Skip to main content

Latency: the hidden HIV-1 challenge

Abstract

Eradication of HIV-1 from an infected individual cannot be achieved by current regimens. Viral reservoirs established early during the infection remain unaffected by anti-retroviral therapy for a long time and are able to replenish systemic infection upon interruption of the treatment. Therapeutic targeting of viral latency will require a better understanding of the basic mechanisms underlying the establishment and long-term maintenance of HIV-1 in resting memory CD4 T cells, the most prominent reservoir of transcriptionally silent provirus. Since the molecular mechanisms that permit long term transcriptional control of proviral gene expression in these cells are still obscure, this review aims at summarizing the various aspects of the problem that need to be considered. In particular, this review will focus the attention on the control of transcription imposed by chromatin through various epigenetic mechanisms. Exploring the molecular details of viral latency will provide new insights for eventual future therapeutics that aim at viral eradication.

Introduction

The major obstacle to HIV-1 eradication is the establishment of a latent infection. In infected individuals, viral production is a dynamic process involving continuous rounds of infection of CD4+ T lymphocytes with rapid turnover of both free virus and virus-producing cells that have a half-life of 1–2 days [1, 2]. The decay curves of plasma viremia following antiretroviral treatment have shown that after an initial fast decay, that wipes out the majority of circulating viruses in 1–2 weeks, plasma virus declines at a lower rate [3, 4]. The half-life of this compartment was estimated to be 1–4 weeks, but the nature of the cellular reservoir responsible for the second phase in the decay curve is still unclear. These cells could be macrophages, which are less sensitive to the cytopathic effect of HIV-1 infection [5] and that once terminally differentiated have a turnover rate of approximately 2 weeks. In addition, cellular reservoirs for HIV-1 could also be CD4+ T lymphocytes not fully activated, which carry the integrated provirus in a non-replicative state until the activation process is complete. Finally, dendritic cells (DCs) may also delay the release of infectious virus, since they are not permissive for HIV infection but can carry the virus trapped on their surfaces [6].

After two months on HAART the plasma levels of genomic RNA falls below the limit of detection in most previously untreated patients. Therefore, it was initially assumed that prolonged treatment might lead to eradication of the virus in these patients [3]. Unfortunately, it is now clear that long-lived reservoirs of HIV-1 can persist for years in the presence of HAART. Although certain tissues like the male urogenital tract or the central nervous system might preserve infectious virus [7, 8] the reservoir that appears to be the major barrier to eradication is composed of latently infected resting memory CD4+ T cells that carry an integrated provirus that is transcriptionally silent [9, 10]. The extremely long half-life of these cells, combined with a tight control of HIV-1 expression, make this reservoir ideally suited to maintain hidden copies of the virus, which are in turn able to trigger a novel systemic infection upon discontinuation of therapy. Given the importance of this reservoir, a lot of effort has been invested to characterize these cells from infected patients. These studies will be discussed in the following chapters, which will also address the problem of choosing appropriate model systems to thoroughly characterize the molecular determinants that allow the provirus to remain silent. Such mechanisms are mostly related to transcriptional control of viral expression and they depend both on the host cells and the virus. Finally, some ideas on how to approach viral eradication in the light of these novel findings will be presented in the final chapter.

Source of latently infected cells

HIV-1 exploits different strategies to persist within infected individuals. In CD4+ T lymphocytes, the replicative state of the virus is dependent upon the cell cycle of the host cell. Whereas HIV-1 entry into activated CD4+ lymphocytes leads to a productive infection [11], the virus encounter several blocks prior to integration in resting CD4+ lymphocytes [12]. Such post-entry blocks have been proposed to result from a delay in completing reverse transcription due to low nucleotide pools and to the inability to import the pre-integration complex into the nucleus [13–16]. Most recently the anti-retroviral deoxycytidine deaminase APOBEC3G has been shown to strongly protect unstimulated peripheral blood CD4+ T cells against HIV-1 infection [17]. Furthermore, in Old World primates, TRIM5α, a component of cytoplasmic bodies, confers a potent block to human immunodeficiency virus type 1 (HIV-1) infection that acts after virus entry into cells, probably at the level of capsid processing [18]. While these blocks delay the production of progeny virus following the infection of CD4+ resting T cells, mitogenic stimuli are able to trigger viral replication and release of infectious virus [13, 14, 19–21]. Although one would expect that activation of the cell per se would allow more efficient reverse transcription and nuclear import, by increasing the pools of DNA precursors and the available ATP used for the active mobilization of the PIC, still it is possible that specific blocks must be removed to allow full recovery of HIV-1 infectivity. In the case of APOBEC3G for example, activation of resting T cells induces the shift of the active low-molecular-mass form of APOBEC3G to an inactive high-molecular-mass complex unable to restrict viral infection [17]. Other types of replication blocks that act after provirus integration in resting T cells, like for example the inhibition of NF-κB activity by Murr1 [22], will be discussed in the following chapters.

Regardless of the kind of restriction imposed by the resting T cells on viral replication, this reservoir of pre-integrated latent virus is relatively labile persisting for weeks and cannot be accounted for the long-term latency observed during HAART.

In addition to CD4+ T lymphocytes, dendritic cells and macrophages are considered reservoirs for HIV-1 infection, but information on the replicative state of the virus within these cells is limited. DCs capture and internalize extracellular virions via the DC-SIGN lectin. Captured virions can subsequently be transmitted to T cells in trans [6]. However, DC-SIGN does not significantly protect captured virions against degradation, leading to loss of infectivity within several hours [23]. In HIV-1-infected monocyte-derived macrophages, mature viral particles can be observed within late endosomes [24]. Virions found within monocyte-derived macrophages persist and retain infectivity for weeks, thus providing an additional mechanism for viral persistence [25]. HIV-1 hidden in DCs and macrophages certainly plays an important role for viral spread and cell-cell transmission, but its involvement in long-term viral persistence has yet to be demonstrated.

A more stable form of latency occurs in CD4+ T cells that carry an integrated provirus. In principle, since integration requires T cell activation to allow efficient reverse-transcription and nuclear import of the pre-integration complex, post-integration latency can result only from the return of an infected activated T cell to a quiescent state. Evidence to this model has come from studies from Siliciano and co-workers who demonstrated the existence of resting memory CD4+ T cells carrying an integrated provirus in vivo [9]. The phenotype of these resting cells carrying a non-productive HIV-1 infection, derived from peripheral blood of patients undergoing antiretroviral therapy with low to undetectable viremia, indicates that they derive from infected CD4+ lymphoblasts that have reverted to a resting memory state. These cells show a specific set of surface markers (such as CD4+, CD25-, CD69-, HLA-DR-) and are positive for the integrated provirus. Most importantly, upon mitogen stimulation, infectious virus can be recovered from these cells, indeed demonstrating that they represent a true, inducible viral reservoir. Hence, in vivo, it appears that HIV-1 gets trapped in T cells that revert to a resting memory state. In some respect long term HIV-1 persistence reflects directly long-term T-cell memory. At any given time, most CD4+ T lymphocytes in the body are in a resting G0 state. In response to antigens, resting T cells undergo a burst of cellular proliferation and differentiation, giving rise to effector cells. Most effector cells die quickly, but a subset survives and reverts to a resting G0 state (contraction phase). These lymphocytes persist as memory cells, with an altered pattern of gene expression enabling long-term survival and rapid responses to the relevant antigen in the future. Activated CD4+ cells are highly susceptible to HIV-1 infection and typically die quickly as a result of the cytopathic effects either of the virus or of the host immune response. However, some activated CD4 cells may become infected and then survive long enough to revert back to a resting state. Unfortunately, our current understanding of the decisive factors that determine if a CD4+ T cells will die or become a memory cell, as well as those that allow the self-renewal of resting memory cells for a lifetime, is still largely incomplete.

Models for latently infected cells

Despite the great wealth of information on the regulation of HIV-1 transcription, the crucial molecular events that control maintenance of the quiescent state in resting T cells remain elusive. Part of the problem depends on the lack of an appropriate model system. Most cell lines carrying an integrated quiescent provirus have been derived from transformed lymphoblasts that have been infected ex vivo with HIV-1. After an initial burst of viral replication and cell death, a population of non-productive cells carrying an integrated provirus remains. Establishment of latency in these cell lines is driven by selection of cells that resist viral replication and has been linked to mutation in viral genes, to certain cellular proteins and to the site of integration [26–29]. Alternatively, T cells can be transduced with an HIV-1 vector carrying Tat and a reporter gene. This method has allowed the characterization of the integration status irrespective of the replication of the virus and has provided useful insights on the status of the integrated provirus. However, the constantly activated and proliferating nature of these cells, either infected or transduced, does not accurately represent the quiescent cellular environment of latently infected cells in vivo. A convenient animal model that recapitulates HIV-1 latency does not exist. In fact, several blocks to HIV-1 infection in mice greatly impair the development of an animal model to study HIV-1 infection amenable to genetic manipulation. One possibility would be the use of the SCID-hu (Thy/Liv) mouse that carries a source of human hematopoietic progenitor cells and human fetal thymus to provide a microenvironment for human T cells lymphopoiesis. Latently HIV-1 infected CD4+ T cells can be obtained in this model, but the exact extent to which it can be applied to natural infection is not known [30]. SIV macaque models of AIDS are well established and have been extremely useful in HIV-1 vaccine development and in advancing the understanding of the pathogenesis of AIDS. The first report exploiting the SIV-macaque model to study viral infection in the course of antiretroviral therapy showed persistence of the virus in resting CD4+ T cells with many similarities to the human situation [31]. This study provides initial evidence for the utility of a closely related retroviral infection for the analysis of HIV-1 persistence during antiretroviral treatment. However, the complexity of the protocol, which also requires antiretroviral drugs especially designed to control SIV infection, makes this model impractical if only for the preclinical evaluation of novel strategies to target viral reservoirs.

Possible molecular mechanisms behind latency

Since the HIV-1 provirus is found integrated into the host genome, regulation of viral gene expression depends on the chromatin environment at the site of integration and on the interaction of the viral Tat trans-activator with host factors. Clearly, multiple mechanisms could concur in this process.

I) Cis- and trans-acting factors involved in HIV-1 silencing

The U3 region of the HIV-1 LTR functions as the viral promoter and contains consensus sequences for several transcription factors, including NFAT and NF-κB, involved also as positive regulators of cell activation in uninfected T cells [32]. NF-κB is a key host transcription factor required for LTR activation [33]. In resting T cells NF-κB is sequestered in the cytoplasm bound to IκB and it is transported to the nucleus following cellular activation by TCR engagement or stimulation by IL-2 or TNFα. NF-κB interacts with two highly conserved binding sites found in the viral LTR and promotes transcriptional activation. Murr1, previously known for its involvement in copper regulation, has been recently shown to inhibit basal and cytokine-stimulated NF-κB [22]. Most importantly, knockdown of Murr1 by RNAi in primary resting CD4+ lymphocytes increased HIV-1 replication. Thus, Murr-1 acts as a host restriction factor that inhibits HIV-1 replication in resting T cells.

In addition to host transcription factors, HIV-1 transcription is boosted by the viral Tat trans-activator, a highly unusual protein that interacts with a cis-acting RNA element (trans-activation-responsive region; TAR) present at the 5' end of each viral transcript [34]. Through this interaction, the protein activates HIV-1 transcription by promoting the assembly of transcriptionally active complexes at the LTR through multiple protein-RNA and protein-protein interactions. Tat interacts directly with cyclin T1, the cyclin component of CDK9, which phosphorylates the carboxy-terminal domain of RNA polymerase II to enhance its processivity [35, 36] (reviewed in: [37]). Tat-induced transcriptional activation of the LTR promoter is concomitant with recruitment of the transcriptional co-activators p300 and the highly homologue cAMP-responsive transcription factor binding protein (CBP) [38–41]. These large proteins are histone acetyl-transferases capable of modulating the interaction of nucleosomes with DNA and with other factors involved in transcription. In fact, besides histones, Tat itself is a substrate for the enzymatic activity of p300/CBP and of the associated factor P/CAF and is regulated by acetylation/deacetylation [42–47]. Furthermore, Tat is also tightly regulated by ubiquitination, further highlighting the intimate interplay between the viral trans-activator and the host cell [48].

Tat-associated proteins could be one of the limiting factors for processive transcription in resting T cells. In this respect, the low levels of P-TEFb kinase activity (CDK9 and Cyclin T1) that have been observed in resting T cells are increased in response to activating stimuli [49]. Tat itself could be the main limiting factor being subject to tight post-translational regulation by acetylation and ubiquitylation [42, 46, 48]. These data fit in a model where limiting availability of host cell's factors and/or the viral trans-activator concur in maintaining the virus transcriptionally silent. Possible mechanisms propose premature termination of transcription due to the absence of sufficient concentrations of Tat and NF-κB [50, 51] or inefficient export of RNAs for structural proteins [52]. A recent report has also shown that fluctuations in Tat expression alone govern stochastic gene expression of the viral LTR [53]. However, when analyzing these studies one should always keep in mind that they should hold true also in resting T cells in vivo.

II) Integration-site-dependent determinants of HIV-1 silencing

HIV-1 is found integrated into the genome of resting memory T cells, hence the chromatin status at the site of integration determines whether the provirus is transcriptionally active, poised for activation or inactive. A recent report [54] has analyzed the integration site of HIV-1 in resting memory CD4+ cells derived from patient on highly active antiretroviral treatment. Surprisingly, HIV-1 has been found in intronic regions of actively transcribed genes. Consistently, HIV-1 sequences were included in the unspliced RNAs of these genes. These findings, although complicated by the high levels of dead integration events observed (i.e. only a small fraction of resting T cells carrying a silent provirus becomes productive when activated), correlate with the observation that HIV-1 integrates in transcriptionally active genes during productive infection of cultured T cells [55], but are in sharp contrast with previous work showing that HIV-1 infected T cell lines selected for a quiescent state of the provirus show preferential integration into heterochromatin [28, 56]. A very recent study helps clarifying this point showing that the integration site of quiescent/inducible HIV-1 vectors in T cell lines could be associated with heterochromatin, as reported, but also with actively transcribing genes, thus confirming the analysis in patients' cells [57]. Another interesting observation of this work is that the inducible state of the HIV-1 provirus could depend also upon integration in intergenic regions of gene-poor chromosomes. However, all these three conditions: integration into heterochromatin, integration into highly transcribed genes and integration into gene poor chromosomes, account for only 40% of the inducible integration events in this system. Notably, stochastic gene expression from the viral LTR, a phenotype dependent on the levels of Tat, occurs with a high frequency within 1 kb of a human endogenous retrovirus LTR [53]. These observations deserve further study since they may indicate that other as yet not identified chromatin environments at the site of provirus integration control transcriptional silencing and reactivation. In fact, a totally different pattern might emerge, particularly considering novel concepts on the relationship between spatial positioning of chromosomes within the nucleus and transcription activity [58]. T cells switching from a memory state to a lymphoblast and vice-versa undergo a program of spatial genome reorganization of specific genes [59–61]. HIV-1 proviruses "trapped" in a gene that is being spatially confined in a silenced state might become inactive and poised for activation from external stimuli [62].

So far we have considered post-integration transcriptional silencing as a passive consequence of chromatin status. However, we might also consider a certain degree of bias on the integration site driven by the association of the pre-integration complex with certain cellular factor such as the high mobility group (HMG) protein A1, the barrier to autointegration (BAF), the INI1 homologous of the SNF5 component of the ATP-dependent chromatin remodeling complex SWI/SNF and LEDGF/p75 [63–65]. BAF binds directly to double-stranded DNA, nuclear LEM-domain proteins, lamin A and transcriptional activators [66]. Recent observations suggest that BAF has structural roles in nuclear assembly and chromatin organization and might interlink chromatin structure, nuclear architecture and gene regulation. Integrase associates also with INI implying a role of transcription-related ATP-remodeling complexes in determining integration. Particularly interesting is the fact that INI associates with the promyelocytic leukemia protein PML, the principal component of the nuclear bodies [67]. These nuclear structures, whose function is still largely unknown, are dynamically associated to the transcriptional Cyclin T1 and to co-activators such as p300/CBP [68]. Intriguingly, IN binds also p300 and this interaction is involved in the integration process through acetylation of IN itself [69]. As a note of caution it should be said that none of the above mentioned factors that interact with the viral integrase has been shown to be functionally expressed in resting CD4+ T cells in relation to HIV infection. Future research will tell us more on these interactions of HIV-1 integrase and their role in HIV-1 integration.

III) A role for RNA interference in HIV-1 silencing?

RNA Interference (RNAi) was first identified as a post-transcriptional response to exogenous double-stranded RNA (dsRNA) introduced in C. elegans, but this mechanism is conserved from plants to nematodes and mammals [70–72]. RNAi is triggered by long dsRNA cleaved by the cytoplasmic RNaselll enzyme Dicer into short, interfering RNAs (siRNAs). One strand of the siRNA is incorporated into the effector complex of RNAi, the RNA-induced Silencing Complex (RISC). The short RNA guides RISC to target complementary mRNA and catalyzes an endonucleolytic cleavage, resulting in post-transcriptional gene silencing (PTGS) of gene expression. In mammalian cells, siRNAs are recognized by the pathway responsible for the activities of a class of endogenous 21–22 nt microRNAs (miRNAs) (for a recent review see [73]). miRNAs are first produced as long hairpinned precursor dsRNAs transcribed by RNA polymerase II and are sequentially processed by the nucleases Drosha and Dicer. The short dsRNA produced are thought to regulate gene expression mainly at the translational level. Many different miRIMA genes have been predicted in humans, and they have been implicated in the regulation of genes involved in development and growth control [74].

RNAi-mediated pathways of transcriptional silencing have also been shown to induce chromatin modifications at the homologous genomic locus in plants and lower eukaryotes (for a review: [75]). Transposable elements and related repeats are primary targets for RNAi-mediated pathways in the nucleus, consistent with a role for RNAi in host defense against invasive viral sequences (for a recent review: [76]). Silencing occurs through CG methylation by specific DNA methyltransferases that are directed to the target by the methylation of lysine 9 of histone 3 (H3K9-Me). These findings have led to a model whereby siRNAs directed de novo DNA methylation through the successive action of a histone methyltransferase and DNA methyltransferases that maintain methylation at target DNA loci (RITS, RNA-induced transcriptional silencing) [77].

Artificial RNAi can efficiently suppress several human viruses, including HIV-1 [78, 79]. However, this effect is shot-term, since the virus is capable of evading the siRNA response by random mutation of the target sequence, thus limiting the efficacy of this antiviral approach [80]. Not surprisingly, several viruses encode also their own miRNA that can modulate viral replication (reviewed in: [81]). Herpesviruses like EBV and HSHV encode miRNAs directed against cellular and viral targets and are believed to regulate the latent/lytic transition of these viruses [82, 83]. Short-hairpin siRNA precursors have also been found in the HIV-1 genome. Such sequences are involved in suppression of viral transcription unless counteracted by the inhibition of endogenous Dicer activity targeted by the viral Tat transactivator [84]. This mechanism is similar to what has been observed for the primate foamy retrovirus PFV-1 that is capable of subverting a cellular miRNA block through the activity of the viral Tas protein [85]. Another HIV-1-encoded miRNA has been identified in the nef gene and has been speculated to be a possible determinant of long-term non-progression to AIDS through inhibition of Nef function [86].

It would be intriguing to speculate also about the existence of a transcriptional pathway of HIV-1 gene silencing, taking into account previous observations that might [87] link CpG methylation at the HIV-1 promoter to transcriptional silencing [88]. As a note of caution, however, it should be observed that at present RNAi mediated transcriptional gene silencing in human cells is highly controversial, and care should be taken in extrapolating data obtained in lower eukaryotes. Nevertheless, models such as HIV-1 could help in disclose these archival protective mechanisms in human cells, if they exist.

Potential therapies to eliminate latently infected cells

Although the implementation of HAART has improved the survival and quality of life of HIV-infected individuals, HIV cannot yet be eradicated from infected individuals. Several studies have demonstrated that in individuals receiving HAART, the frequency of HIV-infected cells is reduced to fewer than one cell per 106 resting CD4+ T cells [10, 89, 90]. However, even after years with viremia below the limit of quantification, the frequency of these infected cells does not decrease further. The therapeutic approaches evaluated to date have failed to demonstrate a significant and persistent decline of this latent viral reservoir [91], which appears small but stable and contains both wild-type and drug-resistant viral species [92]. Several studies have shown that intensive antiretroviral therapy in combination with interleukin-2 or global T cell activators fails to eradicate HIV-1 infection. Global T cell activation may instead induce viral replication and increase the number of susceptible uninfected target cells beyond the threshold that can be contained by antiretroviral therapy [93]. Following this approach, a strategy that selectively activates quiescent proviral genomes with limited effects on the host cell exploited the properties of the phorbol ester Prostratin or the human cytokine interleukin-7 that have been reported to reactivate latent HIV-1 in the absence of cellular proliferation [30, 94–96]. Another promising agent that has been proposed is the histone-deacetylase inhibitor Valproic acid capable of inducing outgrowth of HIV-1 from resting CD4+ cells of aviremic patients without full activation of the cells from the quiescent state [97]. Such treatments, in combination with antiretroviral therapy, should allow outgrowth of latent HIV-1 but avoid the pitfalls of global T cell activation.

Another approach would be to target and destroy CD4+ memory cells. A study has been conducted ex vivo with an anti-CD45RO ricin immunotoxin to decrease the number of latently infected CD4+ T cells obtained from HIV-infected individuals without detectable plasma viremia [98]. Such treatment significantly reduced the frequency of CD4+ memory cells with only a modest effect on the memory responses of CD8+ T cells. Therefore, purging latent cells from infected individuals on highly active antiretroviral therapy might reduce the HIV latent reservoir without seriously compromising CD8+ T cell memory responses. However, this kind of approach targets both infected and non-infected resting T cells and would severely deplete the immunological memory of the patient.

Conclusion

Recent advances have identified a long-lived stable reservoir of HIV-1 in patients on effective antiretroviral therapy that could potentially persist for life. This reservoir consists of a small pool of resting CD4+ T cells carrying an integrated provirus that somehow control viral expression allowing the virus to remain undetectable in plasma (Figure 1). Discontinuation of therapy and activation of cells results in production of infectious virus leading to a novel systemic infection. Hence, elimination of this reservoir by novel therapeutic approaches will be required before eradication can be achieved. Our knowledge of the cellular and molecular mechanisms behind the establishment of this remarkably stable reservoir will depend on appropriate in vitro model systems and in vivo animal models. These should allow the dissection of the pathways leading to transcriptional control of HIV-1 replication. In particular it will be crucial to correlate the activation state of host cells with the transcription state of the provirus. Key elements to be studied are: the site of provirus integration, which is determined by the viral integrase, and the mechanisms that control gene expression at these chromatin loci, which depend mostly on the activity of the Tat transactivator.

Figure 1
figure 1

Schematic description of the HIV-1 life cycle highlighting the various blocks that can delay viral replication leading to prolonged hiding of the virus in the host cell during HAART. These include: (i) pre-integration blocks like the deoxycytidine deaminase APOBEC3G, the cytoplasmic body component TRIM5α (in Old World monkeys), incomplete reverse-transcription and defects in nuclear import; (ii) post-integration blocks such as integration into heterochromatin where transcription is repressed, ineffective RNAPII elongation in the absence of Tat or of key host factors, regulation of NF-kB by Murr-1; (iii) translational blocks induced by RNAi.

It is time to seriously consider alternative therapies that take into account the aim of eradicating infectious virus from the patients. In order to do so an effort is needed toward the understanding of the intimate interplay that is established at the molecular level between the virus and the host cell. From such studies it will be possible to devise novel therapies that selectively target the hidden virus.

Abbreviations

HIV-1:

human immunodeficiency virus type 1

HAART:

highly active antiretroviral therapy

CD4/CD3/CD28:

cluster of differentiation 4, 3, 28

IL-2:

interleukin 2

PHA:

phytohemagglutinin

PIC:

pre-integration complex

IN:

integrase

HMG:

high mobility group

BAF:

barrier to autointegration factor

INI1:

integrase interactor 1

LEDGF:

lens epithelium-derived growth factor

P/CAF:

p300/CBP associated factor

CBP:

CREB binding protein

CDK9:

cyclin-dependent kinase 9

TPA:

12-O-tetradecanoyl-phorbol-13-acetate

LTR:

long terminal repeat TAR, trans-activating response region

SCID:

Severe Combined Immune Deficiency

PML:

promyelocytic leukemia

PTGS:

post-transcriptional gene silencing

RISC:

RNA-induced silencing complex

RITS:

RNA-induced transcriptional silencing

siRNA:

small-inhibitory RNA.

References

  1. Wei X, Ghosh SK, Taylor ME, Johnson VA, Emini EA, Deutsch P, Lifson JD, Bonhoeffer S, Nowak MA, Hahn BH, et al: Viral dynamics in human immunodeficiency virus type 1 infection. Nature. 1995, 373: 117-122. 10.1038/373117a0.

    Article  CAS  PubMed  Google Scholar 

  2. Ho DD, Neumann AU, Perelson AS, Chen W, Leonard JM, Markowitz M: Rapid turnover of plasma virions and CD4 lymphocytes in HIV-1 infection. Nature. 1995, 373: 123-126. 10.1038/373123a0.

    Article  CAS  PubMed  Google Scholar 

  3. Perelson AS, Neumann AU, Markowitz M, Leonard JM, Ho DD: HIV-1 dynamics in vivo: virion clearance rate, infected cell life-span, and viral generation time. Science. 1996, 271: 1582-1586.

    Article  CAS  PubMed  Google Scholar 

  4. Cavert W, Notermans DW, Staskus K, Wietgrefe SW, Zupancic M, Gebhard K, Henry K, Zhang ZQ, Mills R, McDade H, et al: Kinetics of response in lymphoid tissues to antiretroviral therapy of HIV-1 infection. Science. 1997, 276: 960-964. 10.1126/science.276.5314.960.

    Article  CAS  PubMed  Google Scholar 

  5. Ho DD, Rota TR, Hirsch MS: Infection of onocyte/macrophages by human T lymphotropic virus type III. J Clin Invest. 1986, 77: 1712-1715.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  6. Geijtenbeek TB, Torensma R, van Vliet SJ, van Duijnhoven GC, Adema GJ, van Kooyk Y, Figdor CG: Identification of DC-SIGN, a novel dendritic cell-specific ICAM-3 receptor that supports primary immune responses. Cell. 2000, 100: 575-585. 10.1016/S0092-8674(00)80693-5.

    Article  CAS  PubMed  Google Scholar 

  7. Zhu T, Wang N, Carr A, Nam DS, Moor-Jankowski R, Cooper DA, Ho DD: Genetic characterization of human immunodeficiency virus type 1 in blood and genital secretions: evidence for viral compartmentalization and selection during sexual transmission. J Virol. 1996, 70: 3098-3107.

    PubMed Central  CAS  PubMed  Google Scholar 

  8. Cheng-Mayer C, Weiss C, Seto D, Levy JA: Isolates of human immunodeficiency virus type 1 from the brain may constitute a special group of the AIDS virus. Proc Natl Acad Sci USA. 1989, 86: 8575-8579.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  9. Chun TW, Finzi D, Margolick J, Chadwick K, Schwartz D, Siliciano RF: In vivo fate of HIV-1-infected T cells: quantitative analysis of the transition to stable latency. Nat Med. 1995, 1: 1284-1290. 10.1038/nm1295-1284.

    Article  CAS  PubMed  Google Scholar 

  10. Chun TW, Stuyver L, Mizell SB, Ehler LA, Mican JA, Baseler M, Lloyd AL, Nowak MA, Fauci AS: Presence of an inducible HIV-1 latent reservoir during highly active antiretroviral therapy. Proc Natl Acad Sci USA. 1997, 94: 13193-13197. 10.1073/pnas.94.24.13193.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  11. McDougal JS, Mawle A, Cort SP, Nicholson JK, Cross GD, Scheppler-Campbell JA, Hicks D, Sligh J: Cellular tropism of the human retrovirus HTLV-III/LAV. I. Role of T cell activation and expression of the T4 antigen. J Immunol. 1985, 135: 3151-3162.

    CAS  PubMed  Google Scholar 

  12. Wong JK, Hezareh M, Gunthard HF, Havlir DV, Ignacio CC, Spina CA, Richman DD: Recovery of replication-competent HIV despite prolonged suppression of plasma viremia. Science. 1997, 278: 1291-1295. 10.1126/science.278.5341.1291.

    Article  CAS  PubMed  Google Scholar 

  13. Zack JA, Arrigo SJ, Weitsman SR, Go AS, Haislip A, Chen IS: HIV-1 entry into quiescent primary lymphocytes: molecular analysis reveals a labile, latent viral structure. Cell. 1990, 61: 213-222. 10.1016/0092-8674(90)90802-L.

    Article  CAS  PubMed  Google Scholar 

  14. Zack JA, Haislip AM, Krogstad P, Chen IS: Incompletely reverse-transcribed human immunodeficiency virus type 1 genomes in quiescent cells can function as intermediates in the retroviral life cycle. J Virol. 1992, 66: 1717-1725.

    PubMed Central  CAS  PubMed  Google Scholar 

  15. Korin YD, Zack JA: Nonproductive human immunodeficiency virus type 1 infection in nucleoside-treated G0 lymphocytes. J Virol. 1999, 73: 6526-6532.

    PubMed Central  CAS  PubMed  Google Scholar 

  16. Bukrinsky MI, Sharova N, Dempsey MP, Stanwick TL, Bukrinskaya AG, Haggerty S, Stevenson M: Active nuclear import of human immunodeficiency virus type 1 preintegration complexes. Proc Natl Acad Sci USA. 1992, 89: 6580-6584.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  17. Chiu YL, Soros VB, Kreisberg JF, Stopak K, Yonemoto W, Greene WC: Cellular APOBEC3G restricts HIV-1 infection in resting CD4+ T cells. Nature. 2005, 435: 108-114. 10.1038/nature03493.

    Article  CAS  PubMed  Google Scholar 

  18. Stremlau M, Owens CM, Perron MJ, Kiessling M, Autissier P, Sodroski J: The cytoplasmic body component TRIM5alpha restricts HIV-1 infection in Old World monkeys. Nature. 2004, 427: 848-853. 10.1038/nature02343.

    Article  CAS  PubMed  Google Scholar 

  19. Bukrinsky MI, Stanwick TL, Dempsey MP, Stevenson M: Quiescent T lymphocytes as an inducible virus reservoir in HIV-1 infection. Science. 1991, 254: 423-427.

    Article  CAS  PubMed  Google Scholar 

  20. Blankson JN, Persaud D, Siliciano RF: The challenge of viral reservoirs in HIV-1 infection. Annu Rev Med. 2002, 53: 557-593. 10.1146/annurev.med.53.082901.104024.

    Article  CAS  PubMed  Google Scholar 

  21. Finzi D, Hermankova M, Pierson T, Carruth LM, Buck C, Chaisson RE, Quinn TC, 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. 10.1126/science.278.5341.1295.

    Article  CAS  PubMed  Google Scholar 

  22. Ganesh L, Burstein E, Guha-Niyogi A, Louder MK, Mascola JR, Klomp LW, Wijmenga C, Duckett CS, Nabel GJ: The gene product Murr1 restricts HIV-1 replication in resting CD4+ lymphocytes. Nature. 2003, 426: 853-857. 10.1038/nature02171.

    Article  CAS  PubMed  Google Scholar 

  23. Moris A, Nobile C, Buseyne F, Porrot F, Abastado JP, Schwartz O: DC-SIGN promotes exogenous MHC-I-restricted HIV-1 antigen presentation. Blood. 2004, 103: 2648-2654. 10.1182/blood-2003-07-2532.

    Article  CAS  PubMed  Google Scholar 

  24. Pelchen-Matthews A, Kramer B, Marsh M: Infectious HIV-1 assembles in late endosomes in primary macrophages. J Cell Biol. 2003, 162: 443-455. 10.1083/jcb.200304008.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  25. Sharova N, Swingler C, Sharkey M, Stevenson M: Macrophages archive HIV-1 virions for dissemination in trans. Embo J. 2005, 24: 2481-2489. 10.1038/sj.emboj.7600707.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  26. Folks TM, Justement J, Kinter A, Schnittman S, Orenstein J, Poli G, Fauci AS: Characterization of a promonocyte clone chronically infected with HIV and inducible by 13-phorbol-12-myristate acetate. J Immunol. 1988, 140: 1117-1122.

    CAS  PubMed  Google Scholar 

  27. Emiliani S, Fischle W, Ott M, Van Lint C, Amelia CA, Verdin E: Mutations in the tat gene are responsible for human immunodeficiency virus type 1 postintegration latency in the U1 cell line. J Virol. 1998, 72: 1666-1670.

    PubMed Central  CAS  PubMed  Google Scholar 

  28. Jordan A, Defechereux P, Verdin E: The site of HIV-1 integration in the human genome determines basal transcriptional activity and response to Tat transactivation. Embo J. 2001, 20: 1726-1738. 10.1093/emboj/20.7.1726.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  29. Kutsch O, Levy DN, Bates PJ, Decker J, Kosloff BR, Shaw GM, Priebe W, Benveniste EN: Bis-anthracycline antibiotics inhibit human immunodeficiency virus type 1 transcription. Antimicrob Agents Chemother. 2004, 48: 1652-1663. 10.1128/AAC.48.5.1652-1663.2004.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  30. Brooks DG, Hamer DH, Arlen PA, Gao L, Bristol G, Kitchen CM, Berger EA, Zack JA: Molecular characterization, reactivation, and depletion of latent HIV. Immunity. 2003, 19: 413-423. 10.1016/S1074-7613(03)00236-X.

    Article  CAS  PubMed  Google Scholar 

  31. Shen A, Zink MC, Mankowski JL, Chadwick K, Margolick JB, Carruth LM, Li M, Clements JE, Siliciano RF: Resting CD4+ T lymphocytes but not thymocytes provide a latent viral reservoir in a simian immunodeficiency virus-Macaca nemestrina model of human immunodeficiency virus type 1-infected patients on highly active antiretroviral therapy. J Virol. 2003, 77: 4938-4949. 10.1128/JVI.77.8.4938-4949.2003.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  32. Pereira LA, Bentley K, Peeters A, Churchill MJ, Deacon NJ: compilation of cellular transcription factor interactions with the HIV-1 LTR promoter. Nucleic Acids Res. 2000, 28: 663-668. 10.1093/nar/28.3.663.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  33. Nabel G, Baltimore D: An inducible transcription factor activates expression of human immunodeficiency virus in T cells. Nature. 1987, 326: 711-713. 10.1038/326711a0.

    Article  CAS  PubMed  Google Scholar 

  34. Berkhout B, Silverman RH, Jeang KT: Tat trans-activates the human immunodeficiency virus through a nascent RNA target. Cell. 1989, 59: 273-282. 10.1016/0092-8674(89)90289-4.

    Article  CAS  PubMed  Google Scholar 

  35. Wei P, Garber ME, Fang SM, Fischer WH, Jones KA: 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. 10.1016/S0092-8674(00)80939-3.

    Article  CAS  PubMed  Google Scholar 

  36. Isel C, Karn J: Direct evidence that HIV-1 Tat stimulates RNA polymerase II carboxyl-terminal domain hyperphosphorylation during transcriptional elongation. J Mol Biol. 1999, 290: 929-941. 10.1006/jmbi.1999.2933.

    Article  CAS  PubMed  Google Scholar 

  37. Marcello A, Zoppe M, Giacca M: Multiple modes of transcriptional regulation by the HIV-1 Tat transactivator. IUBMB Life. 2001, 51: 175-181.

    Article  CAS  PubMed  Google Scholar 

  38. Benkirane M, Chun RF, Xiao H, Ogryzko VV, Howard BH, Nakatani Y, Jeang KT: Activation of integrated provirus requires histone acetyltransferase. p300 and P/CAF are coactivators for HIV-1 Tat. J Biol Chem. 1998, 273: 24898-24905. 10.1074/jbc.273.38.24898.

    Article  CAS  PubMed  Google Scholar 

  39. Hottiger MO, Nabel GJ: Interaction of human immunodeficiency virus type 1 Tat with the transcriptional coactivators p300 and CREB binding protein. J Virol. 1998, 72: 8252-8256.

    PubMed Central  CAS  PubMed  Google Scholar 

  40. Marzio G, Tyagi M, Gutierrez MI, Giacca M: HIV-1 tat transactivator recruits p300 and CREB-binding protein histone acetyltransferases to the viral promoter. Proc Natl Acad Sci USA. 1998, 95: 13519-13524. 10.1073/pnas.95.23.13519.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  41. 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. 10.1093/emboj/cdg631.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  42. Bres V, Tagami H, Peloponese JM, Loret E, Jeang KT, Nakatani Y, Emiliani S, Benkirane M, Kiernan RE: Differential acetylation of Tat coordinates its interaction with the co-activators cyclin Tl and PCAF. Embo J. 2002, 21: 6811-6819. 10.1093/emboj/cdf669.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  43. Kiernan RE, Vanhulle C, Schiltz L, Adam E, Xiao H, Maudoux F, Calomme C, Burny A, Nakatani Y, Jeang KT, et al: HIV-1 tat transcriptional activity is regulated by acetylation. Embo J. 1999, 18: 6106-6118. 10.1093/emboj/18.21.6106.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  44. Pagans S, Pedal A, North BJ, Kaehlcke K, Marshall BL, Dorr A, Hetzer-Egger C, Henklein P, Frye R, McBurney MW, et al: SIRT1 regulates HIV transcription via Tat deacetylation. PLoS Biol. 2005, 3: e41-10.1371/journal.pbio.0030041.

    Article  PubMed Central  PubMed  Google Scholar 

  45. Ott M, Schnolzer M, Garnica J, Fischle W, Emiliani S, Rackwitz HR, Verdin E: Acetylation of the HIV-1 Tat protein by p300 is important for its transcriptional activity. Curr Biol. 1999, 9: 1489-1492. 10.1016/S0960-9822(00)80120-7.

    Article  CAS  PubMed  Google Scholar 

  46. Kaehlcke K, Dorr A, Hetzer-Egger C, Kiermer V, Henklein P, Schnoelzer M, Loret E, Cole PA, Verdin E, Ott M: Acetylation of Tat defines a cyclinT1-independent step in HIV transactivation. Mol Cell. 2003, 12: 167-176. 10.1016/S1097-2765(03)00245-4.

    Article  PubMed  Google Scholar 

  47. Dorr A, Kiermer V, Pedal A, Rackwitz HR, Henklein P, Schubert U, Zhou MM, Verdin E, Ott M: Transcriptional synergy between Tat and PCAF is dependent on the binding of acetylated Tat to the PCAF bromodomain. Embo J. 2002, 21: 2715-2723. 10.1093/emboj/21.11.2715.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  48. Bres V, Kiernan RE, Linares LK, Chable-Bessia C, Plechakova O, Treand C, Emiliani S, Peloponese JM, Jeang KT, Coux O, et al: A non-proteolytic role for ubiquitin in Tat-mediated transactivation of the HIV-1 promoter. Nat Cell Biol. 2003, 5: 754-761. 10.1038/ncb1023.

    Article  CAS  PubMed  Google Scholar 

  49. Ghose R, Liou LY, Herrmann CH, Rice AP: Induction of TAK (cyclin Tl/P-TEFb) in purified resting CD4(+) T lymphocytes by combination of cytokines. J Virol. 2001, 75: 11336-11343. 10.1128/JVI.75.23.11336-11343.2001.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  50. Kao SY, Caiman AF, Luciw PA, Peterlin BM: Anti-termination of transcription within the long terminal repeat of HIV-1 by tat gene product. Nature. 1987, 330: 489-493. 10.1038/330489a0.

    Article  CAS  PubMed  Google Scholar 

  51. Adams M, Sharmeen L, Kimpton J, Romeo JM, Garcia JV, Peterlin BM, Groudine M, Emerman M: Cellular latency in human immunodeficiency virus-infected individuals with high CD4 levels can be detected by the presence of promoter-proximal transcripts. Proc Natl Acad Sci USA. 1994, 91: 3862-3866.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  52. Pomerantz RJ, Seshamma T, Trono D: Efficient replication of human immunodeficiency virus type 1 requires a threshold level of Rev: potential implications for latency. J Virol. 1992, 66: 1809-1813.

    PubMed Central  CAS  PubMed  Google Scholar 

  53. Weinberger LS, Burnett JC, Toettcher JE, Arkin AP, Schaffer DV: Stochastic Gene Expression in a Lentiviral Positive-Feedback Loop: HIV-1 Tat Fluctuations Drive Phenotypic Diversity. Cell. 2005, 122: 169-182. 10.1016/j.cell.2005.06.006.

    Article  CAS  PubMed  Google Scholar 

  54. Han Y, Lassen K, Monie D, Sedaghat AR, Shimoji S, Liu X, Pierson TC, Margolick JB, Siliciano RF, Siliciano JD: Resting CD4+ T cells from human immunodeficiency virus type 1 (HlV-l)-infected individuals carry integrated HIV-1 genomes within actively transcribed host genes. J Virol. 2004, 78: 6122-6133. 10.1128/JVI.78.12.6122-6133.2004.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  55. Schroder AR, Shinn P, Chen H, Berry C, Ecker JR, Bushman F: HIV-1 integration in the human genome favors active genes and local hotspots. Cell. 2002, 110: 521-529. 10.1016/S0092-8674(02)00864-4.

    Article  CAS  PubMed  Google Scholar 

  56. Jordan A, Bisgrove D, Verdin E: HIV reproducibly establishes a latent infection after acute infection of T cells in vitro. Embo J. 2003, 22: 1868-1877. 10.1093/emboj/cdg188.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  57. Lewinski MK, Bisgrove D, Shinn P, Chen H, Hoffmann C, Hannenhalli S, Verdin E, Berry CC, Ecker JR, Bushman FD: Genome-wide analysis of chromosomal features repressing human immunodeficiency virus transcription. J Virol. 2005, 79: 6610-6619. 10.1128/JVI.79.11.6610-6619.2005.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  58. Misteli T: Spatial positioning; a new dimension in genome function. Cell. 2004, 119: 153-156. 10.1016/j.cell.2004.09.035.

    Article  CAS  PubMed  Google Scholar 

  59. Spilianakis CG, Lalioti MD, Town T, Lee GR, Flavell RA: Interchromosomal associations between alternatively expressed loci. Nature. 2005, 435: 637-645. 10.1038/nature03574.

    Article  CAS  PubMed  Google Scholar 

  60. Kim SH, McQueen PG, Lichtman MK, Shevach EM, Parada LA, Misteli T: Spatial genome organization during T-cell differentiation. Cytogenet Genome Res. 2004, 105: 292-301. 10.1159/000078201.

    Article  CAS  PubMed  Google Scholar 

  61. Brown KE, Baxter J, Graf D, Merkenschlager M, Fisher AG: Dynamic repositioning of genes in the nucleus of lymphocytes preparing for cell division. Mol Cell. 1999, 3: 207-217. 10.1016/S1097-2765(00)80311-1.

    Article  CAS  PubMed  Google Scholar 

  62. Gilbert N, Boyle S, Fiegler H, Woodfme K, Carter NP, Bickmore WA: Chromatin architecture of the human genome: gene-rich domains are enriched in open chromatin fibers. Cell. 2004, 118: 555-566. 10.1016/j.cell.2004.08.011.

    Article  CAS  PubMed  Google Scholar 

  63. Kalpana GV, Marmon S, Wang W, Crabtree GR, Goff SP: Binding and stimulation of HIV-1 integrase by a human homolog of yeast transcription factor SNF5. Science. 1994, 266: 2002-2006.

    Article  CAS  PubMed  Google Scholar 

  64. 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. 10.1074/jbc.M209278200.

    Article  CAS  PubMed  Google Scholar 

  65. Lin CW, Engelman A: The barrier-to-autointegration factor is a component of functional human immunodeficiency virus type 1 preintegration complexes. J Virol. 2003, 77: 5030-5036. 10.1128/JVI.77.8.5030-5036.2003.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  66. Segura-Totten M, Wilson KL: BAF: roles in chromatin, nuclear structure and retrovirus integration. Trends Cell Biol. 2004, 14: 261-266. 10.1016/j.tcb.2004.03.004.

    Article  CAS  PubMed  Google Scholar 

  67. 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. 10.1016/S1097-2765(01)00255-6.

    Article  CAS  PubMed  Google Scholar 

  68. Marcello A, Lusic M, Pegoraro G, Pellegrini V, Beltram F, Giacca M: Nuclear organization and the control of HIV-1 transcription. Gene. 2004, 326: 1-11. 10.1016/j.gene.2003.10.018.

    Article  CAS  PubMed  Google Scholar 

  69. Cereseto A, Manganaro L, Gutierrez MI, Terreni M, Fittipaldi A, Lusic M, Marcello A, Giacca M: Acetylation of HIV-1 integrase by p300 regulates viral integration. Embo J. 2005

    Google Scholar 

  70. Fire A, Xu S, Montgomery MK, Kostas SA, Driver SE, Mello CC: Potent and specific genetic interference by double-stranded RNA in Caenorhabditis elegans. Nature. 1998, 391: 806-811. 10.1038/35888.

    Article  CAS  PubMed  Google Scholar 

  71. Hamilton AJ, Baulcombe DC: species of small antisense RNA in posttranscriptional gene silencing in plants. Science. 1999, 286: 950-952. 10.1126/science.286.5441.950.

    Article  CAS  PubMed  Google Scholar 

  72. Hammond SM, Bernstein E, Beach D, Harmon GJ: An RNA-directed nuclease mediates post-transcriptional gene silencing in Drosophila cells. Nature. 2000, 404: 293-296. 10.1038/35005107.

    Article  CAS  PubMed  Google Scholar 

  73. Bartel DP: MicroRNAs: genomics, biogenesis, mechanism, and function. Cell. 2004, 116: 281-297. 10.1016/S0092-8674(04)00045-5.

    Article  CAS  PubMed  Google Scholar 

  74. Griffiths-Jones S: The microRNA Registry. Nucleic Acids Res. 2004, 32: D109-111. 10.1093/nar/gkh023.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  75. Lippman Z, Martienssen R: The role of RNA interference in heterochromatic silencing. Nature. 2004, 431: 364-370. 10.1038/nature02875.

    Article  CAS  PubMed  Google Scholar 

  76. Mak J: RNA interference: more than a research tool in the vertebrates' adaptive immunity. Retrovirology. 2005, 2: 35-10.1186/1742-4690-2-35.

    Article  PubMed Central  PubMed  Google Scholar 

  77. Volpe TA, Kidner C, Hall IM, Teng G, Grewal SI, Martienssen RA: Regulation of heterochromatic silencing and histone H3 lysine-9 methylation by RNAi. Science. 2002, 297: 1833-1837. 10.1126/science.1074973.

    Article  CAS  PubMed  Google Scholar 

  78. Jacque JM, Triques K, Stevenson M: Modulation of HIV-1 replication by RNA interference. Nature. 2002, 418: 435-438. 10.1038/nature00896.

    Article  CAS  PubMed  Google Scholar 

  79. Novina CD, Murray MF, Dykxhoorn DM, Beresford PJ, Riess J, Lee SK, Collman RG, Lieberman J, Shankar P, Sharp PA: siRNA-directed inhibition of HIV-1 infection. Nat Med. 2002, 8: 681-686.

    CAS  PubMed  Google Scholar 

  80. Das AT, Brummelkamp TR, Westerhout EM, Vink M, Madiredjo M, Bernards R, Berkhout B: Human immunodeficiency virus type 1 escapes from RNA interference-mediated inhibition. J Virol. 2004, 78: 2601-2605. 10.1128/JVI.78.5.2601-2605.2004.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  81. Sullivan CS, Ganem D: MicroRNAs and viral infection. Mol Cell. 2005, 20: 3-7. 10.1016/j.molcel.2005.09.012.

    Article  CAS  PubMed  Google Scholar 

  82. Pfeffer S, Zavolan M, Grasser FA, Chien M, Russo JJ, Ju J, John B, Enright AJ, Marks D, Sander C, Tuschl T: Identification of virus-encoded microRNAs. Science. 2004, 304: 734-736. 10.1126/science.1096781.

    Article  CAS  PubMed  Google Scholar 

  83. Cai X, Lu S, Zhang Z, Gonzalez CM, Damania B, Cullen BR: Kaposi's sarcoma-associated herpesvirus expresses an array of viral microRNAs in latently infected cells. Proc Natl Acad Sci USA. 2005, 102: 5570-5575. 10.1073/pnas.0408192102.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  84. Bennasser Y, Le SY, Benkirane M, Jeang KT: Evidence that HIV-1 Encodes an siRNA and a Suppressor of RNA Silencing. Immunity. 2005, 22: 607-619. 10.1016/j.immuni.2005.03.010.

    Article  CAS  PubMed  Google Scholar 

  85. Lecellier CH, Dunoyer P, Arar K, Lehmann-Che J, Eyquem S, Himber C, Saib A, Voinnet O: A cellular microRNA mediates antiviral defense in human cells. Science. 2005, 308: 557-560. 10.1126/science.1108784.

    Article  CAS  PubMed  Google Scholar 

  86. Omoto S, Ito M, Tsutsumi Y, Ichikawa Y, Okuyama H, Brisibe EA, Saksena NK, Fujii YR: HIV-1 nef suppression by virally encoded microRNA. Retrovirology. 2004, 1: 44-10.1186/1742-4690-1-44.

    Article  PubMed Central  PubMed  Google Scholar 

  87. Pion M, Jordan A, Biancotto A, Dequiedt F, Gondois-Rey F, Rondeau S, Vigne R, Hejnar J, Verdin E, Hirsch I: Transcriptional suppression of in vitro-integrated human immunodeficiency virus type 1 does not correlate with proviral DNA methylation. J Virol. 2003, 77: 4025-4032. 10.1128/JVI.77.7.4025-4032.2003.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  88. Bednarik DP, Cook JA, Pitha PM: Inactivation of the HIV LTR by DNA CpG methylation: evidence for a role in latency. Embo J. 1990, 9: 1157-1164.

    PubMed Central  CAS  PubMed  Google Scholar 

  89. Finzi D, Blankson J, Siliciano JD, Margolick JB, Chadwick K, Pierson T, Smith K, Lisziewicz J, Lori F, Flexner C, et al: Latent infection of CD4+ T cells provides a mechanism for lifelong persistence of HIV-1, even in patients on effective combination therapy. Nat Med. 1999, 5: 512-517. 10.1038/8394.

    Article  CAS  PubMed  Google Scholar 

  90. Siliciano JD, Kajdas J, Finzi D, Quinn TC, Chadwick K, Margolick JB, Kovacs C, Gange SJ, Siliciano RF: 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. 10.1038/nm880.

    Article  CAS  PubMed  Google Scholar 

  91. Davey RT, Bhat N, Yoder C, Chun TW, Metcalf JA, Dewar R, Natarajan V, Lempicki RA, Adelsberger JW, Miller KD, 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. 10.1073/pnas.96.26.15109.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  92. Ruff CT, Ray SC, Kwon P, Zinn R, Pendleton A, Mutton N, Ashworth R, Gange S, Quinn TC, Siliciano RF, Persaud D: Persistence of wild-type virus and lack of temporal structure in the latent reservoir for human immunodeficiency virus type 1 in pediatric patients with extensive antiretroviral exposure. J Virol. 2002, 76: 9481-9492. 10.1128/JVI.76.18.9481-9492.2002.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  93. Chun TW, Engel D, Mizell SB, Hallahan CW, Fischette M, Park S, Davey RT, Dybul M, Kovacs JA, Metcalf JA, et al: Effect of interleukin-2 on the pool of latently infected, resting CD4+ T cells in HIV-1-infected patients receiving highly active anti-retro viral therapy. Nat Med. 1999, 5: 651-655. 10.1038/9498.

    Article  CAS  PubMed  Google Scholar 

  94. Kulkosky J, Culnan DM, Roman J, Dornadula G, Schnell M, Boyd MR, Pomerantz RJ: Prostratin: activation of latent HIV-1 expression suggests a potential inductive adjuvant therapy for HAART. Blood. 2001, 98: 3006-3015. 10.1182/blood.V98.10.3006.

    Article  CAS  PubMed  Google Scholar 

  95. Scripture-Adams DD, Brooks DG, Korin YD, Zack JA: Interleukin-7 induces expression of latent human immunodeficiency virus type 1 with minimal effects on T-cell phenotype. J Virol. 2002, 76: 13077-13082. 10.1128/JVI.76.24.13077-13082.2002.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  96. Korin YD, Brooks DG, Brown S, Korotzer A, Zack JA: Effects of prostratin on T-cell activation and human immunodeficiency virus latency. J Virol. 2002, 76: 8118-8123. 10.1128/JVI.76.16.8118-8123.2002.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  97. Ylisastigui L, Archin NM, Lehrman G, Bosch RJ, Margolis DM: Coaxing HIV-1 from resting CD4 T cells: histone deacetylase inhibition allows latent viral expression. Aids. 2004, 18: 1101-1108. 10.1097/00002030-200405210-00003.

    Article  CAS  PubMed  Google Scholar 

  98. Saavedra-Lozano J, Cao Y, Callison J, Sarode R, Sodora D, Edgar J, Hatfield J, Picker L, Peterson D, Ramilo O, Vitetta ES: An anti-CD45RO immunotoxin kills HIV-latently infected cells from individuals on HAART with little effect on CD8 memory. Proc Natl Acad Sci USA. 2004, 101: 2494-2499. 10.1073/pnas.0308381100.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

I wish to acknowledge the support of the EC STREP consortium n. 012182 "Challenging the hidden HIV: understanding the block on transcriptional reactivation to eradicate infection" that was established to address many of the issues reviewed in this review. Other sources of funding of my work include the Human Frontiers Science Program, the Istituto Superiore di Sanità of Italy and the Ministero Istruzione Università e Ricerca of Italy.

I thank Gianluca Pegoraro and Marina Lusic for critically reading the manuscript as well as Ben Berkhout for helpful suggestions.

I'm particularly grateful to Mauro Giacca for his excellent scientific mentoring.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Alessandro Marcello.

Additional information

Competing interests

The author declares that he has no competing interests.

Authors’ original submitted files for images

Below are the links to the authors’ original submitted files for images.

Authors’ original file for figure 1

Rights and permissions

Open Access This article is published under license to BioMed Central Ltd. This is an Open Access article is distributed under the terms of the Creative Commons Attribution License ( https://creativecommons.org/licenses/by/2.0 ), which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.

Reprints and permissions

About this article

Cite this article

Marcello, A. Latency: the hidden HIV-1 challenge. Retrovirology 3, 7 (2006). https://doi.org/10.1186/1742-4690-3-7

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/1742-4690-3-7

Keywords