Next Article in Journal
Hypermethylation of the VTRNA1-3 Promoter is Associated with Poor Outcome in Lower Risk Myelodysplastic Syndrome Patients
Next Article in Special Issue
Epigenetic Heterogeneity of B-Cell Lymphoma: Chromatin Modifiers
Previous Article in Journal / Special Issue
Epigenetic Therapy for Solid Tumors: Highlighting the Impact of Tumor Hypoxia
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Impact of Chromatin on HIV Replication

1
Department of Medicine, Boston University School of Medicine, Boston, MA 02118, USA
2
Graduate Program in Microbiology, Boston University School of Medicine, Boston, MA 02118, USA
*
Author to whom correspondence should be addressed.
Genes 2015, 6(4), 957-976; https://doi.org/10.3390/genes6040957
Submission received: 13 August 2015 / Revised: 14 September 2015 / Accepted: 22 September 2015 / Published: 30 September 2015
(This article belongs to the Special Issue Chromatin Dynamics)

Abstract

:
Chromatin influences Human Immunodeficiency Virus (HIV) integration and replication. This review highlights critical host factors that influence chromatin structure and organization and that also impact HIV integration, transcriptional regulation and latency. Furthermore, recent attempts to target chromatin associated factors to reduce the HIV proviral load are discussed.

1. Introduction

Retroviruses such as the Human Immunodeficiency Virus (HIV) are defined by their ability to integrate their reverse transcribed genome into the host DNA prior to proviral transcription by the cellular machinery. The role of chromatin in early HIV infection and replication has been extensively studied, revealing that chromatin influences HIV integration as well as the establishment and maintenance of a transcriptionally repressed but inducible reservoir of cells infected with HIV. These latently infected cells present a barrier to curing HIV infection since they are a persistent source of virus following the interruption of antiretroviral treatments. Recent eradication strategies have focused on developing novel compounds and testing new indications for existing drugs that target chromatin remodeling factors. This review will examine the effects of chromatin on the HIV life cycle and discuss the potential of emerging treatments that target chromatin-modifying factors.

2. Chromatin and HIV Integration

Retroviral integration is not a random event and it has been estimated that 70% of HIV provirus is integrated into introns of actively transcribed host genes [1,2,3,4,5,6,7]. This bias in provirus location suggests that local chromatin organization influences the site selection of integration. Integration requires a combination of viral factors including capsid, reverse transcriptase and integrase, and host factors that facilitate the HIV pre-integration complex’s translocation into the nucleus and targeting the viral genome to the host DNA [8,9,10]. Several cellular factors have been identified as being a part of the pre-integration complex or to directly interact with HIV integrase [10,11]. These cellular factors may facilitate nuclear translocation of the preintegration complex through the nuclear pores, or influence targeting of the HIV proviral DNA by interacting with chromatin factors. However, few of these proteins have been validated as playing a role in establishing HIV provirus.
The best studied cellular factor that participates in HIV integration is LEDGF/p75 [12,13,14,15]. LEDGF/p75 is encoded by the PSIP1 gene and has been described as a transcriptional activator and a splicing factor, although, exactly how this relatively ubiquitous transcription factor functions in regulating gene expression is not clear [13,16,17,18,19,20,21]. LEDGF/p75 is part of a larger family of transcription factors, the hepatoma-derived growth factor-related proteins, which are defined by conserved N-terminal PWWP and A/T hook domains [22,23]. A/T hook domains mediate LEDGF/p75 non-specific binding [24] whereas the PWWP motif recognizes specific chromatin marks such as acetylated and methylated lysines in histone tails [25]. LEDGF also contains a C-terminal integrase binding domain (IBD) [26,27,28], which directly interacts with a binding pocket formed by two HIV integrase catalytic core domains (CCD) [28,29]. LEDGF/p75 involvement in HIV integration was initially suspected based on its interaction with integrase [29,30,31] and knockdown experiments validated the role of LEDGF/p75 in HIV infection and integration [32,33,34,35,36]. These studies demonstrated that although depletion of LEDGF/p75 modestly reduced HIV infection, there was an altered integration pattern with significantly reduced HIV integration into actively transcribed genes and a shift towards promoters and CpG islands being targeted by the provirus [9,14]. Structure and function studies are consistent with a model that LEDGF/p75, through its interactions with chromatin, tethers the HIV integrase dimer to the targeted DNA [24,25,29,37,38,39,40]. LEDGF/p75 may also promote integration by protecting HIV integrase from degradation and by stimulating its catalytic activity [24,30,41,42,43]. Recently, factors that interact with LEDGF/p75, including TOX4, NOVA1 [44], JPO2 [41,42,45] and PogZ, have been implicated in HIV integration [46]. In addition to facilitating HIV integration, LEDGF/p75 may silence HIV transcription and contribute to the establishment and maintenance of HIV latency by forming a complex with Spt6 and Iws1 that promotes the recruitment of repressive chromatin to the HIV promoter [47]. Other chromatin interacting proteins may compliment or even be functionally redundant to LEDGF/p75. For example, hepatoma-derived growth factor-related protein 2 (HRP2), which shares homology with LEDGF/p75, binds HIV integrase and rescues LEDGF/p75 activity in knockout cells [31,48,49]. Furthermore, LEDGF/p75 is not a general requirement for retroviral integration. Gammaretroviruses, such as MLV, preferentially integrate into strong transcriptional cis-elements including promoters, enhancers, and DNase hypersensitive sites and this bias is in part mediated by Bromodomain and Extra Terminal (BET) family transcription factors [8,9,50,51]. This indicates that selection of retroviral integration is specific to different retroviruses and influenced by the cellular cofactors that are usurped by the retroviruses.
Ini-1/hSNF5, a SWI/SNF ATP dependent chromatin remodeling factor, has also been suggested to influence HIV replication. The actual function of Ini-1/SNF5 in the context of HIV replication is still unclear, which may reflect the fact that Ini-1/SNF5 has multiple activities at different steps in the HIV life cycle [52,53,54,55]. Ini-1/SNF5 was originally identified as an HIV integrase-interacting partner using a yeast two-hybrid system [56] and was shown to be recruited into HIV virions and pre-integration complexes through a Gag mediated interaction [54,57]. In vitro assays show that Ini-1/hSNF5 facilitates integration by ATP-dependent displacement of nucleosomes to increase DNA accessibility [58]. In addition, disrupting Ini-1/SNF5-integrase interactions has been reported to alter integration [59,60]. Ini-1/SNF5 has also been suggested to have antiviral activities, inhibiting early steps of HIV replication through its interaction with HIV integrase as well as components of the Sin3a-HDAC1 repressor complex [52,54,59]. The functional relevance of Ini-1/SNF5-HIV integrase interactions warrants further investigation.
Recently, the TRIM family protein TRIM28/KAP1 has been shown to physically interact with and inhibit HIV-1 integrase [61]. TRIM28/KAP1 is a transcriptional co-repressor that interacts with KRAB zinc finger DNA proteins through its bromodomain, binding acetylated lysines to mediate gene silencing by recruiting repressor complexes to target promoters [62]. HIV integrase activity is regulated by p300 which acetylates key lysines K264, K266 and K273 within HIV integrase and TRIM28/SNF5 binds these acetylated lysines. TRIM28/KAP1 inhibits integrase activity by recruiting HDAC1 which deacetylates the lysines [61]. Therefore, Trim28/KAP1 coordinates HIV-1 integrase function with chromatin remodeling during proviral integration [61].
Using fluorescence in situ hybridization, immunohistochemistry and microscopy, it has been shown that the nucleus is highly organized and that genes are not randomly distributed, but are partitioned into regions that cluster expressed genes from repressed genes [63,64]. For example, transcriptionally active euchromatin is detected near nucleopore complexes, whereas transcriptionally repressed heterochromatin is enriched in areas between nucleopore complexes [65,66]. This chromatin architecture underneath the nuclear envelope is maintained by nuclear pore complex proteins including Nup153 and Tpr [67]. If expressed genes are juxtaposed to nucleopore complexes then, upon entry into the nucleus, HIV will most likely encounter actively transcribed genes with open chromatin. Furthermore, disrupting nuclear pore complexes would be predicted to compromise nuclear entry and/or integration of the HIV pre-integration complex. The pre-integration complex, HIV capsid and HIV integrase have been shown to interact with several nuclear pore proteins including transportin 3 and nucleoporins NUP153 and NUP62 [68,69]. In addition, a siRNA screen showed that targeting proteins associated with nuclear pore complexes, such as RANBP2, altered the selection of provirus integration sites [70,71]. Other nuclear pore proteins that mediate HIV infection include Nup153 and Nup98, which facilitate nuclear import of the preintegration complex and Tpr, which influences integration site selection by maintaining chromatin architecture, anchoring Nup153 and functionally associating with LEDGF/p75 [67,72,73,74,75,76].
Two recent reports support the possibility that HIV provirus interacts with batteries of actively expressed genes upon translocation into the nucleus. Firstly, it has been observed that HIV genomes co-localize with PML [77]. It is also interesting to note that PML functionally interacts with Ini-1/SNF5 [78]. Secondly, examination of integration sites by three-dimensional immune DNA fluorescence in situ hybridization suggested that integrated virus was strongly associated with chromatin regions at the outer region of the nucleus whereas integration was disfavored in lamin-associated domains located more centrally in the nucleus [79]. This bias in integration sites was dependent on Nup153 and LEDGF/p75. Most importantly, these studies demonstrate the role DNA topography plays in biasing HIV integration site selection [79].

3. Chromatin and HIV Transcription

In addition to influencing sites of integration chromatin architecture directly affects the efficiency of HIV gene expression and silencing. HIV transcription is initiated at the transcriptional start site within the R region of the 5' long terminal repeat (LTR) [80,81,82]. Efficient transcription of proviral DNA is tightly linked to the activation status of the cell and the availability of cellular transcription factors [83,84,85,86]. These factors recognize specific sequences within the HIV LTR and subsequently recruit the RNA polymerase II (RNAP II) transcriptional machinery. Initial transcription of the HIV provirus leads to the production of the virally encoded transcription trans-activator Tat. Tat significantly enhances the efficiency of transcription by binding a RNA stem loop at the beginning of the newly transcribed HIV RNA at the trans-activation response element or TAR. Tat recruits transcription elongation factors to the HIV promoter to facilitate provirus transcription. One critical factor recruited by Tat includes the P-TEFb complex, which phosphorylates the C-terminal domain of RNAP II as well as factors that act as negative elongation factors such as DSIF and NELF, enabling efficient processivity and elongation of the RNA transcript [87,88,89,90]. However, not all of infected cells support productive viral replication. In T cells, a subset of infected cells carries latent proviruses which are not efficiently expressed [80,91,92]. This pool of latently infected resting T cells represent an important barrier to curing HIV infection as these cells can be re-activated to produce virus at any point after the interruption of HAART and lead to the rebound of the viral load in infected individuals [93,94,95]. The transcriptional repression observed in latently infected cells may result from multiple mechanisms including low levels of positive transcription factors, the presence of repressive transcription factors, and epigenetic changes associated with chromatin that limit HIV transcription.
The proviral chromatin architecture includes two nucleosomes, designated as nuc-0 and nuc-1, located immediately upstream of the modulatory region within U3 and downstream of the transcription initiation site, respectively. The location of these nucleosomes were originally mapped in DNAse hyper-sensitivity studies [96]. This organization showed that the upstream cis-elements of the HIV LTR which include Sp-1, NF-κB and Ap-1 sites, are accessible and nucleosome free, whereas Nuc-1, in particular, is in a position to impede RNAP II elongation. The recruitment of histone-modifying complexes targets Histone-3 (H3) and Histone-4 (H4) of Nuc-1 to regulate HIV transcription [97,98,99]. A well characterized modification is acetylation of histones. Acetylation of histones by acetyltransferases (HATs) generally leads to opening of chromatin allowing accessibility of the DNA to the transcription machinery or enhanced RNAP II processivity. Cellular transcription factors including NF-AT, Sp1, NF-κB, AP-1, as well as HIV Tat recruit histone acetyltransferases such as CBP/p300, P/CAF and GCN5 to the HIV LTR (Figure 1). Acetylation of H3 and H4 of nuc-1 has been implicated in the activation of HIV transcription [100,101,102,103]. ChIP-analysis on the acetylation pattern of the nucleosomes following phorbol-ester stimulation of latently infect T cells showed increase acetylation of H3 at lysines 9 and 14 and of Histone-4 at lysines 8 and 16 [101]. Conversely, repressive transcription factors such as NF-κB p50 homodimers, YY-1, LSF, NELF and CBF-1 recruit histone de-acetylases (HDAC), leading to a closed chromatin conformation, limited access to the transcription machinery and decreased RNAP II activity [104] (Figure 1). These modifications are closely associated to the activation state of the host cell and are important for the silencing of HIV transcription and the establishment of latency. HDAC-1 [104,105,106,107,108,109], HDAC-2 [108,110,111] and HDAC-3 [88,110,112] are the main factors that mediate the de-acetylation of the nuc-1 of HIV as evidenced by ChIP analysis of repressive protein complexes, siRNA knock-downs and pharmacological inhibition. In addition to promoting opening and closing of the chromatin structure, the acetylation status of histones provides a platform for the association of additional regulatory proteins. One such protein is BRD4, which recognizes acetylated histones and transcription factors through its bromodomains. BRD4 recruits and activates P-TEFb through its P-TEFb-interacting domain (PID) [113,114]. The interaction of these proteins mediates basal transcription of the provirus in the absence of Tat. Interestingly, BRD4 competes with Tat for P-TEFb, thus acting as an antagonist for efficient proviral transcription elongation. BRD4 further antagonizes Tat transactivation by phosphorylating the CDK9 subunit of P-TEFb, thus inhibiting kinase activity within the HIV transcription initiation complex [115]. Based on these observations BRD4 has been suggested to be a critical factor for limiting HIV transcription and has been suggested to be a target for reactivating latent HIV (see below).
Figure 1. General regulation of HIV transcription and chromatin remodeling. Cellular transcription factors (e.g., p50/p65 NF-κB heterodimers) recognize motifs within the HIV provirus promoter region and recruit cellular co-activators, chromatin remodeling factors and transcription elongation factors. Negative transcription factors (e.g., p50/p50 NF-κB homodimers) recruit cellular co-repressors, chromatin remodeling factors and factors that hinder transcription elongation.
Figure 1. General regulation of HIV transcription and chromatin remodeling. Cellular transcription factors (e.g., p50/p65 NF-κB heterodimers) recognize motifs within the HIV provirus promoter region and recruit cellular co-activators, chromatin remodeling factors and transcription elongation factors. Negative transcription factors (e.g., p50/p50 NF-κB homodimers) recruit cellular co-repressors, chromatin remodeling factors and factors that hinder transcription elongation.
Genes 06 00957 g001
Histones are also modified by methylation and this modification, depending on the context and residue, either contributes to repression or to activation of transcription. Sp1 in complex with CTIP2 recruits SUV39H1 and mediates tri-methylation of H3 at lysine 9 [108]. The methyltransferase G9a also mediates repression of HIV transcription through dimethylation of H3 [116]. It remains unclear which factors mediate activation-related methylation, but it is possible that trimethylation of H3 at lysine 4 and dimethylation at lysine 36 mediated by SET1 and Smyd2 can promote proviral transcription [103,117,118,119]. Methyltransferases have also been shown to modify proviral DNA and HIV Tat creating additional layers of transcriptional regulation [120,121,122,123].
Another important modification to the chromatin structure is mediated by the SWI/SNF complexes. This is a large complex composed of 12 individual proteins [124]. The main function of SWI/SNF is to facilitate the mobilization of nucleosomes along the DNA in an ATP-dependent manner [124], thus facilitating access of DNA to the transcription machinery. In the context of HIV replication, SWI/SNF proteins influence integration (see above) and activate HIV transcription. Upon activation of the cell, Tat becomes acetylated. This modification of Tat allows it to bind to the catalytic subunit of the SWI/SNF complex known as Brg-1, thus recruiting the complex to the HIV promoter. The recruitment of SWI/SNF to the HIV promoter may also be facilitated by the interaction of Brg-1 with acetylated nuc-1 upon phorbol ester stimulation [125].
Chromatin re-assembly factors have also been proposed as limiting HIV transcription. Knockdown experiments of Spt6, Spt16 and Chd1, members of the facilitates chromatin transcription (FACT) complex, induced HIV transcription in latently infected cell lines [126]. These results suggest that the FACT complex promotes silencing of HIV transcription by assembling nucleosomes along the proviral DNA. Further studies are required to confirm the direct recruitment of these factors to the provirus and how their activity is regulated in the context of HIV latency.

4. Targeting Chromatin as a Therapeutic Strategy for HIV

Given the importance of DNA architecture for the efficient integration and transcription of HIV, chromatin modifying factors have been appealing targets for anti-retroviral therapies (Table 1). To block early steps in the HIV life cycle, drugs modulating the interaction between LEDGF/p75 and HIV integrase (IN), called LEDGINs, are under development. Christ et al. identified 2-(quinolin-3-yl) acetic acid compounds that inhibited binding between HIV integrase and LEDGF/p75 [127]. Furthermore, these LEDGINs stabilize IN multimers, allosterically inhibit integrase activity and prevent proper virion assembly [127,128,129,130]. Although the initial LEDGINs displayed IC50 values in the micromolar range, tert-butoxy-(4-phenyl-quinolin-3-yl) acetic acid derivatives are potent in the nanomolar range, increasing the likelihood that they will be brought to clinical trials [128].
Chromatin-targeting strategies for treating HIV also include compounds designed to work at the level of transcription such as “shock and kill” strategies to reactivate latent proviruses in conjunction with antiretroviral therapy [131]. With virus replication blocked by routine antiretroviral therapy, it is expected that the reactivated latent population will be cleared by the host immune system or by the cytopathic effects of viral replication. Many of these latency-reversing agents (LRAs) target regulators of chromatin.
Table 1. Potential HIV therapeutics that target chromatin.
Table 1. Potential HIV therapeutics that target chromatin.
Drug ClassCompoundsReferences
Inhibitors of Integration
LEDGINs2-(quinolin-3-yl) acetic acids, tert-butoxy-(4-phenyl-quinolin-3-yl) acetic acids[127,128,129,130]
Transcriptional Activators
Histone Deacetylase Inhibitors (HDACi)Valproic acid, SAHA, (vorinostat), givinostat, panobinostat, entinostat, romidepsin[132,133,134,135,136,137,138,139,140,141,142,143,144,145,146,147,148,149]
BET Bromodo main InhibitorsJQ1[150,151,152,153,154]
Histone Methyltransferase Inhibitors (HMTi)Chaeotocin, BIX-01294, DZNep[155,156,157]
DNA Methyltransferase Inhibitors (DMTi)5-aza-2'-deoxycytidine[121,122,157]
The most widely studied LRAs are HDAC inhibitors, which maintain an open chromatin structure by histone acetylation. The epilepsy drug, valproic acid (VPA), is a Class I/II HDAC inhibitor [132]. Phase I and II clinical trials for using VPA as a latency-reversing agent have so far shown mixed results regarding changes in the amount of integrated virus as measured by PCR. In the initial proof-of-concept study, three out of four patients obtained a mean reduction of 75% infected units per billion (IUPB) as determined by a viral outgrowth assay (VOA) [133]. A follow-up trial was only able to replicate a statistically significant decrease in IUPB for four out of 11 subjects [134]. In addition, subsequent studies indicated that valproic acid did not alter the amount of integrated latent HIV [135,136,137] and may increase the cognitive impairments associated with AIDS [138].
SAHA or vorinostat, a treatment for T cell lymphomas, is another promising HDAC inhibitor. It induces HIV expression in cellular models of HIV latency, such as J-LAT, J∆K, ACH-2, and U1 cells, as well as ex vivo experiments with patient peripheral blood mononuclear cells [139,140,141]. In a small clinical trial, SAHA treatment substantially increased cell-associated viral RNA levels in patients on ART, several of whom had no detectable HIV RNA before study enrollment [142]. However, a second clinical trial demonstrated that, although oral doses of vorinostat led to a significant increase in cell-associated unspliced HIV RNA for the majority of subjects, there was no change in plasma HIV RNA or the amount of integrated HIV DNA [143].
A large number of other HDAC inhibitors are under investigation, including givinostat, panobinostat, entinostat, and romidepsin [144,145,146,147,148]. Rasmussen et al. [145,149] contrasted the effects of clinically relevant concentrations of panobinostat, givinostat, vorinostat, and valproic acid on U1 and ACH2 latent cell line models as well as latently-infected resting primary cells. Panobinostat, a hydroxamic-acid pan-HDAC inhibitor, significantly increased HIV transcription as compared to the other inhibitors for all cell types [145]. A clinical trial with panobinostat demonstrated its capacity to upregulate cell-associated unspliced HIV RNA and temporarily perturb the amount of viral DNA in patients on ART [149].
Besides HDAC inhibitors, several groups have shown that the BET bromodomain inhibitor JQ1 can reactivate HIV in U1, ACH2, J∆K, and J-Lat cell models [150,151,152,153,154]. In addition, JQ1 induced HIV expression ex vivo in T cells from one out of three individuals on ART [150]. Evidence suggests that JQ1 may inhibit the BET protein BRD4, which represses transcription by sequestering PTEF-b from Tat [113,151]. However, recent studies suggest that JQ1 may function by inhibiting BRD2. Although the mechanism by which BRD2 contributes to HIV latency has not been elucidated, Boehm et al. postulated that BRD2 may target HDACs to the proviral LTR [154]. JQ1 has not yet been brought to clinical trials for the treatment of HIV.
Histone methyltransferases (HMTs) present another potential target for LRAs. Chaeotocin, a fungal toxin and SUV39H1 HMT inhibitor (HMTi), reactivates HIV in latently infected cell models [155,156]. Furthermore, chaeotocin and BIX-01294, a G9a HMTi, can stimulate HIV transcription ex vivo in PBMCs from ART-suppressed HIV+ patients. In contrast, other evidence indicates that DZNep, which inhibits several HMTs including EZH2, is a more potent inducer of HIV transcription than either chaeotocin or BIX-01294 in latently infected Jurkats [157]. These HMTis, however, are too toxic to be used as single agents in clinical trials. DNA methyltransferase inhibitors, while also appealing, have not yet demonstrated the same potential to induce latent HIV expression as HMTi’s. 5-aza-2'-deoxycytidine, which inhibits the cytosine methylation associated with CpG islands, weakly reactivates HIV but can synergize with other LRAs such as the protein kinase C modulator prostratin [121,122,157].
In spite of promising results obtained with cellular models of latency and ex vivo activation of CD4+ T cells from HIV patients, LRAs have had little success in clinical settings. Other than early trials of valproic acid in combination with ART [133], none have successfully reduced the viral reservoir in vivo. However, the potential for global T cell activation, specifically for HDAC inhibitors, may have had significant negative repercussions within these trials and would have limited the dosages, perhaps below the therapeutic threshold. This has led to the search for non-T cell stimulating agents that reactivate latent HIV, like protein kinase C activators. These compounds, which use mechanisms of action unrelated to chromatin dynamics, may have increased efficacy after other LRAs have reorganized the DNA architecture. Indeed, several studies have identified that combinatorial approaches maximally induce HIV expression in latently infected primary cells and cell models [43,140,152,153,157,158,159]. For example, it was demonstrated that several minimally effective small molecules synergize with known activators such as prostratin to induce HIV expression in J-Lat cells [160].
Multifaceted strategies to induce HIV expression will be essential to significantly perturb the HIV reservoir. New evidence indicates that we may be substantially underestimating the size of the viral reservoir. With VOA-based measurements, about one in 106 resting CD4+ T cells are estimated to be inducible [161,162]. In contrast, genomic PCR indicates that nearly 200 in 106 resting CD4+ T cells contain integrated provirus [163] and this proviral load may be dynamic and vary between patients. It has been assumed that the majority of these integrated proviruses are replication-incompetent. However, the Siliciano lab sequenced 213 proviruses from eight patients on ART that were incapable of being activated by the standard LRA phytohemagglutinin and showed that approximately 10% of proviruses deemed non-inducible by VOA were actually capable of infection and replication [164]. Therefore, with this uncertainty surrounding the size and location of the latent reservoir, the potential of LRAs may need to be further evaluated.
Future design of HIV/AIDS therapeutics must incorporate drugs designed to work at the level of chromatin. Host and proviral genomic architecture have profound repercussions for the fate of infected cells. Specific manipulation of chromatin would allow us to modify the balance between the maintenance of latently infected cells and cells containing transcriptionally active proviruses that can then be targeted and finally cleared.

5. Conclusions

Chromatin influences HIV integration and transcription which has implications for efficient HIV replication and latency. Therapeutically targeting chromatin and associated factors provides promising strategies for preventing HIV infection and reducing the size of the latent reservoir.

Acknowledgments

This work was in part supported by NIH grants A1AI097117, DE 023950 and AI108447 awarded to Andrew J. Henderson. Luis M. Agosto and Matthew Gagne were supported by the Immunology Training Grant T32-AI007309.

Author Contributions

All authors contributed equally to scholarship and writing of the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Lewinski, M.K.; Bisgrove, D.; Shinn, P.; Chen, H.; Hoffmann, C.; Hannenhalli, S.; Verdin, E.; Berry, C.C.; Ecker, J.R.; Bushman, F.D. Genome-wide analysis of chromosomal features repressing human immunodeficiency virus transcription. J. Virol. 2005, 79, 6610–6619. [Google Scholar] [CrossRef] [PubMed]
  2. Shan, L.; Yang, H.C.; Rabi, S.A.; Bravo, H.C.; Shroff, N.S.; Irizarry, R.A.; Zhang, H.; Margolick, J.B.; Siliciano, J.D.; Siliciano, R.F. Influence of host gene transcription level and orientation on HIV-1 latency in a primary-cell model. J. Virol. 2011, 85, 5384–5393. [Google Scholar] [CrossRef] [PubMed]
  3. Sherrill-Mix, S.; Lewinski, M.K.; Famiglietti, M.; Bosque, A.; Malani, N.; Ocwieja, K.E.; Berry, C.C.; Looney, D.; Shan, L.; Agosto, L.M.; et al. HIV latency and integration site placement in five cell-based models. Retrovirology 2013. [Google Scholar] [CrossRef]
  4. Lenasi, T.; Contreras, X.; Peterlin, B.M. Transcriptional interference antagonizes proviral gene expression to promote HIV latency. Cell Host Microbe 2008, 4, 123–133. [Google Scholar] [CrossRef] [PubMed]
  5. Ding, D.; Qu, X.; Li, L.; Zhou, X.; Liu, S.; Lin, S.; Wang, P.; Kong, C.; Wang, X.; Liu, L.; et al. Involvement of histone methyltransferase GLP in HIV-1 latency through catalysis of H3K9 dimethylation. Virology 2013, 440, 182–189. [Google Scholar] [CrossRef] [PubMed]
  6. Han, Y.; Lassen, K.; Monie, D.; Sedaghat, A.R.; Shimoji, S.; Liu, X.; Pierson, T.C.; Margolick, J.B.; Siliciano, R.F.; Siliciano, J.D. Resting CD4+ T cells from HIV-1 infected individuals carry integrated HIV-1 genomes within actively transcribed host genes. J. Virol. 2004, 78, 6122–6133. [Google Scholar] [CrossRef] [PubMed]
  7. Rezaei, S.D.; Cameron, P.U. Human immunodeficiency virus (HIV)-1 integration sites in viral latency. Curr. HIV/AIDS Rep. 2015, 12, 88–96. [Google Scholar] [CrossRef] [PubMed]
  8. Krishnan, L.; Engelman, A. Retroviral integrase proteins and HIV-1 DNA integration. J. Biol. Chem. 2012, 287, 40858–40866. [Google Scholar] [CrossRef] [PubMed]
  9. Craigie, R.; Bushman, F.D. HIV DNA integration. Cold Spring Harb. Perspect. Med. 2012. [Google Scholar] [CrossRef] [PubMed]
  10. Van Maele, B.; Busschots, K.; Vandekerckhove, L.; Christ, F.; Debyser, Z. Cellular co-factors of HIV-1 integration. Trends Biochem. Sci. 2006, 31, 98–105. [Google Scholar] [CrossRef] [PubMed]
  11. Vandegraaff, N.; Engelman, A. Molecular mechanisms of HIV integration and therapeutic intervention. Expert Rev. Mol. Med. 2007, 9, 1–19. [Google Scholar] [CrossRef] [PubMed]
  12. Ciuffi, A.; Bushman, F.D. Retroviral DNA integration: HIV and the role of LEDGF/p75. Trends Genet. 2006, 22, 388–395. [Google Scholar] [CrossRef] [PubMed]
  13. Llano, M.; Morrison, J.; Poeschla, E.M. Virological and cellular roles of the transcriptional coactivator LEDGF/p75. Curr. Top. Microbiol. Immunol. 2009, 339, 125–146. [Google Scholar] [PubMed]
  14. Engelman, A.; Cherepanov, P. The lentiviral integrase binding protein LEDGF/p75 and HIV-1 replication. PLoS Pathog. 2008, 4, e1000046. [Google Scholar] [CrossRef] [PubMed]
  15. Christ, F.; Debyser, Z. The LEDGF/p75 integrase interaction, a novel target for anti-HIV therapy. Virology 2013, 435, 102–109. [Google Scholar] [CrossRef] [PubMed]
  16. Ge, H.; Si, Y.; Roeder, R.G. Isolation of cDNAs encoding novel transcription coactivators p52 and p75 reveals an alternate regulatory mechanism of transcriptional activation. EMBO J. 1998, 17, 6723–6729. [Google Scholar] [CrossRef] [PubMed]
  17. Shinohara, T.; Singh, D.P.; Fatma, N. LEDGF, a survival factor, activates stress-related genes. Prog. Retin. Eye Res. 2002, 21, 341–358. [Google Scholar] [CrossRef]
  18. Sutherland, H.G.; Newton, K.; Brownstein, D.G.; Holmes, M.C.; Kress, C.; Semple, C.A.; Bickmore, W.A. Disruption of LEDGF/PSIP1 results in perinatal mortality and homeotic skeletal transformations. Mol. Cell. Biol. 2006, 26, 7201–7210. [Google Scholar] [CrossRef] [PubMed]
  19. Pradeepa, M.M.; Sutherland, H.G.; Ule, J.; Grimes, G.R.; Bickmore, W.A. PSIP1/LEDGF p52 binds methylated histone H3K36 and splicing factors and contributes to the regulation of alternative splicing. PLoS Genet. 2012, 8, e1002717. [Google Scholar] [CrossRef] [PubMed]
  20. Singh, D.P.; Fatma, N.; Kimura, A.; Chylack, L.T., Jr.; Shinohara, T. LEDGF binds to heat shock and stress-related element to activate the expression of stress-related genes. Biochem. Biophys. Res. Commun. 2001, 283, 943–955. [Google Scholar] [CrossRef] [PubMed]
  21. Singh, D.P.; Ohguro, N.; Kikuchi, T.; Sueno, T.; Reddy, V.N.; Yuge, K.; Chylack, L.T., Jr.; Shinohara, T. Lens epithelium-derived growth factor: Effects on growth and survival of lens epithelial cells, keratinocytes, and fibroblasts. Biochem. Biophys. Res. Commun. 2000, 267, 373–381. [Google Scholar] [CrossRef] [PubMed]
  22. Dietz, F.; Franken, S.; Yoshida, K.; Nakamura, H.; Kappler, J.; Gieselmann, V. The family of hepatoma-derived growth factor proteins: Characterization of a new member HRP-4 and classification of its subfamilies. Biochem. J. 2002, 366, 491–500. [Google Scholar] [CrossRef] [PubMed]
  23. Nameki, N.; Tochio, N.; Koshiba, S.; Inoue, M.; Yabuki, T.; Aoki, M.; Seki, E.; Matsuda, T.; Fujikura, Y.; Saito, M.; et al. Solution structure of the PWWP domain of the hepatoma-derived growth factor family. Protein Sci. 2005, 14, 756–764. [Google Scholar] [CrossRef] [PubMed]
  24. Turlure, F.; Maertens, G.; Rahman, S.; Cherepanov, P.; Engelman, A. A tripartite DNA-binding element, comprised of the nuclear localization signal and two AT-hook motifs, mediates the association of LEDGF/p75 with chromatin in vivo. Nucleic Acids Res. 2006, 34, 1653–1665. [Google Scholar] [CrossRef] [PubMed]
  25. Gijsbers, R.; Vets, S.; de Rijck, J.; Ocwieja, K.E.; Ronen, K.; Malani, N.; Bushman, F.D.; Debyser, Z. Role of the PWWP domain of lens epithelium-derived growth factor (LEDGF)/p75 cofactor in lentiviral integration targeting. J. Biol. Chem. 2011, 286, 41812–41825. [Google Scholar] [CrossRef] [PubMed]
  26. Vanegas, M.; Llano, M.; Delgado, S.; Thompson, D.; Peretz, M.; Poeschla, E. Identification of the LEDGF/p75 HIV-1 integrase-interaction domain and NLS reveals NLS-independent chromatin tethering. J. Cell Sci. 2005, 118, 1733–1743. [Google Scholar] [CrossRef] [PubMed]
  27. De Rijck, J.; Vandekerckhove, L.; Gijsbers, R.; Hombrouck, A.; Hendrix, J.; Vercammen, J.; Engelborghs, Y.; Christ, F.; Debyser, Z. Overexpression of the lens epithelium-derived growth factor/p75 integrase binding domain inhibits human immunodeficiency virus replication. J. Virol. 2006, 80, 11498–11509. [Google Scholar] [CrossRef] [PubMed]
  28. Cherepanov, P.; Ambrosio, A.L.; Rahman, S.; Ellenberger, T.; Engelman, A. Structural basis for the recognition between HIV-1 integrase and transcriptional coactivator p75. Proc. Natl. Acad. Sci. USA 2005, 102, 17308–17313. [Google Scholar] [CrossRef] [PubMed]
  29. Cherepanov, P.; Sun, Z.Y.; Rahman, S.; Maertens, G.; Wagner, G.; Engelman, A. Solution structure of the HIV-1 integrase-binding domain in LEDGF/p75. Nat. Struct. Mol. Biol. 2005, 12, 526–532. [Google Scholar] [CrossRef] [PubMed]
  30. 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]
  31. Cherepanov, P.; Devroe, E.; Silver, P.A.; Engelman, A. Identification of an evolutionarily conserved domain in human lens epithelium-derived growth factor/transcriptional co-activator p75 (LEDGF/p75) that binds HIV-1 integrase. J. Biol. Chem. 2004, 279, 48883–48892. [Google Scholar] [CrossRef] [PubMed]
  32. Llano, M.; Saenz, D.T.; Meehan, A.; Wongthida, P.; Peretz, M.; Walker, W.H.; Teo, W.; Poeschla, E.M. An essential role for LEDGF/p75 in HIV integration. Science 2006, 314, 461–464. [Google Scholar] [CrossRef] [PubMed]
  33. 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]
  34. Fadel, H.J.; Morrison, J.H.; Saenz, D.T.; Fuchs, J.R.; Kvaratskhelia, M.; Ekker, S.C.; Poeschla, E.M. TALEN knockout of the PSIP1 gene in human cells: Analyses of HIV-1 replication and allosteric integrase inhibitor mechanism. J. Virol. 2014, 88, 9704–9717. [Google Scholar] [CrossRef] [PubMed]
  35. Badia, R.; Pauls, E.; Riveira-Munoz, E.; Clotet, B.; Este, J.A.; Ballana, E. Zinc finger endonuclease targeting PSIP1 inhibits HIV-1 integration. Antimicrob. Agents Chemother. 2014, 58, 4318–4327. [Google Scholar] [CrossRef] [PubMed]
  36. Shun, M.C.; Raghavendra, N.K.; Vandegraaff, N.; Daigle, J.E.; Hughes, S.; Kellam, P.; Cherepanov, P.; Engelman, A. LEDGF/p75 functions downstream from preintegration complex formation to effect gene-specific HIV-1 integration. Genes Dev. 2007, 21, 1767–1778. [Google Scholar] [CrossRef] [PubMed]
  37. Hare, S.; Shun, M.C.; Gupta, S.S.; Valkov, E.; Engelman, A.; Cherepanov, P. A novel co-crystal structure affords the design of gain-of-function lentiviral integrase mutants in the presence of modified PSIP1/LEDGF/p75. PLoS Pathog. 2009, 5, e1000259. [Google Scholar] [CrossRef] [PubMed]
  38. Botbol, Y.; Raghavendra, N.K.; Rahman, S.; Engelman, A.; Lavigne, M. Chromatinized templates reveal the requirement for the LEDGF/p75 PWWP domain during HIV-1 integration in vitro. Nucleic Acids Res. 2008, 36, 1237–1246. [Google Scholar] [CrossRef]
  39. Rahman, S.; Lu, R.; Vandegraaff, N.; Cherepanov, P.; Engelman, A. Structure-based mutagenesis of the integrase-LEDGF/p75 interface uncouples a strict correlation between in vitro protein binding and HIV-1 fitness. Virology 2007, 357, 79–90. [Google Scholar] [CrossRef] [PubMed]
  40. 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] [PubMed]
  41. Hendrix, J.; van Heertum, B.; Vanstreels, E.; Daelemans, D.; de Rijck, J. Dynamics of the ternary complex formed by c-Myc interactor JPO2, transcriptional co-activator LEDGF/p75, and chromatin. J. Biol. Chem. 2014, 289, 12494–12506. [Google Scholar] [CrossRef] [PubMed]
  42. Maertens, G.N.; Cherepanov, P.; Engelman, A. Transcriptional co-activator p75 binds and tethers the Myc-interacting protein JPO2 to chromatin. J. Cell Sci. 2006, 119, 2563–2571. [Google Scholar] [CrossRef] [PubMed]
  43. Llano, M.; Delgado, S.; Vanegas, M.; Poeschla, E.M. Lens epithelium-derived growth factor/p75 prevents proteasomal degradation of HIV-1 integrase. J. Biol. Chem. 2004, 279, 55570–55577. [Google Scholar] [CrossRef] [PubMed]
  44. Morchikh, M.; Naughtin, M.; di Nunzio, F.; Xavier, J.; Charneau, P.; Jacob, Y.; Lavigne, M. TOX4 and NOVA1 proteins are partners of the LEDGF PWWP domain and affect HIV-1 replication. PLoS ONE 2013, 8, e81217. [Google Scholar] [CrossRef] [PubMed]
  45. Bartholomeeusen, K.; de Rijck, J.; Busschots, K.; Desender, L.; Gijsbers, R.; Emiliani, S.; Benarous, R.; Debyser, Z.; Christ, F. Differential interaction of HIV-1 integrase and JPO2 with the C terminus of LEDGF/p75. J. Mol. Biol. 2007, 372, 407–421. [Google Scholar] [CrossRef] [PubMed]
  46. Bartholomeeusen, K.; Christ, F.; Hendrix, J.; Rain, J.C.; Emiliani, S.; Benarous, R.; Debyser, Z.; Gijsbers, R.; de Rijck, J. Lens epithelium-derived growth factor/p75 interacts with the transposase-derived DDE domain of PogZ. J. Biol. Chem. 2009, 284, 11467–11477. [Google Scholar] [CrossRef] [PubMed]
  47. Gerard, A.; Segeral, E.; Naughtin, M.; Abdouni, A.; Charmeteau, B.; Cheynier, R.; Rain, J.C.; Emiliani, S. The integrase cofactor LEDGF/p75 associates with Iws1 and Spt6 for postintegration silencing of HIV-1 gene expression in latently infected cells. Cell Host Microbe 2015, 17, 107–117. [Google Scholar] [CrossRef] [PubMed]
  48. Wang, H.; Jurado, K.A.; Wu, X.; Shun, M.C.; Li, X.; Ferris, A.L.; Smith, S.J.; Patel, P.A.; Fuchs, J.R.; Cherepanov, P.; et al. HRP2 determines the efficiency and specificity of HIV-1 integration in LEDGF/p75 knockout cells but does not contribute to the antiviral activity of a potent LEDGF/p75-binding site integrase inhibitor. Nucleic Acids Res. 2012, 40, 11518–11530. [Google Scholar] [CrossRef] [PubMed]
  49. Vandegraaff, N.; Devroe, E.; Turlure, F.; Silver, P.A.; Engelman, A. Biochemical and genetic analyses of integrase-interacting proteins lens epithelium-derived growth factor (LEDGF)/p75 and hepatoma-derived growth factor related protein 2 (HRP2) in preintegration complex function and HIV-1 replication. Virology 2006, 346, 415–426. [Google Scholar] [CrossRef] [PubMed]
  50. Gupta, S.S.; Maetzig, T.; Maertens, G.N.; Sharif, A.; Rothe, M.; Weidner-Glunde, M.; Galla, M.; Schambach, A.; Cherepanov, P.; Schulz, T.F. Bromo- and extraterminal domain chromatin regulators serve as cofactors for murine leukemia virus integration. J. Virol. 2013, 87, 12721–12736. [Google Scholar] [CrossRef] [PubMed]
  51. Sharma, A.; Larue, R.C.; Plumb, M.R.; Malani, N.; Male, F.; Slaughter, A.; Kessl, J.J.; Shkriabai, N.; Coward, E.; Aiyer, S.S.; et al. BET proteins promote efficient murine leukemia virus integration at transcription start sites. Proc. Natl. Acad. Sci. USA 2013, 110, 12036–12041. [Google Scholar] [CrossRef]
  52. Sorin, M.; Yung, E.; Wu, X.; Kalpana, G.V. HIV-1 replication in cell lines harboring INI1/hSNF5 mutations. Retrovirology 2006. [Google Scholar] [CrossRef] [PubMed]
  53. Ariumi, Y.; Serhan, F.; Turelli, P.; Telenti, A.; Trono, D. The integrase interactor 1 (INI1) proteins facilitate Tat-mediated human immunodeficiency virus type 1 transcription. Retrovirology 2006. [Google Scholar] [CrossRef] [PubMed]
  54. Cano, J.; Kalpana, G.V. Inhibition of early stages of HIV-1 assembly by INI1/hSNF5 transdominant negative mutant S6. J. Virol. 2011, 85, 2254–2265. [Google Scholar] [CrossRef]
  55. Boese, A.; Sommer, P.; Holzer, D.; Maier, R.; Nehrbass, U. Integrase interactor 1 (Ini1/hSNF5) is a repressor of basal human immunodeficiency virus type 1 promoter activity. J. Gen. Virol. 2009, 90, 2503–2512. [Google Scholar] [CrossRef] [PubMed]
  56. Kalpana, G.V.; Marmon, S.; Wang, W.; Crabtree, G.R.; Goff, S.P. Binding and stimulation of HIV-1 integrase by a human homolog of yeast transcription factor SNF5. Science 1994, 266, 2002–2006. [Google Scholar] [CrossRef] [PubMed]
  57. Yung, E.; Sorin, M.; Pal, A.; Craig, E.; Morozov, A.; Delattre, O.; Kappes, J.; Ott, D.; Kalpana, G.V. Inhibition of HIV-1 virion production by a transdominant mutant of integrase interactor 1. Nat. Med. 2001, 7, 920–926. [Google Scholar] [CrossRef] [PubMed]
  58. Lesbats, P.; Botbol, Y.; Chevereau, G.; Vaillant, C.; Calmels, C.; Arneodo, A.; Andreola, M.L.; Lavigne, M.; Parissi, V. Functional coupling between HIV-1 integrase and the SWI/SNF chromatin remodeling complex for efficient in vitro integration into stable nucleosomes. PLoS Pathog. 2011, 7, e1001280. [Google Scholar] [CrossRef] [PubMed]
  59. Mathew, S.; Nguyen, M.; Wu, X.; Pal, A.; Shah, V.B.; Prasad, V.R.; Aiken, C.; Kalpana, G.V. INI1/hSNF5-interaction defective HIV-1 IN mutants exhibit impaired particle morphology, reverse transcription and integration in vivo. Retrovirology 2013. [Google Scholar] [CrossRef] [PubMed]
  60. Maillot, B.; Levy, N.; Eiler, S.; Crucifix, C.; Granger, F.; Richert, L.; Didier, P.; Godet, J.; Pradeau-Aubreton, K.; Emiliani, S.; et al. Structural and functional role of INI1 and LEDGF in the HIV-1 preintegration complex. PLoS ONE 2013, 8, e60734. [Google Scholar] [CrossRef] [PubMed]
  61. Allouch, A.; di Primio, C.; Alpi, E.; Lusic, M.; Arosio, D.; Giacca, M.; Cereseto, A. The TRIM family protein KAP1 inhibits HIV-1 integration. Cell Host Microbe 2011, 9, 484–495. [Google Scholar] [CrossRef]
  62. Iyengar, S.; Farnham, P.J. KAP1 protein: An enigmatic master regulator of the genome. J. Biol. Chem. 2011, 286, 26267–26276. [Google Scholar] [CrossRef] [PubMed]
  63. Burns, L.T.; Wente, S.R. From hypothesis to mechanism: Uncovering nuclear pore complex links to gene expression. Mol. Cell. Biol. 2014, 34, 2114–2120. [Google Scholar] [CrossRef] [PubMed]
  64. Ibarra, A.; Hetzer, M.W. Nuclear pore proteins and the control of genome functions. Genes Dev. 2015, 29, 337–349. [Google Scholar] [CrossRef] [PubMed]
  65. Torok, D.; Ching, R.W.; Bazett-Jones, D.P. PML nuclear bodies as sites of epigenetic regulation. Front. Biosci. 2009, 14, 1325–1336. [Google Scholar] [CrossRef]
  66. Krull, S.; Dorries, J.; Boysen, B.; Reidenbach, S.; Magnius, L.; Norder, H.; Thyberg, J.; Cordes, V.C. Protein Tpr is required for establishing nuclear pore-associated zones of heterochromatin exclusion. EMBO J. 2010, 29, 1659–1673. [Google Scholar] [CrossRef] [PubMed]
  67. Krull, S.; Thyberg, J.; Bjorkroth, B.; Rackwitz, H.R.; Cordes, V.C. Nucleoporins as components of the nuclear pore complex core structure and Tpr as the architectural element of the nuclear basket. Mol. Biol. Cell 2004, 15, 4261–4277. [Google Scholar] [CrossRef] [PubMed]
  68. Krishnan, L.; Matreyek, K.A.; Oztop, I.; Lee, K.; Tipper, C.H.; Li, X.; Dar, M.J.; Kewalramani, V.N.; Engelman, A. The requirement for cellular transportin 3 (TNPO3 or TRN-SR2) during infection maps to human immunodeficiency virus type 1 capsid and not integrase. J. Virol. 2010, 84, 397–406. [Google Scholar] [CrossRef] [PubMed]
  69. Ao, Z.; Jayappa, K.D.; Wang, B.; Zheng, Y.; Wang, X.; Peng, J.; Yao, X. Contribution of host nucleoporin 62 in HIV-1 integrase chromatin association and viral DNA integration. J. Biol. Chem. 2012, 287, 10544–10555. [Google Scholar] [CrossRef] [PubMed]
  70. Ocwieja, K.E.; Brady, T.L.; Ronen, K.; Huegel, A.; Roth, S.L.; Schaller, T.; James, L.C.; Towers, G.J.; Young, J.A.; Chanda, S.K.; et al. HIV integration targeting: A pathway involving Transportin-3 and the nuclear pore protein RanBP2. PLoS Pathog. 2011, 7, e1001313. [Google Scholar] [CrossRef] [PubMed]
  71. 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] [PubMed]
  72. 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] [PubMed]
  73. 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] [PubMed]
  74. Di Nunzio, F.; Fricke, T.; Miccio, A.; Valle-Casuso, J.C.; Perez, P.; Souque, P.; Rizzi, E.; Severgnini, M.; Mavilio, F.; Charneau, P.; et al. Nup153 and Nup98 bind the HIV-1 core and contribute to the early steps of HIV-1 replication. Virology 2013, 440, 8–18. [Google Scholar] [CrossRef] [PubMed]
  75. Lelek, M.; Casartelli, N.; Pellin, D.; Rizzi, E.; Souque, P.; Severgnini, M.; di Serio, C.; Fricke, T.; Diaz-Griffero, F.; Zimmer, C.; et al. Chromatin organization at the nuclear pore favours HIV replication. Nat. Commun. 2015. [Google Scholar] [CrossRef] [PubMed]
  76. Ebina, H.; Aoki, J.; Hatta, S.; Yoshida, T.; Koyanagi, Y. Role of Nup98 in nuclear entry of human immunodeficiency virus type 1 cDNA. Microbes Infect. 2004, 6, 715–724. [Google Scholar] [CrossRef] [PubMed]
  77. Lusic, M.; Marini, B.; Ali, H.; Lucic, B.; Luzzati, R.; Giacca, M. Proximity to PML nuclear bodies regulates HIV-1 latency in CD4+ T cells. Cell Host Microbe 2013, 13, 665–677. [Google Scholar] [CrossRef] [PubMed]
  78. Turelli, P.; Doucas, V.; Craig, E.; Mangeat, B.; Klages, N.; Evans, R.; Kalpana, G.; Trono, D. Cytoplasmic recruitment of INI1 and PML on incoming HIV preintegration complexes: Interference with early steps of viral replication. Mol. Cell 2001, 7, 1245–1254. [Google Scholar] [CrossRef]
  79. Marini, B.; Kertesz-Farkas, A.; Ali, H.; Lucic, B.; Lisek, K.; Manganaro, L.; Pongor, S.; Luzzati, R.; Recchia, A.; Mavilio, F.; et al. Nuclear architecture dictates HIV-1 integration site selection. Nature 2015, 521, 227–231. [Google Scholar] [CrossRef] [PubMed]
  80. Karn, J.; Stoltzfus, C.M. Transcriptional and posttranscriptional regulation of HIV-1 gene expression. Cold Spring Harb. Perspect. Med. 2012. [Google Scholar] [CrossRef] [PubMed]
  81. Schiralli Lester, G.M.; Henderson, A.J. Mechanisms of HIV Transcriptional Regulation and Their Contribution to Latency. Mol. Biol. Int. 2012. [Google Scholar] [CrossRef] [PubMed]
  82. Van Lint, C.; Bouchat, S.; Marcello, A. HIV-1 transcription and latency: An update. Retrovirology 2013. [Google Scholar] [CrossRef] [PubMed]
  83. Griffin, G.E.; Leung, K.; Folks, T.M.; Kunkel, S.; Nabel, G.J. Activation of HIV gene expression during monocyte differentiation by induction of NF-kB. Nature 1989, 339, 70–73. [Google Scholar] [CrossRef]
  84. Moses, A.V.; Ibanez, C.; Gaynor, R.; Ghazal, P.; Nelson, J.A. Differential role of long terminal repeat control elements for the regulation of basal and Tat-mediated transcription of the human immunodeficiency virus in stimulated and unstimulated primary human macrophages. J. Virol. 1994, 68, 298–307. [Google Scholar] [PubMed]
  85. Nabel, G.; Baltimore, D. An inducible transcription factor activates expression of human immunodeficiency virus in T cells. Nature 1987, 326, 711–713. [Google Scholar] [CrossRef] [PubMed]
  86. Perkins, N.D.; Edwards, N.L.; Duckett, C.S.; Agranoff, A.B.; Schmid, R.M.; Nabel, G.J. A cooperative interaction between NF-kappa B and Sp1 is required for HIV-1 enhancer activation. EMBO J. 1993, 12, 3551–3558. [Google Scholar] [PubMed]
  87. Ping, Y.H.; Rana, T.M. DSIF and NELF interact with RNA polymerase II elongation complex and HIV-1 Tat stimulates P-TEFb-mediated phosphorylation of RNA polymerase II and DSIF during transcription elongation. J. Biol. Chem. 2001, 276, 12951–12958. [Google Scholar] [CrossRef] [PubMed]
  88. Natarajan, M.; Schiralli Lester, G.M.; Lee, C.; Missra, A.; Wasserman, G.A.; Steffen, M.; Gilmour, D.S.; Henderson, A.J. Negative elongation factor (NELF) coordinates RNA polymerase II pausing, premature termination, and chromatin remodeling to regulate HIV transcription. J. Biol. Chem. 2013, 288, 25995–26003. [Google Scholar] [CrossRef] [PubMed]
  89. Zhang, Z.; Klatt, A.; Henderson, A.J.; Gilmour, D.S. Transcription termination factor Pcf11 limits the processivity of Pol II on an HIV provirus to repress gene expression. Genes Dev. 2007, 21, 1609–1614. [Google Scholar] [CrossRef] [PubMed]
  90. Wei, P.; Garber, M.E.; Fang, S.-M.; Fischer, W.H.; Jones, K.A. A novel CDK9-associated C-type cyclin interacts directly with HIV-1 Tat and mediates its high-affinity, loop-specific binding to TAR RNA. Cell 1998, 92, 451–462. [Google Scholar] [CrossRef]
  91. Kaczmarek, K.; Morales, A.; Henderson, A.J. T Cell Transcription Factors and Their Impact on HIV Expression. Virol. Res. Treat. 2013, 2013, 41–47. [Google Scholar]
  92. Siliciano, R.F.; Greene, W.C. HIV latency. Cold Spring Harb. Perspect. Med. 2011. [Google Scholar] [CrossRef] [PubMed]
  93. Davey, R.T.; Bhat, N.; Yoder, C.; Chun, T.-W.; Metcalf, J.A.; Dewar, R.; Natarajan, V.; Lempicki, R.A.; Adelsberger, J.W.; Miller, K.D.; et al. HIV-1 and T cell dynamics after interruption of highly active antiretroviral therapy (HAART) in patients with a history of sustained viral suppression. Proc. Natl. Acad. Sci. USA 1999, 96, 15109–15114. [Google Scholar] [CrossRef] [PubMed]
  94. Chun, T.W.; Stuyver, L.; Mizell, S.B.; Ehler, L.A.; Mican, J.A.; Baseler, M.; Lloyd, A.L.; Nowak, M.A.; Fauci, A.S. Presence of an inducible HIV-1 latent reservoir during highly active antiretroviral therapy. Proc. Natl. Acad. Sci. USA 1997, 94, 13193–13197. [Google Scholar] [CrossRef] [PubMed]
  95. Wong, J.K.; Hezareh, M.; Günthard, H.F.; Havlir, D.V.; Ignacio, C.C.; Spina, C.A.; Richman, D.D. Recovery of replication-competent HIV despite prolonged suppression of plasma viremia. Science 1997, 278, 1291–1295. [Google Scholar] [CrossRef] [PubMed]
  96. Verdin, E.; Paras, P., Jr.; van Lint, C. Chromatin disruption in the promoter of human immunodeficiency virus type 1 during transcriptional activation. EMBO J. 1993, 12, 3249–3259. [Google Scholar] [PubMed]
  97. Van Lint, C.; Emiliani, S.; Ott, M.; Verdin, E. Transcriptional activation and chromatin remodeling of the HIV-1 promoter in response to histone acetylation. EMBO J. 1996, 15, 1112–1120. [Google Scholar] [PubMed]
  98. Van Lint, C.; Emiliani, S.; Verdin, E. The expression of a small fraction of cellular genes is changed in response to histone hyperacetylation. Gene Expr. 1996, 5, 245–253. [Google Scholar] [PubMed]
  99. Quivy, V.; de Walque, S.; van Lint, C. Chromatin-associated regulation of HIV-1 transcription: Implications for the development of therapeutic strategies. Subcell. Biochem. 2007, 41, 371–396. [Google Scholar] [PubMed]
  100. He, G.; Margolis, D.M. Counterregulation of chromatin deacetylation and histone deacetylase occupancy at the integrated promoter of human immunodeficiency virus type 1 (HIV-1) by the HIV-1 repressor YY1 and HIV-1 activator Tat. Mol. Cell. Biol. 2002, 22, 2965–2973. [Google Scholar] [CrossRef] [PubMed]
  101. Lusic, M.; Marcello, A.; Cereseto, A.; Giacca, M. Regulation of HIV-1 gene expression by histone acetylation and factor recruitment at the LTR promoter. EMBO J. 2003, 22, 6550–6561. [Google Scholar] [CrossRef] [PubMed]
  102. Thierry, S.; Marechal, V.; Rosenzwajg, M.; Sabbah, M.; Redeuilh, G.; Nicolas, J.C.; Gozlan, J. Cell cycle arrest in G2 induces human immunodeficiency virus type 1 transcriptional activation through histone acetylation and recruitment of CBP, NF-κB, and c-Jun to the long terminal repeat promoter. J. Virol. 2004, 78, 12198–12206. [Google Scholar] [CrossRef] [PubMed]
  103. Hakre, S.; Chavez, L.; Shirakawa, K.; Verdin, E. Epigenetic regulation of HIV latency. Curr. Opin. HIV AIDS 2011, 6, 19–24. [Google Scholar] [CrossRef] [PubMed]
  104. Tyagi, M.; Karn, J. CBF-1 promotes transcriptional silencing during the establishment of HIV-1 latency. EMBO J. 2007, 26, 4985–4995. [Google Scholar] [CrossRef]
  105. Coull, J.J.; Romerio, F.; Sun, J.M.; Volker, J.L.; Galvin, K.M.; Davie, J.R.; Shi, Y.; Hansen, U.; Margolis, D.M. The human factors YY1 and LSF repress the human immunodeficiency virus type 1 long terminal repeat via recruitment of histone deacetylase 1. J. Virol. 2000, 74, 6790–6799. [Google Scholar] [CrossRef] [PubMed]
  106. Williams, S.A.; Chen, L.F.; Kwon, H.; Ruiz-Jarabo, C.M.; Verdin, E.; Greene, W.C. NF-κB p50 promotes HIV latency through HDAC recruitment and repression of transcriptional initiation. EMBO J. 2006, 25, 139–149. [Google Scholar] [CrossRef] [PubMed]
  107. Imai, K.; Okamoto, T. Transcriptional repression of human immunodeficiency virus type 1 by AP-4. J. Biol. Chem. 2006, 281, 12495–12505. [Google Scholar] [CrossRef] [PubMed]
  108. Marban, C.; Suzanne, S.; Dequiedt, F.; de Walque, S.; Redel, L.; van Lint, C.; Aunis, D.; Rohr, O. Recruitment of chromatin-modifying enzymes by CTIP2 promotes HIV-1 transcriptional silencing. EMBO J. 2007, 26, 412–423. [Google Scholar] [CrossRef] [PubMed]
  109. Jiang, G.; Espeseth, A.; Hazuda, D.J.; Margolis, D.M. c-Myc and Sp1 contribute to proviral latency by recruiting histone deacetylase 1 to the human immunodeficiency virus type 1 promoter. J. Virol. 2007, 81, 10914–10923. [Google Scholar] [CrossRef] [PubMed]
  110. Keedy, K.S.; Archin, N.M.; Gates, A.T.; Espeseth, A.; Hazuda, D.J.; Margolis, D.M. A limited group of class I histone deacetylases acts to repress human immunodeficiency virus type 1 expression. J. Virol. 2009, 83, 4749–4756. [Google Scholar] [CrossRef] [PubMed]
  111. Malcolm, T.; Chen, J.; Chang, C.; Sadowski, I. Induction of chromosomally integrated HIV-1 LTR requires RBF-2 (USF/TFII-I) and Ras/MAPK signaling. Virus Genes 2007, 35, 215–223. [Google Scholar] [CrossRef] [PubMed]
  112. Huber, K.; Doyon, G.; Plaks, J.; Fyne, E.; Mellors, J.W.; Sluis-Cremer, N. Inhibitors of histone deacetylases: Correlation between isoform specificity and reactivation of HIV type 1 (HIV-1) from latently infected cells. J. Biol. Chem. 2011, 286, 22211–22218. [Google Scholar] [CrossRef] [PubMed]
  113. Bisgrove, D.A.; Mahmoudi, T.; Henklein, P.; Verdin, E. Conserved P-TEFb-interacting domain of BRD4 inhibits HIV transcription. Proc. Natl. Acad. Sci. USA 2007, 104, 13690–13695. [Google Scholar] [CrossRef] [PubMed]
  114. Yang, Z.; Yik, J.H.; Chen, R.; He, N.; Jang, M.K.; Ozato, K.; Zhou, Q. Recruitment of P-TEFb for stimulation of transcriptional elongation by the bromodomain protein Brd4. Mol. Cell 2005, 19, 535–545. [Google Scholar] [CrossRef] [PubMed]
  115. Zhou, M.; Huang, K.; Jung, K.J.; Cho, W.K.; Klase, Z.; Kashanchi, F.; Pise-Masison, C.A.; Brady, J.N. Bromodomain protein Brd4 regulates human immunodeficiency virus transcription through phosphorylation of CDK9 at threonine 29. J. Virol. 2009, 83, 1036–1044. [Google Scholar] [CrossRef] [PubMed]
  116. Imai, K.; Togami, H.; Okamoto, T. Involvement of histone H3 lysine 9 (H3K9) methyltransferase G9a in the maintenance of HIV-1 latency and its reactivation by BIX01294. J. Biol. Chem. 2010, 285, 16538–16545. [Google Scholar] [CrossRef] [PubMed]
  117. Zhou, M.; Deng, L.; Lacoste, V.; Park, H.U.; Pumfery, A.; Kashanchi, F.; Brady, J.N.; Kumar, A. Coordination of transcription factor phosphorylation and histone methylation by the P-TEFb kinase during human immunodeficiency virus type 1 transcription. J. Virol. 2004, 78, 13522–13533. [Google Scholar] [CrossRef] [PubMed]
  118. Guccione, E.; Bassi, C.; Casadio, F.; Martinato, F.; Cesaroni, M.; Schuchlautz, H.; Luscher, B.; Amati, B. Methylation of histone H3R2 by PRMT6 and H3K4 by an MLL complex are mutually exclusive. Nature 2007, 449, 933–937. [Google Scholar] [CrossRef] [PubMed]
  119. Brown, M.A.; Sims, R.J., 3rd; Gottlieb, P.D.; Tucker, P.W. Identification and characterization of Smyd2: A split SET/MYND domain-containing histone H3 lysine 36-specific methyltransferase that interacts with the Sin3 histone deacetylase complex. Mol. Cancer 2006. [Google Scholar] [CrossRef] [PubMed]
  120. Sakane, N.; Kwon, H.S.; Pagans, S.; Kaehlcke, K.; Mizusawa, Y.; Kamada, M.; Lassen, K.G.; Chan, J.; Greene, W.C.; Schnoelzer, M.; et al. Activation of HIV transcription by the viral Tat protein requires a demethylation step mediated by lysine-specific demethylase 1 (LSD1/KDM1). PLoS Pathog. 2011, 7, e1002184. [Google Scholar] [CrossRef] [PubMed]
  121. Kauder, S.E.; Bosque, A.; Lindqvist, A.; Planelles, V.; Verdin, E. Epigenetic regulation of HIV-1 latency by cytosine methylation. PLoS Pathog. 2009, 5, e1000495. [Google Scholar] [CrossRef] [PubMed]
  122. Blazkova, J.; Trejbalova, K.; Gondois-Rey, F.; Halfon, P.; Philibert, P.; Guiguen, A.; Verdin, E.; Olive, D.; van Lint, C.; Hejnar, J.; et al. CpG methylation controls reactivation of HIV from latency. PLoS Pathog. 2009, 5, e1000554. [Google Scholar] [CrossRef] [PubMed]
  123. Blazkova, J.; Murray, D.; Justement, J.S.; Funk, E.K.; Nelson, A.K.; Moir, S.; Chun, T.W.; Fauci, A.S. Paucity of HIV DNA methylation in latently infected, resting CD4+ T cells from infected individuals receiving antiretroviral therapy. J. Virol. 2012, 86, 5390–5392. [Google Scholar] [CrossRef] [PubMed]
  124. Bowman, G.D. Mechanisms of ATP-dependent nucleosome sliding. Curr. Opin. Struct. Biol. 2010, 20, 73–81. [Google Scholar] [CrossRef] [PubMed]
  125. Henderson, A.; Holloway, A.; Reeves, R.; Tremethick, D.J. Recruitment of SWI/SNF to the human immunodeficiency virus type 1 promoter. Mol. Cell. Biol. 2004, 24, 389–397. [Google Scholar] [CrossRef] [PubMed]
  126. Gallastegui, E.; Millan-Zambrano, G.; Terme, J.M.; Chavez, S.; Jordan, A. Chromatin reassembly factors are involved in transcriptional interference promoting HIV latency. J. Virol. 2011, 85, 3187–3202. [Google Scholar] [CrossRef] [PubMed]
  127. 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]
  128. Tsiang, M.; Jones, G.S.; Niedziela-Majka, A.; Kan, E.; Lansdon, E.B.; Huang, W.; Hung, M.; Samuel, D.; Novikov, N.; Xu, Y.; et al. New class of HIV-1 integrase (IN) inhibitors with a dual mode of action. J. Biol. Chem. 2012, 287, 21189–21203. [Google Scholar] [CrossRef] [PubMed]
  129. Desimmie, B.A.; Schrijvers, R.; Demeulemeester, J.; Borrenberghs, D.; Weydert, C.; Thys, W.; Vets, S.; van Remoortel, B.; Hofkens, J.; de Rijck, J.; et al. LEDGINs inhibit late stage HIV-1 replication by modulating integrase multimerization in the virions. Retrovirology 2013. [Google Scholar] [CrossRef] [PubMed]
  130. 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] [PubMed]
  131. Hamer, D.H. Can HIV be Cured? Mechanisms of HIV persistence and strategies to combat it. Curr. HIV Res. 2004, 2, 99–111. [Google Scholar] [CrossRef] [PubMed]
  132. Gottlicher, M.; Minucci, S.; Zhu, P.; Kramer, O.H.; Schimpf, A.; Giavara, S.; Sleeman, J.P.; Lo Coco, F.; Nervi, C.; Pelicci, P.G.; et al. Valproic acid defines a novel class of HDAC inhibitors inducing differentiation of transformed cells. EMBO J. 2001, 20, 6969–6978. [Google Scholar] [CrossRef] [PubMed]
  133. Lehrman, G.; Hogue, I.B.; Palmer, S.; Jennings, C.; Spina, C.A.; Wiegand, A.; Landay, A.L.; Coombs, R.W.; Richman, D.D.; Mellors, J.W.; et al. Depletion of latent HIV-1 infection in vivo: A proof-of-concept study. Lancet 2005, 366, 549–555. [Google Scholar] [CrossRef]
  134. Archin, N.M.; Eron, J.J.; Palmer, S.; Hartmann-Duff, A.; Martinson, J.A.; Wiegand, A.; Bandarenko, N.; Schmitz, J.L.; Bosch, R.J.; Landay, A.L.; et al. Valproic acid without intensified antiviral therapy has limited impact on persistent HIV infection of resting CD4+ T cells. AIDS 2008, 22, 1131–1135. [Google Scholar] [CrossRef] [PubMed]
  135. Sagot-Lerolle, N.; Lamine, A.; Chaix, M.L.; Boufassa, F.; Aboulker, J.P.; Costagliola, D.; Goujard, C.; Pallier, C.; Delfraissy, J.F.; Lambotte, O. Prolonged valproic acid treatment does not reduce the size of latent HIV reservoir. AIDS 2008, 22, 1125–1129. [Google Scholar] [CrossRef] [PubMed]
  136. Archin, N.M.; Cheema, M.; Parker, D.; Wiegand, A.; Bosch, R.J.; Coffin, J.M.; Eron, J.; Cohen, M.; Margolis, D.M. Antiretroviral intensification and valproic acid lack sustained effect on residual HIV-1 viremia or resting CD4+ cell infection. PLoS ONE 2010, 5, e9390. [Google Scholar] [CrossRef] [PubMed]
  137. Routy, J.P.; Tremblay, C.L.; Angel, J.B.; Trottier, B.; Rouleau, D.; Baril, J.G.; Harris, M.; Trottier, S.; Singer, J.; Chomont, N.; et al. Valproic acid in association with highly active antiretroviral therapy for reducing systemic HIV-1 reservoirs: Results from a multicentre randomized clinical study. HIV Med. 2012, 13, 291–296. [Google Scholar] [CrossRef] [PubMed]
  138. Cysique, L.A.; Maruff, P.; Brew, B.J. Valproic acid is associated with cognitive decline in HIV-infected individuals: A clinical observational study. BMC Neurol. 2006. [Google Scholar] [CrossRef] [PubMed]
  139. Contreras, X.; Schweneker, M.; Chen, C.S.; McCune, J.M.; Deeks, S.G.; Martin, J.; Peterlin, B.M. Suberoylanilide hydroxamic acid reactivates HIV from latently infected cells. J. Biol. Chem. 2009, 284, 6782–6789. [Google Scholar] [CrossRef] [PubMed]
  140. Reuse, S.; Calao, M.; Kabeya, K.; Guiguen, A.; Gatot, J.S.; Quivy, V.; Vanhulle, C.; Lamine, A.; Vaira, D.; Demonte, D.; et al. Synergistic activation of HIV-1 expression by deacetylase inhibitors and prostratin: Implications for treatment of latent infection. PLoS ONE 2009, 4, e6093. [Google Scholar] [CrossRef] [PubMed]
  141. Archin, N.M.; Espeseth, A.; Parker, D.; Cheema, M.; Hazuda, D.; Margolis, D.M. Expression of latent HIV induced by the potent HDAC inhibitor suberoylanilide hydroxamic acid. AIDS Res. Hum. Retrovir. 2009, 25, 207–212. [Google Scholar] [CrossRef] [PubMed]
  142. Archin, N.M.; Liberty, A.L.; Kashuba, A.D.; Choudhary, S.K.; Kuruc, J.D.; Crooks, A.M.; Parker, D.C.; Anderson, E.M.; Kearney, M.F.; Strain, M.C.; et al. Administration of vorinostat disrupts HIV-1 latency in patients on antiretroviral therapy. Nature 2012, 487, 482–485. [Google Scholar] [CrossRef] [PubMed]
  143. Elliott, J.H.; Wightman, F.; Solomon, A.; Ghneim, K.; Ahlers, J.; Cameron, M.J.; Smith, M.Z.; Spelman, T.; McMahon, J.; Velayudham, P.; et al. Activation of HIV transcription with short-course vorinostat in HIV-infected patients on suppressive antiretroviral therapy. PLoS Pathog. 2014, 10, e1004473. [Google Scholar] [CrossRef] [PubMed]
  144. Matalon, S.; Palmer, B.E.; Nold, M.F.; Furlan, A.; Kassu, A.; Fossati, G.; Mascagni, P.; Dinarello, C.A. The histone deacetylase inhibitor ITF2357 decreases surface CXCR4 and CCR5 expression on CD4+ T-cells and monocytes and is superior to valproic acid for latent HIV-1 expression in vitro. J. Acquir. Immune Defic. Syndr. 2010, 54, 1–9. [Google Scholar] [CrossRef] [PubMed]
  145. Rasmussen, T.A.; Schmeltz Sogaard, O.; Brinkmann, C.; Wightman, F.; Lewin, S.R.; Melchjorsen, J.; Dinarello, C.; Ostergaard, L.; Tolstrup, M. Comparison of HDAC inhibitors in clinical development: Effect on HIV production in latently infected cells and T-cell activation. Hum. Vaccines Immunother. 2013, 9, 993–1001. [Google Scholar] [CrossRef] [PubMed]
  146. Lu, H.K.; Gray, L.R.; Wightman, F.; Ellenberg, P.; Khoury, G.; Cheng, W.J.; Mota, T.M.; Wesselingh, S.; Gorry, P.R.; Cameron, P.U.; et al. Ex vivo response to histone deacetylase (HDAC) inhibitors of the HIV long terminal repeat (LTR) derived from HIV-infected patients on antiretroviral therapy. PLoS ONE 2014, 9, e113341. [Google Scholar] [CrossRef] [PubMed]
  147. Wightman, F.; Lu, H.K.; Solomon, A.E.; Saleh, S.; Harman, A.N.; Cunningham, A.L.; Gray, L.; Churchill, M.; Cameron, P.U.; Dear, A.E.; et al. Entinostat is a histone deacetylase inhibitor selective for class 1 histone deacetylases and activates HIV production from latently infected primary T cells. AIDS 2013, 27, 2853–2862. [Google Scholar] [CrossRef] [PubMed]
  148. Wei, D.G.; Chiang, V.; Fyne, E.; Balakrishnan, M.; Barnes, T.; Graupe, M.; Hesselgesser, J.; Irrinki, A.; Murry, J.P.; Stepan, G.; et al. Histone deacetylase inhibitor romidepsin induces HIV expression in CD4 T cells from patients on suppressive antiretroviral therapy at concentrations achieved by clinical dosing. PLoS Pathog. 2014, 10, e1004071. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  149. Rasmussen, T.A.; Tolstrup, M.; Brinkmann, C.R.; Olesen, R.; Erikstrup, C.; Solomon, A.; Winckelmann, A.; Palmer, S.; Dinarello, C.; Buzon, M.; et al. Panobinostat, a histone deacetylase inhibitor, for latent-virus reactivation in HIV-infected patients on suppressive antiretroviral therapy: A phase 1/2, single group, clinical trial. Lancet HIV 2014, 1, e13–e21. [Google Scholar] [CrossRef]
  150. Banerjee, C.; Archin, N.; Michaels, D.; Belkina, A.C.; Denis, G.V.; Bradner, J.; Sebastiani, P.; Margolis, D.M.; Montano, M. BET bromodomain inhibition as a novel strategy for reactivation of HIV-1. J. Leukoc. Biol. 2012, 92, 1147–1154. [Google Scholar] [CrossRef] [PubMed]
  151. Bartholomeeusen, K.; Xiang, Y.; Fujinaga, K.; Peterlin, B.M. Bromodomain and extra-terminal (BET) bromodomain inhibition activate transcription via transient release of positive transcription elongation factor b (P-TEFb) from 7SK small nuclear ribonucleoprotein. J. Biol. Chem. 2012, 287, 36609–36616. [Google Scholar] [CrossRef] [PubMed]
  152. Zhu, J.; Gaiha, G.D.; John, S.P.; Pertel, T.; Chin, C.R.; Gao, G.; Qu, H.; Walker, B.D.; Elledge, S.J.; Brass, A.L. Reactivation of latent HIV-1 by inhibition of BRD4. Cell Rep. 2012, 2, 807–816. [Google Scholar] [CrossRef] [PubMed]
  153. Li, Z.; Guo, J.; Wu, Y.; Zhou, Q. The BET bromodomain inhibitor JQ1 activates HIV latency through antagonizing Brd4 inhibition of Tat-transactivation. Nucleic Acids Res. 2013, 41, 277–287. [Google Scholar] [CrossRef] [PubMed]
  154. Boehm, D.; Calvanese, V.; Dar, R.D.; Xing, S.; Schroeder, S.; Martins, L.; Aull, K.; Li, P.C.; Planelles, V.; Bradner, J.E.; et al. BET bromodomain-targeting compounds reactivate HIV from latency via a Tat-independent mechanism. Cell Cycle 2013, 12, 452–462. [Google Scholar] [CrossRef] [PubMed]
  155. Bernhard, W.; Barreto, K.; Saunders, A.; Dahabieh, M.S.; Johnson, P.; Sadowski, I. The Suv39H1 methyltransferase inhibitor chaetocin causes induction of integrated HIV-1 without producing a T cell response. FEBS Lett. 2011, 585, 3549–3554. [Google Scholar] [CrossRef] [PubMed]
  156. Bouchat, S.; Gatot, J.S.; Kabeya, K.; Cardona, C.; Colin, L.; Herbein, G.; de Wit, S.; Clumeck, N.; Lambotte, O.; Rouzioux, C.; et al. Histone methyltransferase inhibitors induce HIV-1 recovery in resting CD4+ T cells from HIV-1-infected HAART-treated patients. AIDS 2012, 26, 1473–1482. [Google Scholar] [CrossRef] [PubMed]
  157. Friedman, J.; Cho, W.K.; Chu, C.K.; Keedy, K.S.; Archin, N.M.; Margolis, D.M.; Karn, J. Epigenetic silencing of HIV-1 by the histone H3 lysine 27 methyltransferase enhancer of Zeste 2. J. Virol. 2011, 85, 9078–9089. [Google Scholar] [CrossRef]
  158. Burnett, J.C.; Lim, K.I.; Calafi, A.; Rossi, J.J.; Schaffer, D.V.; Arkin, A.P. Combinatorial latency reactivation for HIV-1 subtypes and variants. J. Virol. 2010, 84, 5958–5974. [Google Scholar] [CrossRef]
  159. Laird, G.M.; Bullen, C.K.; Rosenbloom, D.I.; Martin, A.R.; Hill, A.L.; Durand, C.M.; Siliciano, J.D.; Siliciano, R.F. Ex vivo analysis identifies effective HIV-1 latency-reversing drug combinations. J. Clin. Investig. 2015, 125, 1901–1912. [Google Scholar] [CrossRef] [PubMed]
  160. Dar, R.D.; Hosmane, N.N.; Arkin, M.R.; Siliciano, R.F.; Weinberger, L.S. Screening for noise in gene expression identifies drug synergies. Science 2014, 344, 1392–1396. [Google Scholar] [CrossRef]
  161. Finzi, D.; Hermankova, M.; Pierson, T.; Carruth, L.M.; Buck, C.; Chaisson, R.E.; Quinn, T.C.; Chadwick, K.; Margolick, J.; Brookmeyer, R.; et al. Identification of a reservoir for HIV-1 in patients on highly active antiretroviral therapy. Science 1997, 278, 1295–1300. [Google Scholar] [CrossRef] [PubMed]
  162. Siliciano, J.D.; Kajdas, J.; Finzi, D.; Quinn, T.C.; Chadwick, K.; Margolick, J.B.; Kovacs, C.; Gange, S.J.; Siliciano, R.F. Long-term follow-up studies confirm the stability of the latent reservoir for HIV-1 in resting CD4+ T cells. Nat. Med. 2003, 9, 727–728. [Google Scholar] [CrossRef] [PubMed]
  163. Eriksson, S.; Graf, E.H.; Dahl, V.; Strain, M.C.; Yukl, S.A.; Lysenko, E.S.; Bosch, R.J.; Lai, J.; Chioma, S.; Emad, F.; et al. Comparative analysis of measures of viral reservoirs in HIV-1 eradication studies. PLoS Pathog. 2013, 9, e1003174. [Google Scholar] [CrossRef] [PubMed]
  164. Ho, Y.C.; Shan, L.; Hosmane, N.N.; Wang, J.; Laskey, S.B.; Rosenbloom, D.I.; Lai, J.; Blankson, J.N.; Siliciano, J.D.; Siliciano, R.F. Replication-competent noninduced proviruses in the latent reservoir increase barrier to HIV-1 cure. Cell 2013, 155, 540–551. [Google Scholar] [CrossRef] [PubMed]

Share and Cite

MDPI and ACS Style

Agosto, L.M.; Gagne, M.; Henderson, A.J. Impact of Chromatin on HIV Replication. Genes 2015, 6, 957-976. https://doi.org/10.3390/genes6040957

AMA Style

Agosto LM, Gagne M, Henderson AJ. Impact of Chromatin on HIV Replication. Genes. 2015; 6(4):957-976. https://doi.org/10.3390/genes6040957

Chicago/Turabian Style

Agosto, Luis M., Matthew Gagne, and Andrew J. Henderson. 2015. "Impact of Chromatin on HIV Replication" Genes 6, no. 4: 957-976. https://doi.org/10.3390/genes6040957

Article Metrics

Back to TopTop