting the completion of viral DNA synthesis inside the core [108,116]. Overall, IP6 is an essential element in generating infectious virions by supporting immature particle assembly, Gag and Gag-Pol polyprotein cleavage, and conical core formation [33,115].

3.4. Host Membrane Lipid Rafts in HIV-1 Maturation

The composition of the cell membrane plays a crucial role in cellular processes such as endocytosis and the movement of materials into the cell [117]. The plasma membrane contains a variety of lipids distributed asymmetrically across the bilayer [118]. Phosphatidylinositol-4,5-bisphosphate [PI(4,5)P2] plays a key role in the recruitment of Gag to the plasma membrane [119]. HIV-1 particles bud from lipid rafts, which are microdomains in the plasma membrane of infected cells rich in cholesterol, sphingolipids, and glycolipids [86]. Therefore, the viral membrane is enriched in cholesterol, sphingomyelins (SMs), glycosphingolipids, and phosphoinositides such as PI(4,5)P2 [117]. Neutral sphingomyelinase 2 (nSMase2), an enzyme that generates the bioactive lipid ceramide through the hydrolysis of membrane SMs, plays a critical role in the late stages of HIV-1 maturation and replication [120,121]. nSMase2 regulates the size and stability of membrane microdomains and interacts with HIV-1 Gag to hydrolyze SMs into ceramide, which is necessary for virion maturation [122]. The inhibition of nSMase2 in HIV-1 producer cells results in the production of immature HIV-1 particles with increased SMs and a reduced ceramide content, resulting in defective Gag polyprotein processing and impaired viral envelope formation. This, in turn, severely reduces infectivity [122].

3.5. Environmental Factor in HIV-1 Maturation

Environmental factors, such as pH, can contribute to the maturation efficiency of virions. The proteolysis of the Gag polyprotein in virus-like particles (VLPs) by exogenously added recombinant PR has been found to be optimal at pH 6.0 to 7.0 [123], while purified PR is poorly active at neutral pH [124,125]. The internal pH of the HIV-1 virion is generally predicted to be slightly acidic, ranging from pH 6.0 to 7.0 [126]. This pH range ensures the proper processing of the Gag polyprotein as it is optimal for HIV-1 PR activity. Additionally, the cell entry efficiency of enveloped viruses is affected by pH [127], and most enveloped viruses are inactive and unable to invade target cells under acidic conditions of pH 5.0 to 6.0 [128]. Therefore, HIV-1 virion maturation may be controlled through the pH-dependent activation and deactivation of PR.

4. Tools for Assessing Virion Maturation

Although there are many assays for assessing HIV-1 maturation, including western blotting, mass spectrometry, flow cytometry, RT- and PR-based assays, and viral infectivity assays, microscopy is the most utilized and quantitative method. In this section, we specifically discuss the approaches for studying HIV-1 maturation using electron microscopy (EM) and fluorescence microscopy.

4.1. Electron Microscopy

Microscopy is a critical technique in the analysis of viruses due to its ability to reveal events in the virus replication cycle, even with the relatively small size of virions [129]. Transmission EM (TEM) is a widely used technique due to its ability to determine the number of virions in a sample and characterize their morphology at nanometer-scale resolution by transmitting electrons through a specimen embedded in plastic [130,131,132]. TEM has been applied to HIV research to visualize the morphogenesis and maturation of the virus (Figure 4). TEM can visualize the rearrangement of structural proteins as the virus transforms into a mature, infectious form. Additionally, TEM has been used to evaluate the efficacy of inhibitors that block HIV-1 maturation [32,129,133,134]. TEM contributes to studying the effectiveness of these inhibitors by directly revealing their effect on virion morphology. Insights into the localization of the Gag polyprotein precursor in relation to viral gRNA have also been uncovered by TEM [135]. Cryogenic EM (Cryo-EM) is used to determine high-resolution structures of biological macromolecules in three dimensions (3D) [135]. This technique uses frozen macromolecules fixed in ice to determine their visual projection from different directions. The samples are then processed in two dimensions (2D) before being reconstructed into a 3D form for analysis [135]. Cryo-EM can capture multiple frozen particles in different orientations, providing high-resolution images of individual macromolecular complexes with atomic-level details. On the other hand, cryo-electron tomography (Cryo-ET) takes sequential images of a single particle as it is rotated or tilted along an axis. These series of images are aligned to computationally reconstruct a 3D representation, or tomogram, of the particle.
Cryo-EM/ET has also been used to investigate how viruses operate in the context of viral replication and maturation [136,137,138]. It has also been applied to research surrounding HIV-1 to explore how the virus assembles inside host cells and how different intermediate states form [100]. It has also provided structural insights into the transport of the cone-shaped CA core through the nuclear pore [139] and the variation and positioning of HIV-1 Env on the Gag lattice [140].

4.2. Fluorescence Microscopy

Fluorescence microscopy is a key technology adapted to explore the structure of cells, tissues, and other biological components [141]. Super-resolution microscopy techniques surpass the diffraction limit of traditional fluorescence microscopy, enabling the visualization of structures at a nanometer scale, thereby allowing a detailed examination of biological processes at a molecular level [142]. However, even with super-resolution techniques, fluorescence microscopy does not have sufficient resolution to reveal the structural changes that occur in the virion during maturation.
The Förster resonance energy transfer (FRET) principle is a specific feature of fluorescence microscopy, where energy is transferred from an excited fluorophore (the donor) to another fluorophore (the acceptor) [143]. This process is distance-dependent and occurs through an intermolecular long-range dipole–dipole coupling [144]. FRET is an effective technique for investigating various biological processes and has been applied to HIV studies due to its high sensitivity to changes in molecular proximity [134,144,145]. An automated method for detecting PR activity in single virions using fluorescence microscopy was developed [146]. This system employs tandem fluorescent proteins, separated by HIV-1 PR cleavage sequences, fused in-frame to the N-terminus of HIV-1 Vpr. This method utilizes an FRET pair of fluorescent proteins to tag Vpr, providing a convenient reporter for viral PR activity, indicated by a loss of FRET due to the separation of the donor and acceptor proteins. This single-virion FRET-based technology demonstrates that the FRET levels and the relative intensities of the donor and acceptor signals in individual virions can effectively differentiate between immature and mature virions. Although it results in a modest reduction in infectivity, this bifunctional marker may reduce maturation efficiency. Consequently, it serves well for the qualitative differentiation between mature and immature virions but is not suitable for quantitative assessment. Another fluorescence microscopy tool utilizing the FRET principle (HIV-1 Gag-iFRET) was recently used to distinguish between mature and immature virions (Figure 4) [31,134]. This was accomplished by inserting the efficient intramolecular FRET pair combination, ECFPΔC11-cp173Venus [147,148], into the Gag polyprotein between the MA and CA domains, surrounded by an HIV-1 PR cleavage site, based on the HIV-1 Gag-iGFP construct [149] (Figure 2C). In the HIV-1 Gag-iFRET progeny virions, the Gag polyprotein and FRET pair are cleaved by PR during maturation. Separation of the FRET pair results in a reduction in the FRET signal (Figure 4). Alternatively, if maturation is blocked with the treatment of PIs, the FRET pair remains uncleaved and emits an FRET signal [150,151]. This HIV-1 Gag-iFRET virus did not lose any infectivity compared to the unlabeled virus [134] and directly detected the PR-mediated cleavage of Gag and Gag-Pol polyproteins, making it suitable for quantitative analysis without any artificial disruption of maturation. This new fluorescence microscopy-based system showed that approximately 80% of virions completed the cleavage of the MA-CA junction in the Gag and Gag-Pol polyproteins, indicating that they are predicted to be mature virions. This correlated well with the range of virion maturation ratios assessed by TEM results, which vary from 50% to 99%, depending on the laboratory’s manual procedures for determining the morphology of virions [49,55,152,153,154]. Based on a computational analysis of fluorescence emitted at the single-virion level, this FRET-based method enables an unbiased quantification of the proportion of Gag and Gag-Pol polyprotein cleavage by PR inside progeny virions. HEK293T cells, derived from human kidney cells, are typically used for in vitro virus production and the characterization of virion maturation [155]. However, it has recently been discovered using the HIV-1 Gag-iFRET system that the maturation rate of viruses produced from HEK293T cells is relatively higher than in other cells that are more biologically relevant to HIV production [31]. The Gag processing efficiency in viruses from T-cell lines, such as Jurkat cells, was significantly lower than that of viruses from HEK293T cells (68.7% vs. 81.8%, respectively). This difference was also correlated with the virus infectivity [31]. The virus morphological analysis produced by another T-cell line, MT-4 cells, showed that 62% of the virions were mature [112], which aligned with the results from the HIV-1 Gag-iFRET detection system in Jurkat cells. There is limited information about the efficiency of virion maturation in primary T cells, so further characterization and a careful quantitative analysis of maturation efficiency in relevant cell types is warranted. It is also important to note that Gag processing efficiency is an indirect measure of virus maturation efficiency. Therefore, a combination of techniques, including direct morphological analysis by EM and quantitative fluorescence microscopic assays, will comprehensively address this important matter.

5. Influence of Virion Maturation on HIV-1 Env Function

The HIV-1 Env mediates the fusion between viral and cellular membranes (Figure 1A,B). Upon the binding of gp120 to CD4, conformational changes occur in the Env trimer, exposing the binding sites for co-receptors (CCR5 or CXCR4) on gp120. The subsequent interaction of gp120 with the co-receptor facilitates the insertion of the fusion peptide of gp41 into the target cellular membrane. The fusion between viral and cellular membranes is mediated by the formation of an antiparallel 6-helix bundle involving heptad repeat 1 and 2 (HR1 and HR2) in the ectodomain of gp41 [156]. Single-molecule FRET (smFRET) analysis has revealed that Env exists as metastable trimers on virions and infected cells, sampling at least three intramolecular conformations: state 1 (pre-triggered, closed), state 2 (functional, intermediate), and state 3 (fully CD4-bound, open) [157,158,159,160].
Env is incorporated into virions as a trimer of gp120-gp41 heterodimers [161]. The MA domain of Gag plays a pivotal role in Env incorporation during viral assembly [10]. The MA domain forms a trimeric arrangement [162] and multimerizes into a hexameric lattice of trimers on PI(4.5)P2-containing membranes, as well as within immature and mature virions [42,163]. Several studies indicate that MA trimerization is critical for efficient Env incorporation [164,165]. Although direct structural evidence for the interaction is still lacking, biochemical studies suggest that the long gp41 cytoplasmic tail (CT), which is approximately 150 amino acids in length, interacts with the central gap of the MA lattice [164,165]. Env in immature virions is fusion-incompetent, and the PR-mediated Gag cleavage at the SP1-NC junction, which releases the downstream SP1 from MA, is required to activate fusogenicity [166,167]. The suppression of fusogenicity in immature virions can also be rescued by the deletion of the gp41 CT [166,167], suggesting that interactions between the gp41 CT and MA domain of uncleaved Gag inhibit premature fusion. Fluorescence nanoscopic analyses have shown that Gag cleavage by PR induces the redistribution of Env trimers (Figure 1K and Figure 4). Env trimers are dispersed on immature virions but are clustered on mature virions, a process that is dependent on the gp41 CT [168]. Recent cryo-EM analysis has revealed that gp120 binding to CD4 further clusters Env-CD4 complexes in the membrane–membrane interface, with the clustering influenced by viral maturation status [169]. Despite their dispersion, Env trimers on immature virions can adopt an open conformation upon CD4 binding, similar to those on mature virions [170], suggesting that maturation affects intermolecular Env interactions rather than intramolecular Env conformations. However, further analysis is required to understand the impacts of Gag maturation on the Env conformation. Interestingly, immature virions exhibit greater rigidity than mature virions, and virion stiffness is also regulated by the gp41 CT and is correlated with viral infectivity [171]. The composition of the viral membrane also influences the distribution and function of Env trimers. A recent study demonstrates that the gp41 CT directly binds to viral envelope cholesterol [172]. This interaction is required for the Env clustering on the viral membrane and subsequent fusion events [172,173,174]. Several studies have reported that HIV-1 virions typically contain, on average, eight to ten Env trimers, which is a lower density compared to other enveloped viruses such as measles and influenza viruses [175,176]. The low density of Env complexes on HIV-1 particles may be a strategy for evading immune detection. It has been predicted that multiple Env-receptor interactions are required for viral fusion (Figure 1A,B) [177], underscoring the importance of Env clustering for viral fusion. As described above, HIV-1 MA forms hexamers consisting of six trimer units in both immature and mature virions. However, the arrangement of the MA lattice differs between these states, suggesting that Gag cleavage by PR triggers conformational rearrangements of the MA lattice [42]. These findings suggest that structural rearrangements of the MA lattice may influence virion stiffness and the distribution of Env trimers, which are crucial for facilitating effective fusion events.

6. Targeting Maturation for Therapeutic Intervention: Small Molecule Discovery, Mode of Action, and Clinical Development

The rationale of the pharmaceutical targeting of virion maturation has been explored for over three decades, with multiple successful approaches leading to the identification of different classes of small molecule inhibitors. Some of these classes, such as PIs, were initially marketed during the second decade of the HIV pandemic, while others, including CAIs, maturation inhibitors (MIs), and allosteric integrase inhibitors (ALLINIs), have been developed in recent years. CAIs exhibit the highest antiviral potency and the longest-acting therapeutic profiles achieved to date.

6.1. Protease Inhibitors

The first class of small molecule inhibitors of virus maturation was developed early in the HIV pandemic, with a mechanism that directly blocks the catalytic activity of PR. PIs played a critical role in the development of antiretroviral therapies and were included in three-drug regimens, typically with two NRTIs as the backbone. Soon after the discovery of the PR structure and its function in virus maturation and infectivity, multiple active-site PIs were developed as antiretrovirals, with saquinavir (SQV) being the first approved member of this class in 1995 (Table 1) [178]. A seminal work demonstrated that a single substitution of the aspartic acid residue at position 25 with asparagine (D25N) in the PR catalytic site resulted in the production of immature virions that contained only uncleaved Gag polyprotein while preserving PR expression [179]. These results, along with the availability of crystal structures of PR, demonstrated that the catalytic site of PR was a viable target for the development of inhibitors [180,181]. Seven amino-acid peptides corresponding to specific sequences around Gag cleavage sites are sufficient for the proteolytic activity (Figure 2), suggesting that small peptidic substrates could be investigated as PI candidates [182,183]. Most of the approved PIs were then conceived to mimic the PR substrate and were, for this reason, named peptidomimetics. These inhibitors contain a hydroxyl group that interacts with the carboxyl group of residues in the active site of PR. The inhibition of the catalytic activity of PR blocks Gag and Gag-Pol polyprotein precursor processing, resulting in the formation of immature virus particles that are not infectious. Long-term treatment with first-generation PIs revealed limitations due to the high pill burden, frequent adverse events, and potential for drug–drug interactions. These interactions arise because PIs are substrates for the Cytochrome P450 3A4 (CYP3A4) enzyme, which metabolizes many other drugs in the liver, and, thus, PIs are co-administered with pharmacokinetic boosters such as ritonavir (RTV) and cobicistat (COBI) [184]. RTV, initially approved as a PI, is now used in lower doses primarily for its potent CYP3A4 inhibition [185]. COBI, a structural analog of RTV, lacks antiviral activity but also serves as a strong CYP3A4 inhibitor [186]. Both RTV and COBI increase the plasma concentrations of other antiretrovirals, including PIs, thereby enhancing their efficacy, reducing effective doses, and decreasing pill burden. On the other hand, inhibiting CYP3A4 can lead to increased drug levels in the blood, raising the risk of toxicity and adverse drug interactions by slowing the metabolism and clearance of many medications. Second-generation PIs, such as darunavir (TMC114), which was the last PI approved in 2006 [187] (Table 1), were developed to address some of these issues, offering a greatly improved virologic profile with high genetic barriers to resistance and better tolerance, even though they remain a substrate for CYP3A4. Taken together, while PIs were once central to HIV therapy, they are now less commonly used as first-line regimens due to the side effects, complex dosing requirements, and availability of newer antiretroviral classes with improved safety and efficacy profiles. Continued research is essential for developing next-generation PIs with safer drug profiles [188,189].

6.2. Maturation Inhibitors

MIs block particle maturation by binding at the CA-SP1 junction of Gag, thereby preventing CA-SP1 cleavage, which, as mentioned above, is the last cleavage event in the Gag processing cascade and is critical for liberating mature CA (Figure 2). Thus, this class of inhibitors does not block the intrinsic enzymatic activity of PR but rather prevents the enzyme from accessing its substrate. A screening of natural products derived from leaf extracts of Syzygium claviflorum, a tree in the Myrtaceae family native to Australia and Asia, found that a known triterpenoid, betulinic acid, and its derivatives had antiviral activity against HIV-1 in assays conducted with H9 lymphocytic cells. Although with a half maximal effective concentration (EC50) in the single-digit micromolar range and a modest therapeutic index, the activity of those compounds seemed to be specific against HIV-1 replication [190]. Subsequent medicinal chemistry efforts led to the discovery of the dimethyl succinyl derivative of betulinic acid, known variously as YK-FH312, PA-457, DMB, and now bevirimat (BVM) [191,192,193] (Table 1). The time of addition assays demonstrated that BVM inhibits the late stages of virus replication [191]. A seminal study established that the mode of action of BVM is via inhibiting the conversion of the capsid precursor (CA-SP1 or p25) to the mature form (CA or p24), which induces the formation of virus particles that have an aberrant morphology and are not infectious [193]. Importantly, an early report showed that in vitro resistance selection with BVM identified a single mutation at the N-terminus of SP1, which flanks the CA-SP1 cleavage site [136,194,195,196,197], suggesting that the compound binds at the CA-SP1 junction. Further resistance selections and studies using single amino acid deletions spanning the entire CA-SP1 boundary domain expanded our understanding of the BVM binding pocket, and other critical residues were identified in the C-terminus of CA [198,199,200,201]. By using photoaffinity-labeled analogs of BVM, it was shown that this series targets the Gag major homology region (MHR), which is critical for virus assembly and mediates the binding of Gag to IP6 that stabilizes the CA-SP1 lattice [202]. Overall, these studies confirmed that the target site for BVM resides in the Gag CA-SP1 region. With regards to the antiviral spectrum, BVM was found to be active against a panel of wild-type (WT) and drug-resistant isolates [193] but cannot inhibit simian immunodeficiency virus (SIV) [201,203]. Biochemical assays using recombinant monomeric Gag protein failed to recapitulate BVM activity, suggesting that the BVM binding site is present only in multimeric assembled Gag proteins [193]. Furthermore, it was shown by liquid chromatography–mass spectrometry that the compound can bind immature particles but not bind either immature virions containing BVM resistance mutations or mature virions [204]. Recent insights into BVM-mediated antiviral activity revealed its interaction with a six-helix bundle that forms at the CA-SP1 junction in the immature Gag lattice during particle assembly. The current experimental model is based on a dynamic equilibrium between a stable and a disordered state [32]. Structural information on BVM indicates that the molecule binds within the six-helix bundle, thereby stabilizing it. Since the CA-SP1 cleavage site is buried within the bundle, its stabilization by BVM prevents PR from accessing the CA-SP1 cleavage site [106,136,193,194,196,197,205,206]. As mentioned above, IP6, which also binds and stabilizes the six-helix bundle, is an integral component of proper virus particle maturation. The IP6 binding site in the immature Gag lattice contains two rings of lysine side chains sitting in the six-helix bundle above the BVM binding site. BVM and IP6 have been shown to simultaneously bind to the six-helix bundle [113], restricting the mobility of residues in the CA-SP1 junction. This suggests a cooperative mechanism that restricts PR access to the CA-SP1 cleavage site [205]. Several reports investigating the interdependence between IP6 and MIs suggest an intriguing hypothesis that inhibitors designed to compete with IP6 binding to the bundle may enhance antiviral activity [113].
Five key MIs have undergone evaluation in Phase 1-2b clinical trials. BVM showed a decline in HIV-1 RNA of nearly 0.45 log10 c/mL in 10 days in Phase 2a [207]. However, its potency was compromised by several polymorphisms in the CA-SP1 junction region. Another MI, GSK2838232 (Table 1), demonstrated a more potent antiviral decline of about 1.7 log10 c/mL [208]. On the other hand, its PK properties required boosting with RTV, which led to the termination of its development. The MI GSK3532795 (Table 1) showed a potent antiviral response regardless of any polymorphism and did not require boosting with RTV [209]. Thus, GSK3532795 has been evaluated in two Phase 2b studies; the dose range-finding study showed a Week 24 efficacy rate of approximately 82% on a backbone of two NRTIs in treatment-naïve adults with HIV-1 [210]. This compound was terminated for high rates of resistance, both to the NRTI backbone and MI itself, and high rates of gastrointestinal (GI) intolerability (e.g., abdominal pain and diarrhea). The MI GSK3640254 (Table 1) was studied through two Phase 2b studies in treatment-naïve adults with HIV-1, one in combination with two NRTIs and another solely administered with an INSTI, dolutegravir (DTG). Notably, both trials showed similar clinical profiles to the reference regimen: similar rates of virologic responders at Week 24 (without any treatment-emergent resistance in either trial) and acceptable safety/tolerability profiles [211]. Ultimately, it was determined that an intended Phase 3 fixed-dose combination of GSK3640254 and DTG would not be superior to the existing two-drug daily oral regimens. Thus, GSK3640254 was not advanced into Phase 3. At the 2024 Conference on Retroviruses and Opportunistic Infections (CROI), an additional MI, VH3739937 (VH-937, also known as GSK3739937), was reported to have antiviral properties against HIV-1 strains with diverse Gag sequences and polymorphisms generally conferring resistance to prior MIs (CROI 2024, Abstract 633). In a Phase 1 study, VH-937 administered to healthy participants was shown to be generally well-tolerated [212].

6.3. Other Classes

Although not specifically designed for either the direct or indirect inhibition of PR, two other classes of antivirals have been shown to induce defects in the assembly and function of the viral core, resulting in severely compromised particle infectivity.

6.3.1. Capsid Inhibitors

Members of this class of small molecules have a multimodal activity and block both early and late stages of the virus replication cycle (Figure 1D,H,K). The inhibition that CAIs exert in target cells prior to viral DNA integration predominates at clinically relevant inhibitor concentrations [49,213]. Recent work reported that LEN, the first FDA-approved CAI in this class, increases the rigidity or “brittleness” of the viral core, causing hindrance to its nuclear transport and disassembly, eventually preventing the proper integration of the viral DNA in the host genome (Figure 1D) [214,215]. Although with reduced potency, CAIs are also active in virus-producer cells with mechanisms that are less understood. In this regard, it has been reported that CAIs accelerate the rate of CA assembly in vitro and increase the formation of aberrant, noninfectious particles with spherical, non-conical cores (Figure 1K and Figure 3) [49,213]. Several molecules targeting CA have been developed in the past two decades and can be classified based on their binding site [216]. The most potent compounds are those targeting the hydrophobic pocket that lies between CA monomers that are assembled into a hexamer. This site comprises the α-3 and α-4 helices of the CANTD and the α-8 and α-9 helices of the CACTD in the neighboring monomer. Importantly, this pocket mediates the interactions of CA with FG-containing motifs of host factors, including Nup153 and CPSF6 (Figure 3), which perform critical roles in HIV nuclear trafficking and integration (Section 3.2) [47,217,218]. Members of the class of CAIs that bind the FG site include the prototype PF3450074 (PF-74), BI-1, and BI-2, and the more recently developed GSK878, GS-CA1, and LEN (GS-6207) (Table 1). GSK878, GS-CA1, and LEN display remarkable sub-nanomolar half-maximal inhibitory concentrations (IC50) [49,213,219,220,221,222]. Resistance to LEN emerges rapidly in vitro and, in most cases, selects for substitutions that reduce virus fitness, with the exception of the low-level resistance-associated variant Q67H [49].
Because of its potent antiviral activity and favorable drug-like properties, LEN was tested in clinical trials with long-acting formulations and was generally well-tolerated. In an early monotherapy Phase 1b study, single subcutaneous (SC) doses of LEN displayed potent antiviral activity over 9 days. A genetic mixture of WT CA and CA-Q67H mutant emerged in some study participants treated with a low dose (20 mg), which was associated with a decrease in phenotypic susceptibility but without any viral escape noted by day 9 [49]. Subsequently, LEN has also been tested as a twice-yearly SC injectable treatment in combination with other once-daily oral antiretrovirals in a Phase 2 study in treatment-naïve PLWH (CALIBRATE; NCT04143594). This trial confirmed LEN’s safety and potency and showed that emergent resistance, including the Q67H variant, was rare and attributed to an incomplete adherence to the daily oral agents [223]. The Phase 2/3 CAPELLA study (NCT04150068) is being conducted in heavily treatment-experienced (HTE) PLWH with multi-drug resistance. The trial tested an initial 14-day oral administration of LEN alongside the failing therapy, followed by twice-yearly SC injections of LEN with an optimized background regimen (OBR) during the maintenance period starting on day 15. Interim data confirmed LEN’s safety and potency, showing that resistance emerged when LEN was unintentionally used as functional monotherapy due to either insufficient adherence to the OBR or the absence of fully active antivirals in the OBR [224,225,226]. Although LEN, like other CAIs, may have a lower genetic resistance barrier than other classes of antiretrovirals, such as INSTIs, its potency is the best among all antiretrovirals approved to date, and its drug-like properties are favorable for long-acting oral and injectable formulations. For these reasons, in 2022, LEN received approval from the FDA for clinical use in HTE adults who have limited treatment options, and it is being offered as a twice-yearly injection administered in addition to an OBR of once-daily oral antiretrovirals. However, LEN offers the potential for a long-acting regimen if a suitable partner molecule can be developed. Currently, the longest-duration regimen approved for treatment is a once-every-two-month intramuscular (IM) injection of Cabenuva [227,228], although clinical studies are underway to extend the duration to four-month intervals. In addition, LEN is also being investigated for PrEP, and the PURPOSE 1 Phase 3 clinical trial using a twice-yearly SC administration of LEN demonstrated remarkable efficacy and superiority to a daily oral regimen [229]. Currently, the only long-acting agent approved for PrEP is Apretude, an extended-release injectable suspension of CAB that is administered intramuscularly once every two months [230,231].

6.3.2. Allosteric Integrase Inhibitors

The IN is an enzyme with endonuclease and polynucleotide transferase activities present in the catalytic core domain (CCD) that is located between the N-terminal and C-terminal domains of the mature protein. Two sequential catalytic reactions are carried out by retroviral INs: the first processes the 3′ ends of the newly synthesized double-stranded viral DNA, exposing a dinucleotide sequence, and the second is the DNA strand transfer reaction. The latter mediates a covalent ligation of the processed 3′ ends into the host genome [232,233,234,235]. Several INSTIs have become critical components of modern ART due to their high genetic barrier to resistance and several favorable drug-like properties, including long-acting potential [236]. Relevant to this review is a more recently identified secondary activity that IN exerts late in the virus replication cycle by binding to the viral gRNA and promoting its localization within the viral core [237,238]. Early mutagenesis studies identified residues within the CCD that are critical for the enzymatic activity of IN and that are known as class I mutations. Additionally, other sequences outside of IN’s active site were found, whose mutations pleiotropically inhibited the late stages of virus assembly. These were classified as class II mutations [239]. In virus-producer cells, these substitutions induced the formation of virions in which the viral gRNA failed to be packaged inside the core but rather was present as an electron-dense nucleoid located outside of the fullerene cone [239,240]. In the early 2000s, a host factor, lens epithelium-derived growth factor/p75 (LEDGF/p75), was discovered to bind IN, favoring integration in transcriptionally active regions of the host chromatin [241,242,243,244]. Screening for inhibitors of the LEDGF-IN interaction led to the discovery of compounds that were originally called LEDGF-IN inhibitors (LEDGINs) [245,246,247,248,249]. Upon further study, it became clear that the primary mechanism of action of these compounds was not to inhibit integration but rather to disrupt the IN-mediated packaging of the viral gRNA in the core during maturation. It was shown that ALLINIs induce the hyper-multimerization of IN, resulting in virions whose aberrant morphology resembled that of class II IN mutations mentioned above [250,251,252]. These compounds are now predominantly referred to as ALLINIs. X-ray crystallographic analysis demonstrated that ALLINIs mediate the interface between head-to-tail interactions of CCDs in one IN dimer and the C-terminal domain of a second dimer, resulting in the defective multimerization of the protein [250,253,254,255]. The pharmaceutical development of ALLINIs recently led to pirmitegravir (Table 1), also referred to as STP0404, the first molecule in this class advanced to the clinic, an investigational compound currently in Phase 2 clinical development [256].

7. Conclusions

Several different classes of inhibitors have been discovered to target virion maturation and prevent HIV-1 replication. These include PIs, RTIs, INSTIs, MIs, CAIs, and ALLINIs (Figure 1 and Table 1) [257]. The current standard of HIV treatment follows a daily therapy regimen, which could negatively impact those who may not be able to adhere to the daily regimen or neglect to take medications [258,259]. In 2021, the FDA approved the first long-lasting ART regimen, Cabenuva, which consists of CAB and RPV [7]. This regimen requires IM injections of co-packaged CAB and RPV, administered either once a month or every two months. In the further ACTG A5359 LATITUDE Phase 3 trial, a once-monthly IM administration of CAB/RPV demonstrated superior efficacy compared to the daily oral standard of care in PLWH with a history of adherence challenges (NCT03635788; CROI 2024, Abstract 212). Additionally, the IMPAACT 2017 study (MOCHA; NCT03497676) showed that the use of long-acting CAB/RPV was generally well-tolerated in adolescents with HIV (CROI 2024, Abstract 188). In 2022, the FDA approved the first-in-class CAI, LEN, which involves a twice-yearly dosing in combination with a daily OBR for the treatment of HTE PLWH. LEN has a multi-stage mechanism of action, including targeting the late stage of virus replication and interfering with virion assembly and maturation (Figure 1H,K). LEN is specifically effective for the treatment of multidrug-resistant HIV-1 strains. If combined with a suitable long-acting partner, which is yet to be found, LEN has the potential to be included in long-acting regimens for HIV treatment against multidrug-resistant strains. In addition, LEN is currently under investigation for PrEP [260], which could potentially contribute significantly to the prevention of new infections, ultimately working towards the elimination of HIV from humanity. Overall, targeting virion maturation is important for developing long-acting antivirals, especially against multi-drug-resistant mutants, and for providing new options for the treatment and prevention of HIV infection. Currently, there are no FDA-approved MIs for HIV treatment, but VH-937 has shown potential to become the first approved MI in ongoing clinical trials. Further studies on this new compound are highly anticipated.

Future Prospects

A better understanding of virion maturation is critical for the development of antivirals. For this purpose, the establishment of non-biased experimental tools is required to systematically investigate HIV-1 maturation. Though TEM is the most direct experimental technique used to address this question, the rates of virion maturation have been reported in a varied range and have not been clearly determined [49,55,152,153,154]. As described in Section 4, an unbiased virion visualization system utilizing the FRET principle to quantitatively detect the cleavage of Gag and Gag-Pol polyproteins by PR, particularly at the MA-CA junction, has been established [134]. This system serves as an indicator to distinguish between mature and immature HIV-1 virions. (Figure 4). However, the current system visualizing HIV-1 Gag-iFRET labeling viruses is limited, as it can only detect the initial step of maturation for Gag and Gag-Pol polyprotein cleavage by PR. To overcome this limitation, a second-generation HIV-1 Gag-iFRET system that can detect proper conical core formation is under development. When the GFP fluorescent protein is fused with CA and is not cleaved by PR during maturation, the GFP-CA fusion protein is incorporated in the conical core (HIV-1 Gag-iHGFP) [55]. Additionally, the free GFP-CA interacts with viral gRNA that is packaged in the closed core [55]. Based on these features, it is hypothesized that engineering cp173Venus in the HIV-1 Gag-iFRET construct, in fusion with CA and lacking the PR cleavage site, can label closed conical cores after Gag polyprotein cleavage. This new construct expands on the HIV-1 Gag-iFRET system by its ability to detect both (1) the processing of the Gag and Gag-Pol polyprotein and (2) the proper CA assembly to form a core during HIV-1 virion maturation [32]. Through the establishment of this new system, a non-biased, semi-automatic, and quantitative assessment of the full maturation process can be achieved in combination with EM analysis. An FRET tool based on single virion analysis can be applied to better understand the mechanism of action of CAIs and MIs currently under development and to further investigate next-generation antivirals targeting virion maturation. This approach aids in the development of long-acting anti-HIV drugs that can improve the quality of life for PLWH.

Author Contributions

All authors contributed in writing the manuscript, drawing the figures, and reviewing the draft. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by National Institutes of Health grant number 7R15AI172610-02 and District of Columbia Center for AIDS Research grant number P30AI117970.

Acknowledgments

We thank all the administrators and leadership of the authors’ institutions in academia (American University), industry (ViiV Healthcare), and government (National Cancer Institute) for their continued support of this publication. Research in the Freed laboratory is supported by the Intramural Research Program of the Center for Cancer Research, National Cancer Institute, NIH. This review article is partially supported by a research start-up fund to T.I. Additional funding was provided by the National Institutes of Health, specifically through a grant from the National Institute of Allergy and Infectious Diseases (grant number: 7R15AI172610-02 to TI), and in part by a 2024 award from the District of Columbia Center for AIDS Research, an NIH-funded program (P30AI117970), which is supported by the following NIH Co-Funding and Participating Institutes and Centers: NIAID, NCI, NICHD, NHLBI, NIDA, NIMH, NIA, NIDDK, NIMHD, NIDCR, NINR, FIC, and OAR. The content is solely the responsibility of the authors and does not necessarily represent the official views of the NIH. All authors have read and agreed to the published version of the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Siliciano, R.F.; Greene, W.C. HIV latency. Cold Spring Harb. Perspect. Med. 2011, 1, a007096. [Google Scholar] [CrossRef] [PubMed]
  2. Ta, T.M.; Malik, S.; Anderson, E.M.; Jones, A.D.; Perchik, J.; Freylikh, M.; Sardo, L.; Klase, Z.A.; Izumi, T. Insights into Persistent HIV-1 Infection and Functional Cure: Novel Capabilities and Strategies. Front. Microbiol. 2022, 13, 862270. [Google Scholar] [CrossRef] [PubMed]
  3. Sardo, L.; Parolin, C.; Yoshida, T.; Garzino-Demo, A.; Izumi, T. Editorial: Novel Insights into a Functional HIV Cure. Front. Microbiol. 2021, 12, 797570. [Google Scholar] [CrossRef] [PubMed]
  4. Briz, V.; Poveda, E.; Soriano, V. HIV entry inhibitors: Mechanisms of action and resistance pathways. J. Antimicrob. Chemother. 2006, 57, 619–627. [Google Scholar] [CrossRef] [PubMed]
  5. Sharma, A. In PrEP: Long-acting antivirals for HIV prevention. Cell Host Microbe 2022, 30, 148–150. [Google Scholar] [CrossRef]
  6. Goebel, M.C.; Guajardo, E.; Giordano, T.P.; Patel, S.M. The New Era of Long-Acting Antiretroviral Therapy: When and Why to Make the Switch. Curr. HIV/AIDS Rep. 2023, 20, 271–285. [Google Scholar] [CrossRef]
  7. Sension, M.G.; Brunet, L.; Hsu, R.K.; Fusco, J.S.; Cochran, Q.; Uranaka, C.; Sridhar, G.; Vannappagari, V.; Van Wyk, J.; McCurdy, L.; et al. Cabotegravir + Rilpivirine Long-Acting Injections for HIV Treatment in the US: Real World Data from the OPERA Cohort. Infect. Dis. Ther. 2023, 12, 2807–2817. [Google Scholar] [CrossRef]
  8. Dvory-Sobol, H.; Shaik, N.; Callebaut, C.; Rhee, M.S. Lenacapavir: A first-in-class HIV-1 capsid inhibitor. Curr. Opin. HIV AIDS 2022, 17, 15–21. [Google Scholar] [CrossRef]
  9. Novikova, M.; Zhang, Y.; Freed, E.O.; Peng, K. Multiple Roles of HIV-1 Capsid during the Virus Replication Cycle. Virol. Sin. 2019, 34, 119–134. [Google Scholar] [CrossRef]
  10. Freed, E.O. HIV-1 assembly, release and maturation. Nat. Rev. Microbiol. 2015, 13, 484–496. [Google Scholar] [CrossRef]
  11. Klingler, J.; Anton, H.; Real, E.; Zeiger, M.; Moog, C.; Mely, Y.; Boutant, E. How HIV-1 Gag Manipulates Its Host Cell Proteins: A Focus on Interactors of the Nucleocapsid Domain. Viruses 2020, 12, 888. [Google Scholar] [CrossRef] [PubMed]
  12. Olety, B.; Ono, A. Roles played by acidic lipids in HIV-1 Gag membrane binding. Virus Res. 2014, 193, 108–115. [Google Scholar] [CrossRef]
  13. Rossi, E.; Meuser, M.E.; Cunanan, C.J.; Cocklin, S. Structure, Function, and Interactions of the HIV-1 Capsid Protein. Life 2021, 11, 100. [Google Scholar] [CrossRef]
  14. Yu, K.L.; Lee, S.H.; Lee, E.S.; You, J.C. HIV-1 nucleocapsid protein localizes efficiently to the nucleus and nucleolus. Virology 2016, 492, 204–212. [Google Scholar] [CrossRef]
  15. Levin, J.G.; Mitra, M.; Mascarenhas, A.; Musier-Forsyth, K. Role of HIV-1 nucleocapsid protein in HIV-1 reverse transcription. RNA Biol. 2010, 7, 754–774. [Google Scholar] [CrossRef]
  16. Thomas, J.A.; Gorelick, R.J. Nucleocapsid protein function in early infection processes. Virus Res 2008, 134, 39–63. [Google Scholar] [CrossRef] [PubMed]
  17. Sundquist, W.I.; Krausslich, H.G. HIV-1 assembly, budding, and maturation. Cold Spring Harb. Perspect. Med. 2012, 2, a006924. [Google Scholar] [CrossRef]
  18. Meulendyke, K.A.; Croteau, J.D.; Zink, M.C. HIV life cycle, innate immunity and autophagy in the central nervous system. Curr. Opin. HIV AIDS 2014, 9, 565–571. [Google Scholar] [CrossRef] [PubMed]
  19. Gill, M.S.A.; Hassan, S.S.; Ahemad, N. Evolution of HIV-1 reverse transcriptase and integrase dual inhibitors: Recent advances and developments. Eur. J. Med. Chem. 2019, 179, 423–448. [Google Scholar] [CrossRef] [PubMed]
  20. Li, C.; Burdick, R.C.; Nagashima, K.; Hu, W.S.; Pathak, V.K. HIV-1 cores retain their integrity until minutes before uncoating in the nucleus. Proc. Natl. Acad. Sci. USA 2021, 118, e2019467118. [Google Scholar] [CrossRef]
  21. Burdick, R.C.; Duchon, A.; Hu, W.S.; Pathak, V.K. Imaging HIV-1 Nuclear Import, Uncoating, and Proviral Transcription. Methods Mol. Biol. 2024, 2807, 15–30. [Google Scholar] [PubMed]
  22. Ajasin, D.; Eugenin, E.A. HIV-1 Tat: Role in Bystander Toxicity. Front. Cell. Infect. Microbiol. 2020, 10, 517857. [Google Scholar] [CrossRef] [PubMed]
  23. Reeder, J.E.; Kwak, Y.T.; McNamara, R.P.; Forst, C.V.; D’Orso, I. HIV Tat controls RNA Polymerase II and the epigenetic landscape to transcriptionally reprogram target immune cells. Elife 2015, 4, e08955. [Google Scholar] [CrossRef] [PubMed]
  24. Beemon, K.L. Retroviral RNA Processing. Viruses 2022, 14, 1113. [Google Scholar] [CrossRef] [PubMed]
  25. Zhang, X.; Li, J.; Zhang, M.; Bai, B.; Ma, W.; Lin, Y.; Guo, X.; Wang, X.F.; Wang, X. A Novel, Fully Spliced, Accessory Gene in Equine Lentivirus with Distinct Rev-Responsive Element. J. Virol. 2022, 96, e0098622. [Google Scholar] [CrossRef]
  26. Kuzembayeva, M.; Dilley, K.; Sardo, L.; Hu, W.S. Life of psi: How full-length HIV-1 RNAs become packaged genomes in the viral particles. Virology 2014, 454, 362–370. [Google Scholar] [CrossRef]
  27. Coffin, J.M.; Hughes, S.H.; Varmus, H.E. Retroviruses; Cold Spring Harbor: New York, NY, USA, 1997. [Google Scholar]
  28. Chen, J.; Liu, Y.; Wu, B.; Nikolaitchik, O.A.; Mohan, P.R.; Chen, J.; Pathak, V.K.; Hu, W.S. Visualizing the translation and packaging of HIV-1 full-length RNA. Proc. Natl. Acad. Sci. USA 2020, 117, 6145–6155. [Google Scholar] [CrossRef]
  29. Cullen, B.R. Nuclear mRNA export: Insights from virology. Trends Biochem. Sci. 2003, 28, 419–424. [Google Scholar] [CrossRef]
  30. Affranchino, J.L.; Gonzalez, S.A. Understanding the process of envelope glycoprotein incorporation into virions in simian and feline immunodeficiency viruses. Viruses 2014, 6, 264–283. [Google Scholar] [CrossRef]
  31. McGraw, A.; Hillmer, G.; Choi, J.; Narayan, K.; Mehedincu, S.M.; Marquez, D.; Tibebe, H.; DeCicco-Skinner, K.L.; Izumi, T. Evaluating HIV-1 Infectivity and Virion Maturation across Varied Producer Cells with a Novel FRET-Based Detection and Quantification Assay. Int. J. Mol. Sci. 2024, 25, 6396. [Google Scholar] [CrossRef]
  32. Kleinpeter, A.B.; Freed, E.O. HIV-1 Maturation: Lessons Learned from Inhibitors. Viruses 2020, 12, 940. [Google Scholar] [CrossRef]
  33. Biswas, P.; Jiang, X.; Pacchia, A.L.; Dougherty, J.P.; Peltz, S.W. The human immunodeficiency virus type 1 ribosomal frameshifting site is an invariant sequence determinant and an important target for antiviral therapy. J. Virol. 2004, 78, 2082–2087. [Google Scholar] [CrossRef]
  34. Korniy, N.; Goyal, A.; Hoffmann, M.; Samatova, E.; Peske, F.; Pohlmann, S.; Rodnina, M.V. Modulation of HIV-1 Gag/Gag-Pol frameshifting by tRNA abundance. Nucleic Acids Res. 2019, 47, 5210–5222. [Google Scholar] [CrossRef]
  35. Balvay, L.; Lopez Lastra, M.; Sargueil, B.; Darlix, J.L.; Ohlmann, T. Translational control of retroviruses. Nat. Rev. Microbiol. 2007, 5, 128–140. [Google Scholar] [CrossRef]
  36. Das, A.T.; Harwig, A.; Berkhout, B. The HIV-1 Tat protein has a versatile role in activating viral transcription. J. Virol. 2011, 85, 9506–9516. [Google Scholar] [CrossRef] [PubMed]
  37. Liu, R.D.; Wu, J.; Shao, R.; Xue, Y.H. Mechanism and factors that control HIV-1 transcription and latency activation. J. Zhejiang Univ. Sci. B 2014, 15, 455–465. [Google Scholar] [CrossRef] [PubMed]
  38. Pornillos, O.; Ganser-Pornillos, B.K. Maturation of retroviruses. Curr. Opin. Virol. 2019, 36, 47–55. [Google Scholar] [CrossRef] [PubMed]
  39. Jacks, T.; Power, M.D.; Masiarz, F.R.; Luciw, P.A.; Barr, P.J.; Varmus, H.E. Characterization of ribosomal frameshifting in HIV-1 gag-pol expression. Nature 1988, 331, 280–283. [Google Scholar] [CrossRef] [PubMed]
  40. Lin, Y.R.; Chu, S.M.; Yu, F.H.; Huang, K.J.; Wang, C.T. Effects of reduced gag cleavage efficiency on HIV-1 Gag-Pol package. BMC Microbiol. 2022, 22, 94. [Google Scholar] [CrossRef]
  41. Centazzo, M.; Manganaro, L.; Alvisi, G. Cellular Targets of HIV-1 Protease: Just the Tip of the Iceberg? Viruses 2023, 15, 712. [Google Scholar] [CrossRef]
  42. Qu, K.; Ke, Z.; Zila, V.; Anders-Osswein, M.; Glass, B.; Mucksch, F.; Muller, R.; Schultz, C.; Muller, B.; Krausslich, H.G.; et al. Maturation of the matrix and viral membrane of HIV-1. Science 2021, 373, 700–704. [Google Scholar] [CrossRef] [PubMed]
  43. Zhao, G.; Perilla, J.R.; Yufenyuy, E.L.; Meng, X.; Chen, B.; Ning, J.; Ahn, J.; Gronenborn, A.M.; Schulten, K.; Aiken, C.; et al. Mature HIV-1 capsid structure by cryo-electron microscopy and all-atom molecular dynamics. Nature 2013, 497, 643–646. [Google Scholar] [CrossRef]
  44. Nicastro, G.; Lucci, M.; Oregioni, A.; Kelly, G.; Frenkiel, T.A.; Taylor, I.A. CP-MAS and Solution NMR Studies of Allosteric Communication in CA-assemblies of HIV-1. J. Mol. Biol. 2022, 434, 167691. [Google Scholar] [CrossRef] [PubMed]
  45. Schirra, R.T.; Dos Santos, N.F.B.; Zadrozny, K.K.; Kucharska, I.; Ganser-Pornillos, B.K.; Pornillos, O. A molecular switch modulates assembly and host factor binding of the HIV-1 capsid. Nat. Struct. Mol. Biol. 2023, 30, 383–390. [Google Scholar] [CrossRef] [PubMed]
  46. Mallery, D.L.; Faysal, K.M.R.; Kleinpeter, A.; Wilson, M.S.C.; Vaysburd, M.; Fletcher, A.J.; Novikova, M.; Bocking, T.; Freed, E.O.; Saiardi, A.; et al. Cellular IP(6) Levels Limit HIV Production while Viruses that Cannot Efficiently Package IP(6) Are Attenuated for Infection and Replication. Cell. Rep. 2019, 29, 3983–3996.e4. [Google Scholar] [CrossRef]
  47. Price, A.J.; Jacques, D.A.; McEwan, W.A.; Fletcher, A.J.; Essig, S.; Chin, J.W.; Halambage, U.D.; Aiken, C.; James, L.C. Host cofactors and pharmacologic ligands share an essential interface in HIV-1 capsid that is lost upon disassembly. PLoS Pathog. 2014, 10, e1004459. [Google Scholar] [CrossRef]
  48. Li, S.; Patel, J.S.; Yang, J.; Crabtree, A.M.; Rubenstein, B.M.; Lund-Andersen, P.K.; Ytreberg, F.M.; Rowley, P.A. Defining the HIV Capsid Binding Site of Nucleoporin 153. mSphere 2022, 7, e0031022. [Google Scholar] [CrossRef]
  49. Link, J.O.; Rhee, M.S.; Tse, W.C.; Zheng, J.; Somoza, J.R.; Rowe, W.; Begley, R.; Chiu, A.; Mulato, A.; Hansen, D.; et al. Clinical targeting of HIV capsid protein with a long-acting small molecule. Nature 2020, 584, 614–618. [Google Scholar] [CrossRef]
  50. Noser, J.A.; Towers, G.J.; Sakuma, R.; Dumont, J.M.; Collins, M.K.; Ikeda, Y. Cyclosporine increases human immunodeficiency virus type 1 vector transduction of primary mouse cells. J. Virol. 2006, 80, 7769–7774. [Google Scholar] [CrossRef]
  51. Yoo, S.; Myszka, D.G.; Yeh, C.; McMurray, M.; Hill, C.P.; Sundquist, W.I. Molecular recognition in the HIV-1 capsid/cyclophilin A complex. J. Mol. Biol. 1997, 269, 780–795. [Google Scholar] [CrossRef]
  52. Gamble, T.R.; Vajdos, F.F.; Yoo, S.; Worthylake, D.K.; Houseweart, M.; Sundquist, W.I.; Hill, C.P. Crystal structure of human cyclophilin A bound to the amino-terminal domain of HIV-1 capsid. Cell 1996, 87, 1285–1294. [Google Scholar] [CrossRef]
  53. Luban, J. Absconding with the chaperone: Essential cyclophilin-Gag interaction in HIV-1 virions. Cell 1996, 87, 1157–1159. [Google Scholar] [CrossRef]
  54. Braaten, D.; Franke, E.K.; Luban, J. Cyclophilin A is required for an early step in the life cycle of human immunodeficiency virus type 1 before the initiation of reverse transcription. J. Virol. 1996, 70, 3551–3560. [Google Scholar] [CrossRef] [PubMed]
  55. Burdick, R.C.; Li, C.; Munshi, M.; Rawson, J.M.O.; Nagashima, K.; Hu, W.S.; Pathak, V.K. HIV-1 uncoats in the nucleus near sites of integration. Proc. Natl. Acad. Sci. USA 2020, 117, 5486–5493. [Google Scholar] [CrossRef] [PubMed]
  56. Dwivedi, R.; Prakash, P.; Kumbhar, B.V.; Balasubramaniam, M.; Dash, C. HIV-1 capsid and viral DNA integration. mBio 2024, 15, e0021222. [Google Scholar] [CrossRef] [PubMed]
  57. Sokolskaja, E.; Sayah, D.M.; Luban, J. Target cell cyclophilin A modulates human immunodeficiency virus type 1 infectivity. J. Virol. 2004, 78, 12800–12808. [Google Scholar] [CrossRef]
  58. Kim, K.; Dauphin, A.; Komurlu, S.; McCauley, S.M.; Yurkovetskiy, L.; Carbone, C.; Diehl, W.E.; Strambio-De-Castillia, C.; Campbell, E.M.; Luban, J. Cyclophilin A protects HIV-1 from restriction by human TRIM5alpha. Nat. Microbiol. 2019, 4, 2044–2051. [Google Scholar] [CrossRef]
  59. Li, Y.; Kar, A.K.; Sodroski, J. Target cell type-dependent modulation of human immunodeficiency virus type 1 capsid disassembly by cyclophilin A. J. Virol. 2009, 83, 10951–10962. [Google Scholar] [CrossRef]
  60. Shah, V.B.; Shi, J.; Hout, D.R.; Oztop, I.; Krishnan, L.; Ahn, J.; Shotwell, M.S.; Engelman, A.; Aiken, C. The host proteins transportin SR2/TNPO3 and cyclophilin A exert opposing effects on HIV-1 uncoating. J. Virol. 2013, 87, 422–432. [Google Scholar] [CrossRef]
  61. Rasaiyaah, J.; Tan, C.P.; Fletcher, A.J.; Price, A.J.; Blondeau, C.; Hilditch, L.; Jacques, D.A.; Selwood, D.L.; James, L.C.; Noursadeghi, M.; et al. HIV-1 evades innate immune recognition through specific cofactor recruitment. Nature 2013, 503, 402–405. [Google Scholar] [CrossRef]
  62. Lu, M.; Hou, G.; Zhang, H.; Suiter, C.L.; Ahn, J.; Byeon, I.J.; Perilla, J.R.; Langmead, C.J.; Hung, I.; Gor’kov, P.L.; et al. Dynamic allostery governs cyclophilin A-HIV capsid interplay. Proc. Natl. Acad. Sci. USA 2015, 112, 14617–14622. [Google Scholar] [CrossRef] [PubMed]
  63. Luban, J. TRIM5 and the Regulation of HIV-1 Infectivity. Mol. Biol. Int. 2012, 2012, 426840. [Google Scholar] [CrossRef]
  64. Grutter, M.G.; Luban, J. TRIM5 structure, HIV-1 capsid recognition, and innate immune signaling. Curr. Opin. Virol. 2012, 2, 142–150. [Google Scholar] [CrossRef]
  65. Kono, K.; Song, H.; Shingai, Y.; Shioda, T.; Nakayama, E.E. Comparison of anti-viral activity of rhesus monkey and cynomolgus monkey TRIM5alphas against human immunodeficiency virus type 2 infection. Virology 2008, 373, 447–456. [Google Scholar] [CrossRef] [PubMed]
  66. Zhao, G.; Ke, D.; Vu, T.; Ahn, J.; Shah, V.B.; Yang, R.; Aiken, C.; Charlton, L.M.; Gronenborn, A.M.; Zhang, P. Rhesus TRIM5alpha disrupts the HIV-1 capsid at the inter-hexamer interfaces. PLoS Pathog. 2011, 7, e1002009. [Google Scholar] [CrossRef]
  67. Sakuma, R.; Noser, J.A.; Ohmine, S.; Ikeda, Y. Rhesus monkey TRIM5alpha restricts HIV-1 production through rapid degradation of viral Gag polyproteins. Nat. Med. 2007, 13, 631–635. [Google Scholar] [CrossRef]
  68. Soll, S.J.; Wilson, S.J.; Kutluay, S.B.; Hatziioannou, T.; Bieniasz, P.D. Assisted evolution enables HIV-1 to overcome a high TRIM5alpha-imposed genetic barrier to rhesus macaque tropism. PLoS Pathog. 2013, 9, e1003667. [Google Scholar] [CrossRef] [PubMed]
  69. Santelli, E.; Bankston, L.A.; Leppla, S.H.; Liddington, R.C. Crystal structure of a complex between anthrax toxin and its host cell receptor. Nature 2004, 430, 905–908. [Google Scholar] [CrossRef]
  70. Sokolskaja, E.; Berthoux, L.; Luban, J. Cyclophilin A and TRIM5alpha independently regulate human immunodeficiency virus type 1 infectivity in human cells. J. Virol. 2006, 80, 2855–2862. [Google Scholar] [CrossRef]
  71. Strebel, K.; Luban, J.; Jeang, K.T. Human cellular restriction factors that target HIV-1 replication. BMC Med. 2009, 7, 48. [Google Scholar] [CrossRef]
  72. Sayah, D.M.; Sokolskaja, E.; Berthoux, L.; Luban, J. Cyclophilin A retrotransposition into TRIM5 explains owl monkey resistance to HIV-1. Nature 2004, 430, 569–573. [Google Scholar] [CrossRef] [PubMed]
  73. Jimenez-Moyano, E.; Ruiz, A.; Kloverpris, H.N.; Rodriguez-Plata, M.T.; Pena, R.; Blondeau, C.; Selwood, D.L.; Izquierdo-Useros, N.; Moris, A.; Clotet, B.; et al. Nonhuman TRIM5 Variants Enhance Recognition of HIV-1-Infected Cells by CD8+ T Cells. J. Virol. 2016, 90, 8552–8562. [Google Scholar] [CrossRef] [PubMed]
  74. Cloherty, A.P.M.; Rader, A.G.; Compeer, B.; Ribeiro, C.M.S. Human TRIM5alpha: Autophagy Connects Cell-Intrinsic HIV-1 Restriction and Innate Immune Sensor Functioning. Viruses 2021, 13, 320. [Google Scholar] [CrossRef]
  75. Ganser-Pornillos, B.K.; Pornillos, O. Restriction of HIV-1 and other retroviruses by TRIM5. Nat. Rev. Microbiol. 2019, 17, 546–556. [Google Scholar] [CrossRef]
  76. Achuthan, V.; Perreira, J.M.; Sowd, G.A.; Puray-Chavez, M.; McDougall, W.M.; Paulucci-Holthauzen, A.; Wu, X.; Fadel, H.J.; Poeschla, E.M.; Multani, A.S.; et al. Capsid-CPSF6 Interaction Licenses Nuclear HIV-1 Trafficking to Sites of Viral DNA Integration. Cell Host Microbe 2018, 24, 392–404.e8. [Google Scholar] [CrossRef]
  77. McDonald, D.; Vodicka, M.A.; Lucero, G.; Svitkina, T.M.; Borisy, G.G.; Emerman, M.; Hope, T.J. Visualization of the intracellular behavior of HIV in living cells. J. Cell Biol. 2002, 159, 441–452. [Google Scholar] [CrossRef] [PubMed]
  78. Zhong, Z.; Ning, J.; Boggs, E.A.; Jang, S.; Wallace, C.; Telmer, C.; Bruchez, M.P.; Ahn, J.; Engelman, A.N.; Zhang, P.; et al. Cytoplasmic CPSF6 Regulates HIV-1 Capsid Trafficking and Infection in a Cyclophilin A-Dependent Manner. mBio 2021, 12, e03142-20. [Google Scholar] [CrossRef]
  79. Lukic, Z.; Dharan, A.; Fricke, T.; Diaz-Griffero, F.; Campbell, E.M. HIV-1 uncoating is facilitated by dynein and kinesin 1. J. Virol. 2014, 88, 13613–13625. [Google Scholar] [CrossRef]
  80. Malikov, V.; Naghavi, M.H. Localized Phosphorylation of a Kinesin-1 Adaptor by a Capsid-Associated Kinase Regulates HIV-1 Motility and Uncoating. Cell Rep. 2017, 20, 2792–2799. [Google Scholar] [CrossRef]
  81. Fernandez, J.; Portilho, D.M.; Danckaert, A.; Munier, S.; Becker, A.; Roux, P.; Zambo, A.; Shorte, S.; Jacob, Y.; Vidalain, P.O.; et al. Microtubule-associated proteins 1 (MAP1) promote human immunodeficiency virus type I (HIV-1) intracytoplasmic routing to the nucleus. J. Biol. Chem. 2015, 290, 4631–4646. [Google Scholar] [CrossRef]
  82. Carnes, S.K.; Zhou, J.; Aiken, C. HIV-1 Engages a Dynein-Dynactin-BICD2 Complex for Infection and Transport to the Nucleus. J. Virol. 2018, 92, e00358-18. [Google Scholar] [CrossRef]
  83. Dharan, A.; Opp, S.; Abdel-Rahim, O.; Keceli, S.K.; Imam, S.; Diaz-Griffero, F.; Campbell, E.M. Bicaudal D2 facilitates the cytoplasmic trafficking and nuclear import of HIV-1 genomes during infection. Proc. Natl. Acad. Sci. USA 2017, 114, E10707–E10716. [Google Scholar] [CrossRef] [PubMed]
  84. Huang, P.T.; Summers, B.J.; Xu, C.; Perilla, J.R.; Malikov, V.; Naghavi, M.H.; Xiong, Y. FEZ1 Is Recruited to a Conserved Cofactor Site on Capsid to Promote HIV-1 Trafficking. Cell Rep. 2019, 28, 2373–2385.e7. [Google Scholar] [CrossRef]
  85. Malikov, V.; da Silva, E.S.; Jovasevic, V.; Bennett, G.; de Souza Aranha Vieira, D.A.; Schulte, B.; Diaz-Griffero, F.; Walsh, D.; Naghavi, M.H. HIV-1 capsids bind and exploit the kinesin-1 adaptor FEZ1 for inward movement to the nucleus. Nat. Commun. 2015, 6, 6660. [Google Scholar] [CrossRef]
  86. Dharan, A.; Talley, S.; Tripathi, A.; Mamede, J.I.; Majetschak, M.; Hope, T.J.; Campbell, E.M. KIF5B and Nup358 Cooperatively Mediate the Nuclear Import of HIV-1 during Infection. PLoS Pathog. 2016, 12, e1005700. [Google Scholar] [CrossRef]
  87. Schaller, T.; Ocwieja, K.E.; Rasaiyaah, J.; Price, A.J.; Brady, T.L.; Roth, S.L.; Hue, S.; Fletcher, A.J.; Lee, K.; KewalRamani, V.N.; et al. HIV-1 capsid-cyclophilin interactions determine nuclear import pathway, integration targeting and replication efficiency. PLoS Pathog. 2011, 7, e1002439. [Google Scholar] [CrossRef]
  88. Bichel, K.; Price, A.J.; Schaller, T.; Towers, G.J.; Freund, S.M.; James, L.C. HIV-1 capsid undergoes coupled binding and isomerization by the nuclear pore protein NUP358. Retrovirology 2013, 10, 81. [Google Scholar] [CrossRef] [PubMed]
  89. Wei, G.; Iqbal, N.; Courouble, V.V.; Francis, A.C.; Singh, P.K.; Hudait, A.; Annamalai, A.S.; Bester, S.; Huang, S.W.; Shkriabai, N.; et al. Prion-like low complexity regions enable avid virus-host interactions during HIV-1 infection. Nat. Commun. 2022, 13, 5879. [Google Scholar] [CrossRef]
  90. Shen, Q.; Kumari, S.; Xu, C.; Jang, S.; Shi, J.; Burdick, R.C.; Levintov, L.; Xiong, Q.; Wu, C.; Devarkar, S.C.; et al. The capsid lattice engages a bipartite NUP153 motif to mediate nuclear entry of HIV-1 cores. Proc. Natl. Acad. Sci. USA 2023, 120, e2202815120. [Google Scholar] [CrossRef] [PubMed]
  91. Matreyek, K.A.; Yucel, S.S.; Li, X.; Engelman, A. Nucleoporin NUP153 phenylalanine-glycine motifs engage a common binding pocket within the HIV-1 capsid protein to mediate lentiviral infectivity. PLoS Pathog. 2013, 9, e1003693. [Google Scholar] [CrossRef]
  92. Bejarano, D.A.; Peng, K.; Laketa, V.; Borner, K.; Jost, K.L.; Lucic, B.; Glass, B.; Lusic, M.; Muller, B.; Krausslich, H.G. HIV-1 nuclear import in macrophages is regulated by CPSF6-capsid interactions at the nuclear pore complex. Elife 2019, 8, e41800. [Google Scholar] [CrossRef] [PubMed]
  93. King, M.C.; Raposo, G.; Lemmon, M.A. Inhibition of nuclear import and cell-cycle progression by mutated forms of the dynamin-like GTPase MxB. Proc. Natl. Acad. Sci. USA 2004, 101, 8957–8962. [Google Scholar] [CrossRef] [PubMed]
  94. Ding, J.; Zhao, J.; Yang, Z.; Ma, L.; Mi, Z.; Wu, Y.; Guo, J.; Zhou, J.; Li, X.; Guo, Y.; et al. Microbial Natural Product Alternariol 5-O-Methyl Ether Inhibits HIV-1 Integration by Blocking Nuclear Import of the Pre-Integration Complex. Viruses 2017, 9, 105. [Google Scholar] [CrossRef]
  95. Raghavendra, N.K.; Shkriabai, N.; Graham, R.; Hess, S.; Kvaratskhelia, M.; Wu, L. Identification of host proteins associated with HIV-1 preintegration complexes isolated from infected CD4+ cells. Retrovirology 2010, 7, 66. [Google Scholar] [CrossRef] [PubMed]
  96. Fricke, T.; White, T.E.; Schulte, B.; de Souza Aranha Vieira, D.A.; Dharan, A.; Campbell, E.M.; Brandariz-Nunez, A.; Diaz-Griffero, F. MxB binds to the HIV-1 core and prevents the uncoating process of HIV-1. Retrovirology 2014, 11, 68. [Google Scholar] [CrossRef]
  97. Kane, M.; Yadav, S.S.; Bitzegeio, J.; Kutluay, S.B.; Zang, T.; Wilson, S.J.; Schoggins, J.W.; Rice, C.M.; Yamashita, M.; Hatziioannou, T.; et al. MX2 is an interferon-induced inhibitor of HIV-1 infection. Nature 2013, 502, 563–566. [Google Scholar] [CrossRef]
  98. Goujon, C.; Moncorge, O.; Bauby, H.; Doyle, T.; Ward, C.C.; Schaller, T.; Hue, S.; Barclay, W.S.; Schulz, R.; Malim, M.H. Human MX2 is an interferon-induced post-entry inhibitor of HIV-1 infection. Nature 2013, 502, 559–562. [Google Scholar] [CrossRef]
  99. Liu, Z.; Pan, Q.; Ding, S.; Qian, J.; Xu, F.; Zhou, J.; Cen, S.; Guo, F.; Liang, C. The interferon-inducible MxB protein inhibits HIV-1 infection. Cell Host Microbe 2013, 14, 398–410. [Google Scholar] [CrossRef]
  100. Haller, O.; Staeheli, P.; Schwemmle, M.; Kochs, G. Mx GTPases: Dynamin-like antiviral machines of innate immunity. Trends Microbiol. 2015, 23, 154–163. [Google Scholar] [CrossRef]
  101. Xie, L.; Chen, L.; Zhong, C.; Yu, T.; Ju, Z.; Wang, M.; Xiong, H.; Zeng, Y.; Wang, J.; Hu, H.; et al. MxB impedes the NUP358-mediated HIV-1 pre-integration complex nuclear import and viral replication cooperatively with CPSF6. Retrovirology 2020, 17, 16. [Google Scholar] [CrossRef]
  102. Lu, L.; Yang, Y.; Yang, Z.; Wu, Y.; Liu, X.; Li, X.; Chen, L.; Han, Y.; Song, X.; Kong, Z.; et al. Altered plasma metabolites and inflammatory networks in HIV-1 infected patients with different immunological responses after long-term antiretroviral therapy. Front. Immunol. 2023, 14, 1254155. [Google Scholar] [CrossRef]
  103. Mikaeloff, F.; Svensson Akusjarvi, S.; Ikomey, G.M.; Krishnan, S.; Sperk, M.; Gupta, S.; Magdaleno, G.D.V.; Escos, A.; Lyonga, E.; Okomo, M.C.; et al. Trans cohort metabolic reprogramming towards glutaminolysis in long-term successfully treated HIV-infection. Commun. Biol. 2022, 5, 27. [Google Scholar] [CrossRef] [PubMed]
  104. Aiken, C.; Rousso, I. The HIV-1 capsid and reverse transcription. Retrovirology 2021, 18, 29. [Google Scholar] [CrossRef] [PubMed]
  105. Renner, N.; Kleinpeter, A.; Mallery, D.L.; Albecka, A.; Rifat Faysal, K.M.; Bocking, T.; Saiardi, A.; Freed, E.O.; James, L.C. HIV-1 is dependent on its immature lattice to recruit IP6 for mature capsid assembly. Nat. Struct. Mol. Biol. 2023, 30, 370–382. [Google Scholar] [CrossRef] [PubMed]
  106. Pak, A.J.; Purdy, M.D.; Yeager, M.; Voth, G.A. Preservation of HIV-1 Gag Helical Bundle Symmetry by Bevirimat Is Central to Maturation Inhibition. J. Am. Chem. Soc. 2021, 143, 19137–19148. [Google Scholar] [CrossRef]
  107. Dick, R.A.; Zadrozny, K.K.; Xu, C.; Schur, F.K.M.; Lyddon, T.D.; Ricana, C.L.; Wagner, J.M.; Perilla, J.R.; Ganser-Pornillos, B.K.; Johnson, M.C.; et al. Inositol phosphates are assembly co-factors for HIV-1. Nature 2018, 560, 509–512. [Google Scholar] [CrossRef]
  108. Mallery, D.L.; Marquez, C.L.; McEwan, W.A.; Dickson, C.F.; Jacques, D.A.; Anandapadamanaban, M.; Bichel, K.; Towers, G.J.; Saiardi, A.; Bocking, T.; et al. IP6 is an HIV pocket factor that prevents capsid collapse and promotes DNA synthesis. Elife 2018, 7, e35335. [Google Scholar] [CrossRef]
  109. Marie, V.; Gordon, M.L. The HIV-1 Gag Protein Displays Extensive Functional and Structural Roles in Virus Replication and Infectivity. Int. J. Mol. Sci. 2022, 23, 7569. [Google Scholar] [CrossRef]
  110. Evilevitch, A.; Sae-Ueng, U. Mechanical Capsid Maturation Facilitates the Resolution of Conflicting Requirements for Herpesvirus Assembly. J. Virol. 2022, 96, e0183121. [Google Scholar] [CrossRef]
  111. Dick, R.A.; Xu, C.; Morado, D.R.; Kravchuk, V.; Ricana, C.L.; Lyddon, T.D.; Broad, A.M.; Feathers, J.R.; Johnson, M.C.; Vogt, V.M.; et al. Structures of immature EIAV Gag lattices reveal a conserved role for IP6 in lentivirus assembly. PLoS Pathog. 2020, 16, e1008277. [Google Scholar] [CrossRef]
  112. Sowd, G.A.; Aiken, C. Inositol phosphates promote HIV-1 assembly and maturation to facilitate viral spread in human CD4+ T cells. PLoS Pathog. 2021, 17, e1009190, Erratum in PLoS Pathog. 2021, 17, e1009389. [Google Scholar]
  113. Mallery, D.L.; Kleinpeter, A.B.; Renner, N.; Faysal, K.M.R.; Novikova, M.; Kiss, L.; Wilson, M.S.C.; Ahsan, B.; Ke, Z.; Briggs, J.A.G.; et al. A stable immature lattice packages IP(6) for HIV capsid maturation. Sci. Adv. 2021, 7, eabe4716. [Google Scholar] [CrossRef] [PubMed]
  114. Ricana, C.L.; Lyddon, T.D.; Dick, R.A.; Johnson, M.C. Primate lentiviruses require Inositol hexakisphosphate (IP6) or inositol pentakisphosphate (IP5) for the production of viral particles. PLoS Pathog. 2020, 16, e1008646. [Google Scholar] [CrossRef]
  115. Obr, M.; Schur, F.K.M.; Dick, R.A. A Structural Perspective of the Role of IP6 in Immature and Mature Retroviral Assembly. Viruses 2021, 13, 1853. [Google Scholar] [CrossRef] [PubMed]
  116. Jennings, J.; Bracey, H.; Nguyen, D.T.; Dasgupta, R.; Rivera, A.V.; Sluis-Cremer, N.; Shi, J.; Aiken, C. The HIV-1 capsid serves as a nanoscale reaction vessel for reverse transcription. PLoS Pathog. 2024, 20, e1011810. [Google Scholar] [CrossRef]
  117. Roncato, R.; Angelini, J.; Pani, A.; Talotta, R. Lipid rafts as viral entry routes and immune platforms: A double-edged sword in SARS-CoV-2 infection? Biochim. Biophys. Acta Mol. Cell Biol. Lipids 2022, 1867, 159140. [Google Scholar] [CrossRef] [PubMed]
  118. Yang, N.J.; Hinner, M.J. Getting across the cell membrane: An overview for small molecules, peptides, and proteins. Methods Mol. Biol. 2015, 1266, 29–53. [Google Scholar]
  119. Ono, A.; Ablan, S.D.; Lockett, S.J.; Nagashima, K.; Freed, E.O. Phosphatidylinositol (4,5) bisphosphate regulates HIV-1 Gag targeting to the plasma membrane. Proc. Natl. Acad. Sci. USA 2004, 101, 14889–14894. [Google Scholar] [CrossRef]
  120. Shamseddine, A.A.; Airola, M.V.; Hannun, Y.A. Roles and regulation of neutral sphingomyelinase-2 in cellular and pathological processes. Adv. Biol. Regul. 2015, 57, 24–41. [Google Scholar] [CrossRef]
  121. Waheed, A.A.; Zhu, Y.; Agostino, E.; Naing, L.; Hikichi, Y.; Soheilian, F.; Yoo, S.W.; Song, Y.; Zhang, P.; Slusher, B.S.; et al. Neutral sphingomyelinase 2 is required for HIV-1 maturation. Proc. Natl. Acad. Sci. USA 2023, 120, e2219475120. [Google Scholar] [CrossRef]
  122. Yoo, S.W.; Waheed, A.A.; Deme, P.; Tohumeken, S.; Rais, R.; Smith, M.D.; DeMarino, C.; Calabresi, P.A.; Kashanchi, F.; Freed, E.O.; et al. Inhibition of neutral sphingomyelinase 2 impairs HIV-1 envelope formation and substantially delays or eliminates viral rebound. Proc. Natl. Acad. Sci. USA 2023, 120, e2219543120. [Google Scholar] [CrossRef] [PubMed]
  123. Konvalinka, J.; Heuser, A.M.; Hruskova-Heidingsfeldova, O.; Vogt, V.M.; Sedlacek, J.; Strop, P.; Krausslich, H.G. Proteolytic processing of particle-associated retroviral polyproteins by homologous and heterologous viral proteinases. Eur. J. Biochem. 1995, 228, 191–198. [Google Scholar] [CrossRef] [PubMed]
  124. Darke, P.L.; Leu, C.T.; Davis, L.J.; Heimbach, J.C.; Diehl, R.E.; Hill, W.S.; Dixon, R.A.; Sigal, I.S. Human immunodeficiency virus protease. Bacterial expression and characterization of the purified aspartic protease. J. Biol. Chem. 1989, 264, 2307–2312. [Google Scholar] [CrossRef] [PubMed]
  125. Hyland, L.J.; Tomaszek, T.A., Jr.; Meek, T.D. Human immunodeficiency virus-1 protease. 2. Use of pH rate studies and solvent kinetic isotope effects to elucidate details of chemical mechanism. Biochemistry 1991, 30, 8454–8463. [Google Scholar] [CrossRef]
  126. Perilla, J.R.; Schulten, K. Physical properties of the HIV-1 capsid from all-atom molecular dynamics simulations. Nat. Commun. 2017, 8, 15959. [Google Scholar] [CrossRef]
  127. Cosset, F.L.; Lavillette, D. Cell entry of enveloped viruses. Adv. Genet. 2011, 73, 121–183. [Google Scholar]
  128. Roche, S.; Gaudin, Y. Characterization of the equilibrium between the native and fusion-inactive conformation of rabies virus glycoprotein indicates that the fusion complex is made of several trimers. Virology 2002, 297, 128–135. [Google Scholar] [CrossRef]
  129. Romero-Brey, I.; Bartenschlager, R. Viral Infection at High Magnification: 3D Electron Microscopy Methods to Analyze the Architecture of Infected Cells. Viruses 2015, 7, 6316–6345. [Google Scholar] [CrossRef]
  130. Roingeard, P.; Raynal, P.I.; Eymieux, S.; Blanchard, E. Virus detection by transmission electron microscopy: Still useful for diagnosis and a plus for biosafety. Rev. Med. Virol. 2019, 29, e2019. [Google Scholar] [CrossRef]
  131. Richert-Poggeler, K.R.; Franzke, K.; Hipp, K.; Kleespies, R.G. Electron Microscopy Methods for Virus Diagnosis and High Resolution Analysis of Viruses. Front. Microbiol. 2018, 9, 3255. [Google Scholar] [CrossRef]
  132. Graham, L.; Orenstein, J.M. Processing tissue and cells for transmission electron microscopy in diagnostic pathology and research. Nat. Protoc. 2007, 2, 2439–2450. [Google Scholar] [CrossRef] [PubMed]
  133. Rey, J.S.; Li, W.; Bryer, A.J.; Beatson, H.; Lantz, C.; Engelman, A.N.; Perilla, J.R. Deep-learning in situ classification of HIV-1 virion morphology. Comput. Struct. Biotechnol. J. 2021, 19, 5688–5700. [Google Scholar] [CrossRef]
  134. Sarca, A.D.; Sardo, L.; Fukuda, H.; Matsui, H.; Shirakawa, K.; Horikawa, K.; Takaori-Kondo, A.; Izumi, T. FRET-Based Detection and Quantification of HIV-1 Virion Maturation. Front. Microbiol. 2021, 12, 647452. [Google Scholar] [CrossRef] [PubMed]
  135. Benjin, X.; Ling, L. Developments, applications, and prospects of cryo-electron microscopy. Protein Sci. 2020, 29, 872–882. [Google Scholar] [CrossRef]
  136. Mendonca, L.; Sun, D.; Ning, J.; Liu, J.; Kotecha, A.; Olek, M.; Frosio, T.; Fu, X.; Himes, B.A.; Kleinpeter, A.B.; et al. CryoET structures of immature HIV Gag reveal six-helix bundle. Commun. Biol. 2021, 4, 481. [Google Scholar] [CrossRef]
  137. Schoehn, G.; Chenavier, F.; Crepin, T. Advances in Structural Virology via Cryo-EM in 2022. Viruses 2023, 15, 1315. [Google Scholar] [CrossRef] [PubMed]
  138. Ward, A.E.; Kiessling, V.; Pornillos, O.; White, J.M.; Ganser-Pornillos, B.K.; Tamm, L.K. HIV-cell membrane fusion intermediates are restricted by Serincs as revealed by cryo-electron and TIRF microscopy. J. Biol. Chem. 2020, 295, 15183–15195. [Google Scholar] [CrossRef]
  139. Zila, V.; Margiotta, E.; Turonova, B.; Muller, T.G.; Zimmerli, C.E.; Mattei, S.; Allegretti, M.; Borner, K.; Rada, J.; Muller, B.; et al. Cone-shaped HIV-1 capsids are transported through intact nuclear pores. Cell 2021, 184, 1032–1046.e18. [Google Scholar] [CrossRef]
  140. Mangala Prasad, V.; Leaman, D.P.; Lovendahl, K.N.; Croft, J.T.; Benhaim, M.A.; Hodge, E.A.; Zwick, M.B.; Lee, K.K. Cryo-ET of Env on intact HIV virions reveals structural variation and positioning on the Gag lattice. Cell 2022, 185, 641–653.e17. [Google Scholar] [CrossRef]
  141. Golm, S.K.; Hubner, W.; Muller, K.M. Fluorescence Microscopy in Adeno-Associated Virus Research. Viruses 2023, 15, 1174. [Google Scholar] [CrossRef]
  142. Huang, B.; Bates, M.; Zhuang, X. Super-resolution fluorescence microscopy. Annu. Rev. Biochem. 2009, 78, 993–1016. [Google Scholar] [CrossRef]
  143. Chen, G.W.; Song, F.L.; Xiong, X.Q.; Peng, X.J. Fluorescent Nanosensors Based on Fluorescence Resonance Energy Transfer (FRET). Ind. Eng. Chem. Res. 2013, 52, 11228–11245. [Google Scholar] [CrossRef]
  144. Sekar, R.B.; Periasamy, A. Fluorescence resonance energy transfer (FRET) microscopy imaging of live cell protein localizations. J. Cell Biol. 2003, 160, 629–633. [Google Scholar] [CrossRef] [PubMed]
  145. Jin, S.; Ellis, E.; Veetil, J.V.; Yao, H.; Ye, K. Visualization of human immunodeficiency virus protease inhibition using a novel Forster resonance energy transfer molecular probe. Biotechnol. Prog. 2011, 27, 1107–1114. [Google Scholar] [CrossRef] [PubMed]
  146. Sood, C.; Francis, A.C.; Desai, T.M.; Melikyan, G.B. An improved labeling strategy enables automated detection of single-virus fusion and assessment of HIV-1 protease activity in single virions. J. Biol. Chem. 2017, 292, 20196–20207. [Google Scholar] [CrossRef]
  147. Nagai, T.; Yamada, S.; Tominaga, T.; Ichikawa, M.; Miyawaki, A. Expanded dynamic range of fluorescent indicators for Ca(2+) by circularly permuted yellow fluorescent proteins. Proc. Natl. Acad. Sci. USA 2004, 101, 10554–10559. [Google Scholar] [CrossRef]
  148. Chiu, T.Y.; Yang, D.M. Intracellular Pb2+ content monitoring using a protein-based Pb2+ indicator. Toxicol. Sci. 2012, 126, 436–445. [Google Scholar] [CrossRef]
  149. Hubner, W.; Chen, P.; Del Portillo, A.; Liu, Y.; Gordon, R.E.; Chen, B.K. Sequence of human immunodeficiency virus type 1 (HIV-1) Gag localization and oligomerization monitored with live confocal imaging of a replication-competent, fluorescently tagged HIV-1. J. Virol. 2007, 81, 12596–12607. [Google Scholar] [CrossRef]
  150. Albertazzi, L.; Arosio, D.; Marchetti, L.; Ricci, F.; Beltram, F. Quantitative FRET analysis with the EGFP-mCherry fluorescent protein pair. Photochem. Photobiol. 2009, 85, 287–297. [Google Scholar] [CrossRef]
  151. Bhuckory, S.; Kays, J.C.; Dennis, A.M. In Vivo Biosensing Using Resonance Energy Transfer. Biosensors 2019, 9, 76. [Google Scholar] [CrossRef]
  152. Fontana, J.; Jurado, K.A.; Cheng, N.; Ly, N.L.; Fuchs, J.R.; Gorelick, R.J.; Engelman, A.N.; Steven, A.C. Distribution and Redistribution of HIV-1 Nucleocapsid Protein in Immature, Mature, and Integrase-Inhibited Virions: A Role for Integrase in Maturation. J. Virol. 2015, 89, 9765–9780. [Google Scholar] [CrossRef] [PubMed]
  153. Jurado, K.A.; Wang, H.; Slaughter, A.; Feng, L.; Kessl, J.J.; Koh, Y.; Wang, W.; Ballandras-Colas, A.; Patel, P.A.; Fuchs, J.R.; et al. Allosteric integrase inhibitor potency is determined through the inhibition of HIV-1 particle maturation. Proc. Natl. Acad. Sci. USA 2013, 110, 8690–8695. [Google Scholar] [CrossRef]
  154. Checkley, M.A.; Luttge, B.G.; Soheilian, F.; Nagashima, K.; Freed, E.O. The capsid-spacer peptide 1 Gag processing intermediate is a dominant-negative inhibitor of HIV-1 maturation. Virology 2010, 400, 137–144. [Google Scholar] [CrossRef] [PubMed]
  155. Tan, E.; Chin, C.S.H.; Lim, Z.F.S.; Ng, S.K. HEK293 Cell Line as a Platform to Produce Recombinant Proteins and Viral Vectors. Front. Bioeng. Biotechnol. 2021, 9, 796991. [Google Scholar] [CrossRef]
  156. Chen, B. Molecular Mechanism of HIV-1 Entry. Trends Microbiol. 2019, 27, 878–891. [Google Scholar] [CrossRef]
  157. Ma, X.; Lu, M.; Gorman, J.; Terry, D.S.; Hong, X.; Zhou, Z.; Zhao, H.; Altman, R.B.; Arthos, J.; Blanchard, S.C.; et al. HIV-1 Env trimer opens through an asymmetric intermediate in which individual protomers adopt distinct conformations. Elife 2018, 7, e34271. [Google Scholar] [CrossRef] [PubMed]
  158. Munro, J.B.; Gorman, J.; Ma, X.; Zhou, Z.; Arthos, J.; Burton, D.R.; Koff, W.C.; Courter, J.R.; Smith, A.B., 3rd; Kwong, P.D.; et al. Conformational dynamics of single HIV-1 envelope trimers on the surface of native virions. Science 2014, 346, 759–763. [Google Scholar] [CrossRef]
  159. Herschhorn, A.; Ma, X.; Gu, C.; Ventura, J.D.; Castillo-Menendez, L.; Melillo, B.; Terry, D.S.; Smith, A.B., 3rd; Blanchard, S.C.; Munro, J.B.; et al. Release of gp120 Restraints Leads to an Entry-Competent Intermediate State of the HIV-1 Envelope Glycoproteins. mBio 2016, 7, e01598-16. [Google Scholar] [CrossRef]
  160. Alsahafi, N.; Bakouche, N.; Kazemi, M.; Richard, J.; Ding, S.; Bhattacharyya, S.; Das, D.; Anand, S.P.; Prevost, J.; Tolbert, W.D.; et al. An Asymmetric Opening of HIV-1 Envelope Mediates Antibody-Dependent Cellular Cytotoxicity. Cell Host Microbe 2019, 25, 578–587.e5. [Google Scholar] [CrossRef]
  161. Checkley, M.A.; Luttge, B.G.; Freed, E.O. HIV-1 envelope glycoprotein biosynthesis, trafficking, and incorporation. J. Mol. Biol. 2011, 410, 582–608. [Google Scholar] [CrossRef]
  162. Hill, C.P.; Worthylake, D.; Bancroft, D.P.; Christensen, A.M.; Sundquist, W.I. Crystal structures of the trimeric human immunodeficiency virus type 1 matrix protein: Implications for membrane association and assembly. Proc. Natl. Acad. Sci. USA 1996, 93, 3099–3104. [Google Scholar] [CrossRef] [PubMed]
  163. Alfadhli, A.; Barklis, R.L.; Barklis, E. HIV-1 matrix organizes as a hexamer of trimers on membranes containing phosphatidylinositol-(4,5)-bisphosphate. Virology 2009, 387, 466–472. [Google Scholar] [CrossRef] [PubMed]
  164. Tedbury, P.R.; Ablan, S.D.; Freed, E.O. Global rescue of defects in HIV-1 envelope glycoprotein incorporation: Implications for matrix structure. PLoS Pathog. 2013, 9, e1003739. [Google Scholar] [CrossRef] [PubMed]
  165. Tedbury, P.R.; Novikova, M.; Ablan, S.D.; Freed, E.O. Biochemical evidence of a role for matrix trimerization in HIV-1 envelope glycoprotein incorporation. Proc. Natl. Acad. Sci. USA 2016, 113, E182–E190. [Google Scholar] [CrossRef] [PubMed]
  166. Wyma, D.J.; Jiang, J.; Shi, J.; Zhou, J.; Lineberger, J.E.; Miller, M.D.; Aiken, C. Coupling of human immunodeficiency virus type 1 fusion to virion maturation: A novel role of the gp41 cytoplasmic tail. J. Virol. 2004, 78, 3429–3435. [Google Scholar] [CrossRef]
  167. Murakami, T.; Ablan, S.; Freed, E.O.; Tanaka, Y. Regulation of human immunodeficiency virus type 1 Env-mediated membrane fusion by viral protease activity. J. Virol. 2004, 78, 1026–1031. [Google Scholar] [CrossRef]
  168. Chojnacki, J.; Staudt, T.; Glass, B.; Bingen, P.; Engelhardt, J.; Anders, M.; Schneider, J.; Müller, B.; Hell, S.W.; Kräusslich, H.G. Maturation-dependent HIV-1 surface protein redistribution revealed by fluorescence nanoscopy. Science 2012, 338, 524–528. [Google Scholar] [CrossRef]
  169. Li, W.; Qin, Z.; Nand, E.; Grunst, M.W.; Grover, J.R.; Bess, J.W., Jr.; Lifson, J.D.; Zwick, M.B.; Tagare, H.D.; Uchil, P.D.; et al. HIV-1 Env trimers asymmetrically engage CD4 receptors in membranes. Nature 2023, 623, 1026–1033. [Google Scholar] [CrossRef]
  170. Carlon-Andres, I.; Malinauskas, T.; Padilla-Parra, S. Structure dynamics of HIV-1 Env trimers on native virions engaged with living T cells. Commun. Biol. 2021, 4, 1228. [Google Scholar] [CrossRef]
  171. Kol, N.; Shi, Y.; Tsvitov, M.; Barlam, D.; Shneck, R.Z.; Kay, M.S.; Rousso, I. A stiffness switch in human immunodeficiency virus. Biophys. J. 2007, 92, 1777–1783. [Google Scholar] [CrossRef]
  172. Nieto-Garai, J.A.; Arboleya, A.; Otaegi, S.; Chojnacki, J.; Casas, J.; Fabrias, G.; Contreras, F.X.; Krausslich, H.G.; Lorizate, M. Cholesterol in the Viral Membrane is a Molecular Switch Governing HIV-1 Env Clustering. Adv. Sci. 2021, 8, 2003468. [Google Scholar] [CrossRef] [PubMed]
  173. Liao, Z.; Cimakasky, L.M.; Hampton, R.; Nguyen, D.H.; Hildreth, J.E. Lipid rafts and HIV pathogenesis: Host membrane cholesterol is required for infection by HIV type 1. AIDS Res. Hum. Retroviruses 2001, 17, 1009–1019. [Google Scholar] [CrossRef]
  174. Waheed, A.A.; Ablan, S.D.; Mankowski, M.K.; Cummins, J.E.; Ptak, R.G.; Schaffner, C.P.; Freed, E.O. Inhibition of HIV-1 replication by amphotericin B methyl ester: Selection for resistant variants. J. Biol. Chem. 2006, 281, 28699–28711. [Google Scholar] [CrossRef] [PubMed]
  175. Zhu, P.; Chertova, E.; Bess, J., Jr.; Lifson, J.D.; Arthur, L.O.; Liu, J.; Taylor, K.A.; Roux, K.H. Electron tomography analysis of envelope glycoprotein trimers on HIV and simian immunodeficiency virus virions. Proc. Natl. Acad. Sci. USA 2003, 100, 15812–15817. [Google Scholar] [CrossRef]
  176. Klein, J.S.; Bjorkman, P.J. Few and far between: How HIV may be evading antibody avidity. PLoS Pathog. 2010, 6, e1000908. [Google Scholar] [CrossRef] [PubMed]
  177. Brandenberg, O.F.; Magnus, C.; Regoes, R.R.; Trkola, A. The HIV-1 Entry Process: A Stoichiometric View. Trends Microbiol. 2015, 23, 763–774. [Google Scholar] [CrossRef]
  178. Weber, I.T.; Wang, Y.F.; Harrison, R.W. HIV Protease: Historical Perspective and Current Research. Viruses 2021, 13, 839. [Google Scholar] [CrossRef]
  179. Kohl, N.E.; Emini, E.A.; Schleif, W.A.; Davis, L.J.; Heimbach, J.C.; Dixon, R.A.; Scolnick, E.M.; Sigal, I.S. Active human immunodeficiency virus protease is required for viral infectivity. Proc. Natl. Acad. Sci. USA 1988, 85, 4686–4690. [Google Scholar] [CrossRef]
  180. Navia, M.A.; Fitzgerald, P.M.; McKeever, B.M.; Leu, C.T.; Heimbach, J.C.; Herber, W.K.; Sigal, I.S.; Darke, P.L.; Springer, J.P. Three-dimensional structure of aspartyl protease from human immunodeficiency virus HIV-1. Nature 1989, 337, 615–620. [Google Scholar] [CrossRef]
  181. Wlodawer, A.; Miller, M.; Jaskolski, M.; Sathyanarayana, B.K.; Baldwin, E.; Weber, I.T.; Selk, L.M.; Clawson, L.; Schneider, J.; Kent, S.B. Conserved folding in retroviral proteases: Crystal structure of a synthetic HIV-1 protease. Science 1989, 245, 616–621. [Google Scholar] [CrossRef]
  182. Darke, P.L.; Nutt, R.F.; Brady, S.F.; Garsky, V.M.; Ciccarone, T.M.; Leu, C.T.; Lumma, P.K.; Freidinger, R.M.; Veber, D.F.; Sigal, I.S. HIV-1 protease specificity of peptide cleavage is sufficient for processing of gag and pol polyproteins. Biochem. Biophys. Res. Commun. 1988, 156, 297–303. [Google Scholar] [CrossRef]
  183. Kotler, M.; Katz, R.A.; Danho, W.; Leis, J.; Skalka, A.M. Synthetic peptides as substrates and inhibitors of a retroviral protease. Proc. Natl. Acad. Sci. USA 1988, 85, 4185–4189. [Google Scholar] [CrossRef] [PubMed]
  184. Atta, M.G.; De Seigneux, S.; Lucas, G.M. Clinical Pharmacology in HIV Therapy. Clin. J. Am. Soc. Nephrol. 2019, 14, 435–444. [Google Scholar] [CrossRef]
  185. Loos, N.H.C.; Beijnen, J.H.; Schinkel, A.H. The Mechanism-Based Inactivation of CYP3A4 by Ritonavir: What Mechanism? Int. J. Mol. Sci. 2022, 23, 9866. [Google Scholar] [CrossRef] [PubMed]
  186. Nathan, B.; Bayley, J.; Waters, L.; Post, F.A. Cobicistat: A Novel Pharmacoenhancer for Co-Formulation with HIV Protease and Integrase Inhibitors. Infect. Dis. Ther. 2013, 2, 111–122. [Google Scholar] [CrossRef]
  187. Koh, Y.; Nakata, H.; Maeda, K.; Ogata, H.; Bilcer, G.; Devasamudram, T.; Kincaid, J.F.; Boross, P.; Wang, Y.F.; Tie, Y.; et al. Novel bis-tetrahydrofuranylurethane-containing nonpeptidic protease inhibitor (PI) UIC-94017 (TMC114) with potent activity against multi-PI-resistant human immunodeficiency virus in vitro. Antimicrob. Agents Chemother. 2003, 47, 3123–3129. [Google Scholar] [CrossRef]
  188. Spielvogel, E.; Lee, S.K.; Zhou, S.; Lockbaum, G.J.; Henes, M.; Sondgeroth, A.; Kosovrasti, K.; Nalivaika, E.A.; Ali, A.; Yilmaz, N.K.; et al. Selection of HIV-1 for resistance to fifth-generation protease inhibitors reveals two independent pathways to high-level resistance. Elife 2023, 12, e80328. [Google Scholar] [CrossRef] [PubMed]
  189. Ghosh, A.K.; Weber, I.T.; Mitsuya, H. Beyond darunavir: Recent development of next generation HIV-1 protease inhibitors to combat drug resistance. Chem. Commun. 2022, 58, 11762–11782. [Google Scholar] [CrossRef]
  190. Fujioka, T.; Kashiwada, Y.; Kilkuskie, R.E.; Cosentino, L.M.; Ballas, L.M.; Jiang, J.B.; Janzen, W.P.; Chen, I.S.; Lee, K.H. Anti-AIDS agents, 11. Betulinic acid and platanic acid as anti-HIV principles from Syzigium claviflorum, and the anti-HIV activity of structurally related triterpenoids. J. Nat. Prod. 1994, 57, 243–247. [Google Scholar] [CrossRef]
  191. Kanamoto, T.; Kashiwada, Y.; Kanbara, K.; Gotoh, K.; Yoshimori, M.; Goto, T.; Sano, K.; Nakashima, H. Anti-human immunodeficiency virus activity of YK-FH312 (a betulinic acid derivative), a novel compound blocking viral maturation. Antimicrob. Agents Chemother. 2001, 45, 1225–1230. [Google Scholar] [CrossRef]
  192. Kashiwada, Y.; Hashimoto, F.; Cosentino, L.M.; Chen, C.H.; Garrett, P.E.; Lee, K.H. Betulinic acid and dihydrobetulinic acid derivatives as potent anti-HIV agents. J. Med. Chem. 1996, 39, 1016–1017. [Google Scholar] [CrossRef]
  193. Li, F.; Goila-Gaur, R.; Salzwedel, K.; Kilgore, N.R.; Reddick, M.; Matallana, C.; Castillo, A.; Zoumplis, D.; Martin, D.E.; Orenstein, J.M.; et al. PA-457: A potent HIV inhibitor that disrupts core condensation by targeting a late step in Gag processing. Proc. Natl. Acad. Sci. USA 2003, 100, 13555–13560. [Google Scholar] [CrossRef] [PubMed]
  194. Schur, F.K.; Obr, M.; Hagen, W.J.; Wan, W.; Jakobi, A.J.; Kirkpatrick, J.M.; Sachse, C.; Krausslich, H.G.; Briggs, J.A. An atomic model of HIV-1 capsid-SP1 reveals structures regulating assembly and maturation. Science 2016, 353, 506–508. [Google Scholar] [CrossRef]
  195. Wagner, J.M.; Zadrozny, K.K.; Chrustowicz, J.; Purdy, M.D.; Yeager, M.; Ganser-Pornillos, B.K.; Pornillos, O. Crystal structure of an HIV assembly and maturation switch. Elife 2016, 5, e17063. [Google Scholar] [CrossRef]
  196. Purdy, M.D.; Shi, D.; Chrustowicz, J.; Hattne, J.; Gonen, T.; Yeager, M. MicroED structures of HIV-1 Gag CTD-SP1 reveal binding interactions with the maturation inhibitor bevirimat. Proc. Natl. Acad. Sci. USA 2018, 115, 13258–13263. [Google Scholar] [CrossRef] [PubMed]
  197. Gupta, S.; Louis, J.M.; Tycko, R. Effects of an HIV-1 maturation inhibitor on the structure and dynamics of CA-SP1 junction helices in virus-like particles. Proc. Natl. Acad. Sci. USA 2020, 117, 10286–10293. [Google Scholar] [CrossRef]
  198. Adamson, C.S.; Ablan, S.D.; Boeras, I.; Goila-Gaur, R.; Soheilian, F.; Nagashima, K.; Li, F.; Salzwedel, K.; Sakalian, M.; Wild, C.T.; et al. In vitro resistance to the human immunodeficiency virus type 1 maturation inhibitor PA-457 (Bevirimat). J. Virol. 2006, 80, 10957–10971. [Google Scholar] [CrossRef] [PubMed]
  199. Li, F.; Zoumplis, D.; Matallana, C.; Kilgore, N.R.; Reddick, M.; Yunus, A.S.; Adamson, C.S.; Salzwedel, K.; Martin, D.E.; Allaway, G.P.; et al. Determinants of activity of the HIV-1 maturation inhibitor PA-457. Virology 2006, 356, 217–224. [Google Scholar] [CrossRef]
  200. Zhou, J.; Chen, C.H.; Aiken, C. Human immunodeficiency virus type 1 resistance to the small molecule maturation inhibitor 3-O-(3′,3′-dimethylsuccinyl)-betulinic acid is conferred by a variety of single amino acid substitutions at the CA-SP1 cleavage site in Gag. J Virol 2006, 80, 12095–12101. [Google Scholar] [CrossRef]
  201. Zhou, J.; Yuan, X.; Dismuke, D.; Forshey, B.M.; Lundquist, C.; Lee, K.H.; Aiken, C.; Chen, C.H. Small-molecule inhibition of human immunodeficiency virus type 1 replication by specific targeting of the final step of virion maturation. J. Virol. 2004, 78, 922–929. [Google Scholar] [CrossRef]
  202. Nguyen, A.T.; Feasley, C.L.; Jackson, K.W.; Nitz, T.J.; Salzwedel, K.; Air, G.M.; Sakalian, M. The prototype HIV-1 maturation inhibitor, bevirimat, binds to the CA-SP1 cleavage site in immature Gag particles. Retrovirology 2011, 8, 101. [Google Scholar] [CrossRef] [PubMed]
  203. Zhou, J.; Chen, C.H.; Aiken, C. The sequence of the CA-SP1 junction accounts for the differential sensitivity of HIV-1 and SIV to the small molecule maturation inhibitor 3-O-3′,3′-dimethylsuccinyl-betulinic acid. Retrovirology 2004, 1, 15. [Google Scholar] [CrossRef]
  204. Zhou, J.; Huang, L.; Hachey, D.L.; Chen, C.H.; Aiken, C. Inhibition of HIV-1 maturation via drug association with the viral Gag protein in immature HIV-1 particles. J. Biol. Chem. 2005, 280, 42149–42155. [Google Scholar] [CrossRef]
  205. Sarkar, S.; Zadrozny, K.K.; Zadorozhnyi, R.; Russell, R.W.; Quinn, C.M.; Kleinpeter, A.; Ablan, S.; Meshkin, H.; Perilla, J.R.; Freed, E.O.; et al. Structural basis of HIV-1 maturation inhibitor binding and activity. Nat. Commun. 2023, 14, 1237. [Google Scholar] [CrossRef]
  206. Prabu-Jeyabalan, M.; Nalivaika, E.; Schiffer, C.A. Substrate shape determines specificity of recognition for HIV-1 protease: Analysis of crystal structures of six substrate complexes. Structure 2002, 10, 369–381. [Google Scholar] [CrossRef] [PubMed]
  207. Smith, P.F.; Ogundele, A.; Forrest, A.; Wilton, J.; Salzwedel, K.; Doto, J.; Allaway, G.P.; Martin, D.E. Phase I and II study of the safety, virologic effect, and pharmacokinetics/pharmacodynamics of single-dose 3-o-(3′,3′-dimethylsuccinyl)betulinic acid (bevirimat) against human immunodeficiency virus infection. Antimicrob. Agents Chemother. 2007, 51, 3574–3581. [Google Scholar] [CrossRef]
  208. DeJesus, E.; Harward, S.; Jewell, R.C.; Johnson, M.; Dumont, E.; Wilches, V.; Halliday, F.; Talarico, C.L.; Jeffrey, J.; Gan, J.; et al. A Phase IIa Study Evaluating Safety, Pharmacokinetics, and Antiviral Activity of GSK2838232, a Novel, Second-generation Maturation Inhibitor, in Participants With Human Immunodeficiency Virus Type 1 Infection. Clin. Infect. Dis. 2020, 71, 1255–1262. [Google Scholar] [CrossRef] [PubMed]
  209. Hwang, C.; Schurmann, D.; Sobotha, C.; Boffito, M.; Sevinsky, H.; Ray, N.; Ravindran, P.; Xiao, H.; Keicher, C.; Huser, A.; et al. Antiviral Activity, Safety, and Exposure-Response Relationships of GSK3532795, a Second-Generation Human Immunodeficiency Virus Type 1 Maturation Inhibitor, Administered as Monotherapy or in Combination With Atazanavir With or Without Ritonavir in a Phase 2a Randomized, Dose-Ranging, Controlled Trial (AI468002). Clin. Infect. Dis. 2017, 65, 442–452. [Google Scholar]
  210. Morales-Ramirez, J.; Bogner, J.R.; Molina, J.M.; Lombaard, J.; Dicker, I.B.; Stock, D.A.; DeGrosky, M.; Gartland, M.; Pene Dumitrescu, T.; Min, S.; et al. Safety, efficacy, and dose response of the maturation inhibitor GSK3532795 (formerly known as BMS-955176) plus tenofovir/emtricitabine once daily in treatment-naive HIV-1-infected adults: Week 24 primary analysis from a randomized Phase IIb trial. PLoS ONE 2018, 13, e0205368. [Google Scholar] [CrossRef]
  211. Abstracts of the 19th European AIDS Conference (#EACS2023), October 18–21, 2023, Warsaw, Poland. HIV Med. 2023, 24 (Suppl. 5), 3–788.
  212. Benn, P.D.; Zhang, Y.; Kahl, L.; Greene, T.J.; Bainbridge, V.; Fishman, C.; Wen, B.; Gartland, M. A phase I, first-in-human study investigating the safety, tolerability, and pharmacokinetics of the maturation inhibitor GSK3739937. Pharmacol. Res. Perspect. 2023, 11, e01093. [Google Scholar] [CrossRef] [PubMed]
  213. Yant, S.R.; Mulato, A.; Hansen, D.; Tse, W.C.; Niedziela-Majka, A.; Zhang, J.R.; Stepan, G.J.; Jin, D.; Wong, M.H.; Perreira, J.M.; et al. A highly potent long-acting small-molecule HIV-1 capsid inhibitor with efficacy in a humanized mouse model. Nat. Med. 2019, 25, 1377–1384. [Google Scholar] [CrossRef] [PubMed]
  214. Faysal, K.M.R.; Walsh, J.C.; Renner, N.; Marquez, C.L.; Shah, V.B.; Tuckwell, A.J.; Christie, M.P.; Parker, M.W.; Turville, S.G.; Towers, G.J.; et al. Pharmacologic hyperstabilisation of the HIV-1 capsid lattice induces capsid failure. Elife 2024, 13, e83605. [Google Scholar] [PubMed]
  215. Marquis, K.A.; Everett, J.; Cantu, A.; McFarland, A.; Sherrill-Mix, S.; Krystal, M.; Parcella, K.; Gillis, E.; Fridell, R.A.; Bushman, F.D. The HIV-1 Capsid-Targeted Inhibitor GSK878 Alters Selection of Target Sites for HIV DNA Integration. AIDS Res. Hum. Retroviruses 2024, 40, 114–126. [Google Scholar] [CrossRef]
  216. McFadden, W.M.; Snyder, A.A.; Kirby, K.A.; Tedbury, P.R.; Raj, M.; Wang, Z.; Sarafianos, S.G. Rotten to the core: Antivirals targeting the HIV-1 capsid core. Retrovirology 2021, 18, 41. [Google Scholar] [CrossRef]
  217. Zhou, J.; Price, A.J.; Halambage, U.D.; James, L.C.; Aiken, C. HIV-1 Resistance to the Capsid-Targeting Inhibitor PF74 Results in Altered Dependence on Host Factors Required for Virus Nuclear Entry. J. Virol. 2015, 89, 9068–9079. [Google Scholar] [CrossRef]
  218. Bhattacharya, A.; Alam, S.L.; Fricke, T.; Zadrozny, K.; Sedzicki, J.; Taylor, A.B.; Demeler, B.; Pornillos, O.; Ganser-Pornillos, B.K.; Diaz-Griffero, F.; et al. Structural basis of HIV-1 capsid recognition by PF74 and CPSF6. Proc. Natl. Acad. Sci. USA 2014, 111, 18625–18630. [Google Scholar] [CrossRef]
  219. Lamorte, L.; Titolo, S.; Lemke, C.T.; Goudreau, N.; Mercier, J.F.; Wardrop, E.; Shah, V.B.; von Schwedler, U.K.; Langelier, C.; Banik, S.S.; et al. Discovery of novel small-molecule HIV-1 replication inhibitors that stabilize capsid complexes. Antimicrob. Agents Chemother. 2013, 57, 4622–4631. [Google Scholar] [CrossRef]
  220. Blair, W.S.; Pickford, C.; Irving, S.L.; Brown, D.G.; Anderson, M.; Bazin, R.; Cao, J.; Ciaramella, G.; Isaacson, J.; Jackson, L.; et al. HIV capsid is a tractable target for small molecule therapeutic intervention. PLoS Pathog. 2010, 6, e1001220. [Google Scholar] [CrossRef]
  221. Wang, C.; Huang, H.; Mallon, K.; Valera, L.; Parcella, K.; Cockett, M.I.; Kadow, J.F.; Gillis, E.P.; Krystal, M.; Fridell, R.A. Antiviral Properties of HIV-1 Capsid Inhibitor GSK878. Antimicrob. Agents Chemother. 2023, 67, e0169422. [Google Scholar] [CrossRef]
  222. Gillis, E.P.; Parcella, K.; Bowsher, M.; Cook, J.H.; Iwuagwu, C.; Naidu, B.N.; Patel, M.; Peese, K.; Huang, H.; Valera, L.; et al. Potent Long-Acting Inhibitors Targeting the HIV-1 Capsid Based on a Versatile Quinazolin-4-one Scaffold. J. Med. Chem. 2023, 66, 1941–1954. [Google Scholar] [CrossRef]
  223. Gupta, S.K.; Berhe, M.; Crofoot, G.; Benson, P.; Ramgopal, M.; Sims, J.; McDonald, C.; Ruane, P.; Sanchez, W.E.; Scribner, A.; et al. Lenacapavir administered every 26 weeks or daily in combination with oral daily antiretroviral therapy for initial treatment of HIV: A randomised, open-label, active-controlled, phase 2 trial. Lancet HIV 2023, 10, e15–e23. [Google Scholar] [CrossRef] [PubMed]
  224. Margot, N.; Pennetzdorfer, N.; Naik, V.; Rhee, M.; Callebaut, C. Cross-resistance to entry inhibitors and lenacapavir resistance through Week 52 in study CAPELLA. Antivir. Ther. 2023, 28, 13596535231220754. [Google Scholar] [CrossRef] [PubMed]
  225. Margot, N.A.; Naik, V.; VanderVeen, L.; Anoshchenko, O.; Singh, R.; Dvory-Sobol, H.; Rhee, M.S.; Callebaut, C. Resistance Analyses in Highly Treatment-Experienced People with Human Immunodeficiency Virus (HIV) Treated with the Novel Capsid HIV Inhibitor Lenacapavir. J. Infect. Dis. 2022, 226, 1985–1991. [Google Scholar] [CrossRef] [PubMed]
  226. Segal-Maurer, S.; DeJesus, E.; Stellbrink, H.J.; Castagna, A.; Richmond, G.J.; Sinclair, G.I.; Siripassorn, K.; Ruane, P.J.; Berhe, M.; Wang, H.; et al. Capsid Inhibition with Lenacapavir in Multidrug-Resistant HIV-1 Infection. N. Engl. J. Med. 2022, 386, 1793–1803. [Google Scholar] [CrossRef] [PubMed]
  227. Orkin, C.; Schapiro, J.M.; Perno, C.F.; Kuritzkes, D.R.; Patel, P.; DeMoor, R.; Dorey, D.; Wang, Y.; Han, K.; Van Eygen, V.; et al. Expanded Multivariable Models to Assist Patient Selection for Long-Acting Cabotegravir + Rilpivirine Treatment: Clinical Utility of a Combination of Patient, Drug Concentration, and Viral Factors Associated With Virologic Failure. Clin. Infect. Dis. 2023, 77, 1423–1431. [Google Scholar] [CrossRef]
  228. Cutrell, A.G.; Schapiro, J.M.; Perno, C.F.; Kuritzkes, D.R.; Quercia, R.; Patel, P.; Polli, J.W.; Dorey, D.; Wang, Y.; Wu, S.; et al. Exploring predictors of HIV-1 virologic failure to long-acting cabotegravir and rilpivirine: A multivariable analysis. AIDS 2021, 35, 1333–1342. [Google Scholar] [CrossRef]
  229. Bekker, L.G.; Das, M.; Abdool Karim, Q.; Ahmed, K.; Batting, J.; Brumskine, W.; Gill, K.; Harkoo, I.; Jaggernath, M.; Kigozi, G.; et al. Twice-Yearly Lenacapavir or Daily F/TAF for HIV Prevention in Cisgender Women. N. Engl. J. Med. 2024; ahead of print. [Google Scholar] [CrossRef] [PubMed]
  230. Landovitz, R.J.; Donnell, D.; Clement, M.E.; Hanscom, B.; Cottle, L.; Coelho, L.; Cabello, R.; Chariyalertsak, S.; Dunne, E.F.; Frank, I.; et al. Cabotegravir for HIV Prevention in Cisgender Men and Transgender Women. N. Engl. J. Med. 2021, 385, 595–608. [Google Scholar] [CrossRef]
  231. Delany-Moretlwe, S.; Hughes, J.P.; Bock, P.; Ouma, S.G.; Hunidzarira, P.; Kalonji, D.; Kayange, N.; Makhema, J.; Mandima, P.; Mathew, C.; et al. Cabotegravir for the prevention of HIV-1 in women: Results from HPTN 084, a phase 3, randomised clinical trial. Lancet 2022, 399, 1779–1789. [Google Scholar] [CrossRef]
  232. Fujiwara, T.; Mizuuchi, K. Retroviral DNA integration: Structure of an integration intermediate. Cell 1988, 54, 497–504. [Google Scholar] [CrossRef]
  233. Brown, P.O.; Bowerman, B.; Varmus, H.E.; Bishop, J.M. Retroviral integration: Structure of the initial covalent product and its precursor, and a role for the viral IN protein. Proc. Natl. Acad. Sci. USA 1989, 86, 2525–2529. [Google Scholar] [CrossRef] [PubMed]
  234. Roth, M.J.; Schwartzberg, P.L.; Goff, S.P. Structure of the termini of DNA intermediates in the integration of retroviral DNA: Dependence on IN function and terminal DNA sequence. Cell 1989, 58, 47–54. [Google Scholar] [CrossRef] [PubMed]
  235. Sherman, P.A.; Fyfe, J.A. Human immunodeficiency virus integration protein expressed in Escherichia coli possesses selective DNA cleaving activity. Proc. Natl. Acad. Sci. USA 1990, 87, 5119–5123. [Google Scholar] [CrossRef]
  236. Zhao, A.V.; Crutchley, R.D.; Guduru, R.C.; Ton, K.; Lam, T.; Min, A.C. A clinical review of HIV integrase strand transfer inhibitors (INSTIs) for the prevention and treatment of HIV-1 infection. Retrovirology 2022, 19, 22. [Google Scholar] [CrossRef]
  237. Kessl, J.J.; Kutluay, S.B.; Townsend, D.; Rebensburg, S.; Slaughter, A.; Larue, R.C.; Shkriabai, N.; Bakouche, N.; Fuchs, J.R.; Bieniasz, P.D.; et al. HIV-1 Integrase Binds the Viral RNA Genome and Is Essential during Virion Morphogenesis. Cell 2016, 166, 1257–1268.e12. [Google Scholar] [CrossRef]
  238. Allen, P.; Worland, S.; Gold, L. Isolation of high-affinity RNA ligands to HIV-1 integrase from a random pool. Virology 1995, 209, 327–336. [Google Scholar] [CrossRef]
  239. Engelman, A. In vivo analysis of retroviral integrase structure and function. Adv. Virus. Res. 1999, 52, 411–426. [Google Scholar] [PubMed]
  240. Engelman, A.; Englund, G.; Orenstein, J.M.; Martin, M.A.; Craigie, R. Multiple effects of mutations in human immunodeficiency virus type 1 integrase on viral replication. J. Virol. 1995, 69, 2729–2736. [Google Scholar] [CrossRef]
  241. Llano, M.; Vanegas, M.; Hutchins, N.; Thompson, D.; Delgado, S.; Poeschla, E.M. Identification and characterization of the chromatin-binding domains of the HIV-1 integrase interactor LEDGF/p75. J. Mol. Biol. 2006, 360, 760–773. [Google Scholar] [CrossRef]
  242. Ciuffi, A.; Llano, M.; Poeschla, E.; Hoffmann, C.; Leipzig, J.; Shinn, P.; Ecker, J.R.; Bushman, F. A role for LEDGF/p75 in targeting HIV DNA integration. Nat. Med. 2005, 11, 1287–1289. [Google Scholar] [CrossRef] [PubMed]
  243. Maertens, G.; Cherepanov, P.; Pluymers, W.; Busschots, K.; De Clercq, E.; Debyser, Z.; Engelborghs, Y. LEDGF/p75 is essential for nuclear and chromosomal targeting of HIV-1 integrase in human cells. J. Biol. Chem. 2003, 278, 33528–33539. [Google Scholar] [CrossRef]
  244. Cherepanov, P.; Maertens, G.; Proost, P.; Devreese, B.; Van Beeumen, J.; Engelborghs, Y.; De Clercq, E.; Debyser, Z. HIV-1 integrase forms stable tetramers and associates with LEDGF/p75 protein in human cells. J. Biol. Chem. 2003, 278, 372–381. [Google Scholar] [CrossRef] [PubMed]
  245. Desimmie, B.A.; Demeulemeester, J.; Christ, F.; Debyser, Z. Rational design of LEDGINs as first allosteric integrase inhibitors for the treatment of HIV infection. Drug Discov. Today Technol. 2013, 10, e517–e522. [Google Scholar] [CrossRef]
  246. Le Rouzic, E.; Bonnard, D.; Chasset, S.; Bruneau, J.M.; Chevreuil, F.; Le Strat, F.; Nguyen, J.; Beauvoir, R.; Amadori, C.; Brias, J.; et al. Dual inhibition of HIV-1 replication by integrase-LEDGF allosteric inhibitors is predominant at the post-integration stage. Retrovirology 2013, 10, 144. [Google Scholar] [CrossRef]
  247. Sharma, A.; Slaughter, A.; Jena, N.; Feng, L.; Kessl, J.J.; Fadel, H.J.; Malani, N.; Male, F.; Wu, L.; Poeschla, E.; et al. A new class of multimerization selective inhibitors of HIV-1 integrase. PLoS Pathog. 2014, 10, e1004171. [Google Scholar] [CrossRef]
  248. Fader, L.D.; Malenfant, E.; Parisien, M.; Carson, R.; Bilodeau, F.; Landry, S.; Pesant, M.; Brochu, C.; Morin, S.; Chabot, C.; et al. Discovery of BI 224436, a Noncatalytic Site Integrase Inhibitor (NCINI) of HIV-1. ACS Med. Chem. Lett. 2014, 5, 422–427. [Google Scholar] [CrossRef]
  249. Christ, F.; Voet, A.; Marchand, A.; Nicolet, S.; Desimmie, B.A.; Marchand, D.; Bardiot, D.; Van der Veken, N.J.; Van Remoortel, B.; Strelkov, S.V.; et al. Rational design of small-molecule inhibitors of the LEDGF/p75-integrase interaction and HIV replication. Nat. Chem. Biol. 2010, 6, 442–448. [Google Scholar] [CrossRef] [PubMed]
  250. Gupta, K.; Turkki, V.; Sherrill-Mix, S.; Hwang, Y.; Eilers, G.; Taylor, L.; McDanal, C.; Wang, P.; Temelkoff, D.; Nolte, R.T.; et al. Structural Basis for Inhibitor-Induced Aggregation of HIV Integrase. PLoS Biol. 2016, 14, e1002584. [Google Scholar] [CrossRef]
  251. Gupta, K.; Brady, T.; Dyer, B.M.; Malani, N.; Hwang, Y.; Male, F.; Nolte, R.T.; Wang, L.; Velthuisen, E.; Jeffrey, J.; et al. Allosteric inhibition of human immunodeficiency virus integrase: Late block during viral replication and abnormal multimerization involving specific protein domains. J. Biol. Chem. 2014, 289, 20477–20488. [Google Scholar] [CrossRef]
  252. Balakrishnan, M.; Yant, S.R.; Tsai, L.; O’Sullivan, C.; Bam, R.A.; Tsai, A.; Niedziela-Majka, A.; Stray, K.M.; Sakowicz, R.; Cihlar, T. Non-catalytic site HIV-1 integrase inhibitors disrupt core maturation and induce a reverse transcription block in target cells. PLoS ONE 2013, 8, e74163. [Google Scholar]
  253. Eilers, G.; Gupta, K.; Allen, A.; Montermoso, S.; Murali, H.; Sharp, R.; Hwang, Y.; Bushman, F.D.; Van Duyne, G. Structure of a HIV-1 IN-Allosteric inhibitor complex at 2.93 A resolution: Routes to inhibitor optimization. PLoS Pathog. 2023, 19, e1011097. [Google Scholar] [CrossRef]
  254. Singer, M.R.; Dinh, T.; Levintov, L.; Annamalai, A.S.; Rey, J.S.; Briganti, L.; Cook, N.J.; Pye, V.E.; Taylor, I.A.; Kim, K.; et al. The Drug-Induced Interface That Drives HIV-1 Integrase Hypermultimerization and Loss of Function. mBio 2023, 14, e0356022. [Google Scholar] [CrossRef] [PubMed]
  255. Bonnard, D.; Le Rouzic, E.; Singer, M.R.; Yu, Z.; Le Strat, F.; Batisse, C.; Batisse, J.; Amadori, C.; Chasset, S.; Pye, V.E.; et al. Biological and Structural Analyses of New Potent Allosteric Inhibitors of HIV-1 Integrase. Antimicrob. Agents Chemother. 2023, 67, e0046223. [Google Scholar] [CrossRef]
  256. Maehigashi, T.; Ahn, S.; Kim, U.I.; Lindenberger, J.; Oo, A.; Koneru, P.C.; Mahboubi, B.; Engelman, A.N.; Kvaratskhelia, M.; Kim, K.; et al. A highly potent and safe pyrrolopyridine-based allosteric HIV-1 integrase inhibitor targeting host LEDGF/p75-integrase interaction site. PLoS Pathog. 2021, 17, e1009671. [Google Scholar] [CrossRef]
  257. Hitchcock, A.M.; Kufel, W.D.; Dwyer, K.A.M.; Sidman, E.F. Lenacapavir: A novel injectable HIV-1 capsid inhibitor. Int. J. Antimicrob. Agents 2024, 63, 107009. [Google Scholar] [CrossRef] [PubMed]
  258. Cihlar, T.; Fordyce, M. Current status and prospects of HIV treatment. Curr. Opin. Virol. 2016, 18, 50–56. [Google Scholar] [CrossRef] [PubMed]
  259. Barthold, D.; Saldarriaga, E.M.; Brah, A.T.; Hauber, B.; Banerjee, P.; Fuller, S.M.; McCaslin, D.; Moldoveanu, A.M.; Marconi, V.C.; Simoni, J.M.; et al. Preference for daily oral pills over long-acting antiretroviral therapy options among people with HIV. AIDS 2023, 37, 1545–1553. [Google Scholar] [CrossRef]
  260. Swanstrom, A.E.; Gorelick, R.J.; Welker, J.L.; Schmidt, F.; Lu, B.; Wang, K.; Rowe, W.; Breed, M.W.; Killoran, K.E.; Kramer, J.A.; et al. Long-acting lenacapavir protects macaques against intravenous challenge with simian-tropic HIV. EBioMedicine 2023, 95, 104764. [Google Scholar] [CrossRef]
Figure 1. HIV Replication Cycle. Schematic representation of the HIV replication cycle. (A) HIV-1 Env binds to plasma membrane receptors, CD4, and CCR5 or CXCR4 coreceptors. (B) Upon interactions between Env and host receptors, viral and host membrane fusion occurs, releasing the viral core into the host cell cytoplasm. (C) Reverse transcription is initiated shortly after virus entry. The HIV-1 CA core is transported toward the nucleus through the host cell’s microtubule network by dynein and kinesin motor proteins. (D) Reverse-transcribed intermediate products inside the intact core are imported into the nucleus through the nuclear pore. (E) Upon completion of reverse transcription, viral DNA induces uncoating of the core structure in the nucleus, and IN mediates viral DNA integration into the host genome. (F) Proviral DNA is transcribed by the host machinery, enhanced by HIV-1 Tat. (G) Viral mRNAs are translated by free ribosomes or ER-associated ribosomes specific for spliced forms encoding the env gene. (H) Gag and Gag-Pol polyproteins assemble at the plasma membrane and recruit Env in progeny virions. (I) Full-length viral RNAs migrate to the plasma membrane and are packaged into progeny virions. (J) Progeny virions bud from virus-producer cells. (K) PR cleaves Gag and Gag-Pol in immature virions, releasing CA proteins that form the conical core harboring viral gRNA. Viral Env distribution changes during maturation. The currently developed antivirals disrupt multiple steps of the viral replication cycle, including (B) Fusion, (C) Reverse Transcription, (D) Nuclear Transport/Import, (E) Integration, (H) Assembly, and (K) Maturation.
Figure 1. HIV Replication Cycle. Schematic representation of the HIV replication cycle. (A) HIV-1 Env binds to plasma membrane receptors, CD4, and CCR5 or CXCR4 coreceptors. (B) Upon interactions between Env and host receptors, viral and host membrane fusion occurs, releasing the viral core into the host cell cytoplasm. (C) Reverse transcription is initiated shortly after virus entry. The HIV-1 CA core is transported toward the nucleus through the host cell’s microtubule network by dynein and kinesin motor proteins. (D) Reverse-transcribed intermediate products inside the intact core are imported into the nucleus through the nuclear pore. (E) Upon completion of reverse transcription, viral DNA induces uncoating of the core structure in the nucleus, and IN mediates viral DNA integration into the host genome. (F) Proviral DNA is transcribed by the host machinery, enhanced by HIV-1 Tat. (G) Viral mRNAs are translated by free ribosomes or ER-associated ribosomes specific for spliced forms encoding the env gene. (H) Gag and Gag-Pol polyproteins assemble at the plasma membrane and recruit Env in progeny virions. (I) Full-length viral RNAs migrate to the plasma membrane and are packaged into progeny virions. (J) Progeny virions bud from virus-producer cells. (K) PR cleaves Gag and Gag-Pol in immature virions, releasing CA proteins that form the conical core harboring viral gRNA. Viral Env distribution changes during maturation. The currently developed antivirals disrupt multiple steps of the viral replication cycle, including (B) Fusion, (C) Reverse Transcription, (D) Nuclear Transport/Import, (E) Integration, (H) Assembly, and (K) Maturation.
Viruses 16 01423 g001
Figure 2. Gag and Gag-Pol polyprotein processing by PR. Schematic representation of (A) Gag and (B) Gag-Pol polyprotein cleavage pattern by PR. (A) The first cleavage occurs between the SP1-NC region within the RQA|NFLG protease recognition sequence, followed by cleavages between MA-CA within SQNY|PIV and SP2-p6 within PGNF|LQS. Finally, SP1 and SP2 short peptides are removed from CA and NC, respectively, within the ARVL|AEA and RQAN|FLG protease recognition sequences. (B) The Gag-Pol polyprotein is formed by a frameshifting event at the NC/SP1 boundary, which translates the p6 domain in an alternative reading frame (Transframe/TF). The initial cleavage events separate Gag and Pol regions by cleaving between TF and PR within the SFNF|PQIT protease recognition sequence, carried out by the poorly active PR precursor. This process frees the protease from the rest of the precursor, enabling the ordered processing of the Gag polyprotein. (C) FRET indicator fluorescent proteins inserted between the MA and CA domains, bridged by the SQNY|PIV protease cleavage sequence. CFP and YFP separate when MA-CA is cleaved by PR.
Figure 2. Gag and Gag-Pol polyprotein processing by PR. Schematic representation of (A) Gag and (B) Gag-Pol polyprotein cleavage pattern by PR. (A) The first cleavage occurs between the SP1-NC region within the RQA|NFLG protease recognition sequence, followed by cleavages between MA-CA within SQNY|PIV and SP2-p6 within PGNF|LQS. Finally, SP1 and SP2 short peptides are removed from CA and NC, respectively, within the ARVL|AEA and RQAN|FLG protease recognition sequences. (B) The Gag-Pol polyprotein is formed by a frameshifting event at the NC/SP1 boundary, which translates the p6 domain in an alternative reading frame (Transframe/TF). The initial cleavage events separate Gag and Pol regions by cleaving between TF and PR within the SFNF|PQIT protease recognition sequence, carried out by the poorly active PR precursor. This process frees the protease from the rest of the precursor, enabling the ordered processing of the Gag polyprotein. (C) FRET indicator fluorescent proteins inserted between the MA and CA domains, bridged by the SQNY|PIV protease cleavage sequence. CFP and YFP separate when MA-CA is cleaved by PR.
Viruses 16 01423 g002
Figure 3. Structural representation of the HIV-1 CA Core. Approximately 250 hexamers and exactly 12 pentamers of CA assemble into a conical core (PDB: 3J3Y) [43,44]. The N-terminal domain from residues 1–145, the linker between residues 146–150, and the C-terminal domain from residues 151–231 of CA are colored in brown, gray, and bright orange, respectively. Twelve pentamers are specifically highlighted in sky blue. IP6 (green) is incorporated into both capsid hexamers (PDB: 7ZUD) [44,45] and pentamers (PDB: 7URN) [45]. In the absence of IP6, IP5 (blue) is incorporated into the capsid hexamers (PDB: 6R8C) [46]. NUP153 (pink in PDB: 4U0C) [47,48]), CPSF6 (pale green in PDB: 7ZUD), and LEN (green in PDB: 6V2F) [49] associated with the capsid hexamer are highlighted. To dynamically visualize these host proteins and small molecule interactions with the CA hexamer, the angles are rotated by 90 degrees from the outer surface angle on the x-axis.
Figure 3. Structural representation of the HIV-1 CA Core. Approximately 250 hexamers and exactly 12 pentamers of CA assemble into a conical core (PDB: 3J3Y) [43,44]. The N-terminal domain from residues 1–145, the linker between residues 146–150, and the C-terminal domain from residues 151–231 of CA are colored in brown, gray, and bright orange, respectively. Twelve pentamers are specifically highlighted in sky blue. IP6 (green) is incorporated into both capsid hexamers (PDB: 7ZUD) [44,45] and pentamers (PDB: 7URN) [45]. In the absence of IP6, IP5 (blue) is incorporated into the capsid hexamers (PDB: 6R8C) [46]. NUP153 (pink in PDB: 4U0C) [47,48]), CPSF6 (pale green in PDB: 7ZUD), and LEN (green in PDB: 6V2F) [49] associated with the capsid hexamer are highlighted. To dynamically visualize these host proteins and small molecule interactions with the CA hexamer, the angles are rotated by 90 degrees from the outer surface angle on the x-axis.
Viruses 16 01423 g003
Figure 4. Schematic, TEM, and fluorescence images of HIV-1 Gag-iFRET viruses. While immature virions, depicted by both schematic and TEM images, exhibit a higher FRET signal (red in the ratio color; yellow arrow), mature virions show a low FRET signal (blue in the ratio color; white arrow). Under electron microscopy, HIV-1 immature virions are characterized by an incomplete, spherical structure, while mature virions display a conical core indicative of successful proteolytic processing and proper assembly of the capsid proteins. Under fluorescence microscopy, immature virions express a red signal, which is reduced in mature virions in ratio view. This results in mature virions emitting a blue and green signal. Several drugs, including PIs, MIs, CAIs, and ALLINIs, interrupt virion maturation.
Figure 4. Schematic, TEM, and fluorescence images of HIV-1 Gag-iFRET viruses. While immature virions, depicted by both schematic and TEM images, exhibit a higher FRET signal (red in the ratio color; yellow arrow), mature virions show a low FRET signal (blue in the ratio color; white arrow). Under electron microscopy, HIV-1 immature virions are characterized by an incomplete, spherical structure, while mature virions display a conical core indicative of successful proteolytic processing and proper assembly of the capsid proteins. Under fluorescence microscopy, immature virions express a red signal, which is reduced in mature virions in ratio view. This results in mature virions emitting a blue and green signal. Several drugs, including PIs, MIs, CAIs, and ALLINIs, interrupt virion maturation.
Viruses 16 01423 g004
Table 1. Antivirals targeting virion maturation.
Table 1. Antivirals targeting virion maturation.
ClassStructureStage of Development
Protease Inhibitors (PIs):
Saquinavir (SQV)Viruses 16 01423 i001FDA approved, 1995
Darunavir (DRV)Viruses 16 01423 i002FDA approved, 2006
Maturation Inhibitors (MIs):
Bevirimat (BVM)Viruses 16 01423 i003Clinical, terminated
GSK2838232Viruses 16 01423 i004Clinical, terminated
GSK3532795Viruses 16 01423 i005Clinical, terminated
GSK3640254Viruses 16 01423 i006Clinical, terminated
Capsid Inhibitors (CAIs):
PF-3450074 (PF74)Viruses 16 01423 i007Discovery
BI-1Viruses 16 01423 i008Discovery
BI-2Viruses 16 01423 i009Discovery
GSK878Viruses 16 01423 i010Pre-clinical
GS-CA1Viruses 16 01423 i011Pre-clinical
Lenacapavir (LEN)Viruses 16 01423 i012FDA approved, 2022
Allosteric Integrase Inhibitors (ALLINIs):
Pirmitegravir (PIR)Viruses 16 01423 i013Clinical, Phase 2
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

McGraw, A.; Hillmer, G.; Medehincu, S.M.; Hikichi, Y.; Gagliardi, S.; Narayan, K.; Tibebe, H.; Marquez, D.; Mei Bose, L.; Keeting, A.; et al. Exploring HIV-1 Maturation: A New Frontier in Antiviral Development. Viruses 2024, 16, 1423. https://doi.org/10.3390/v16091423

AMA Style

McGraw A, Hillmer G, Medehincu SM, Hikichi Y, Gagliardi S, Narayan K, Tibebe H, Marquez D, Mei Bose L, Keeting A, et al. Exploring HIV-1 Maturation: A New Frontier in Antiviral Development. Viruses. 2024; 16(9):1423. https://doi.org/10.3390/v16091423

Chicago/Turabian Style

McGraw, Aidan, Grace Hillmer, Stefania M. Medehincu, Yuta Hikichi, Sophia Gagliardi, Kedhar Narayan, Hasset Tibebe, Dacia Marquez, Lilia Mei Bose, Adleigh Keeting, and et al. 2024. "Exploring HIV-1 Maturation: A New Frontier in Antiviral Development" Viruses 16, no. 9: 1423. https://doi.org/10.3390/v16091423

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