Skip to main content

Regulation of HIV-1 transcription in cells of the monocyte-macrophage lineage

Abstract

Human immunodeficiency virus type 1 (HIV-1) has been shown to replicate productively in cells of the monocyte-macrophage lineage, although replication occurs to a lesser extent than in infected T cells. As cells of the monocyte-macrophage lineage become differentiated and activated and subsequently travel to a variety of end organs, they become a source of infectious virus and secreted viral proteins and cellular products that likely initiate pathological consequences in a number of organ systems. During this process, alterations in a number of signaling pathways, including the level and functional properties of many cellular transcription factors, alter the course of HIV-1 long terminal repeat (LTR)-directed gene expression. This process ultimately results in events that contribute to the pathogenesis of HIV-1 infection. First, increased transcription leads to the upregulation of infectious virus production, and the increased production of viral proteins (gp120, Tat, Nef, and Vpr), which have additional activities as extracellular proteins. Increased viral production and the presence of toxic proteins lead to enhanced deregulation of cellular functions increasing the production of toxic cellular proteins and metabolites and the resulting organ-specific pathologic consequences such as neuroAIDS. This article reviews the structural and functional features of the cis-acting elements upstream and downstream of the transcriptional start site in the retroviral LTR. It also includes a discussion of the regulation of the retroviral LTR in the monocyte-macrophage lineage during virus infection of the bone marrow, the peripheral blood, the lymphoid tissues, and end organs such as the brain. The impact of genetic variation on LTR-directed transcription during the course of retrovirus disease is also reviewed.

Introduction

Approximately 33.2 million people are infected with the human immunodeficiency virus type 1 (HIV-1) worldwide, including 2.5 million people who were newly infected in 2007 [1]. Although fewer people are currently infected with HIV type 2 (HIV-2), this virus is spreading from its origin in West Africa to the Americas, Asia, and Europe [2] and reviewed in [3–5]). In addition to being the causative agent of the acquired immunodeficiency syndrome (AIDS), HIV-1 can cause neurological problems, ranging in severity from minor cognitive/motor dysfunction (MCMD) to HIV-1-associated dementia (HAD) (reviewed in [6–9]).

Cells of the monocyte-macrophage lineage play an important role in the transmission and pathogenesis of HIV [10–12]. When transmission occurs vaginally, rectally, or orally, the primary cells involved in the transmission event are dendritic cells [13]. However, during mucosal trauma, inflammation, and ulceration, the epithelial barrier may be disrupted and provide HIV with direct access to the mucosal microcirculation and/or provide directional signals to recruit highly susceptible, activated, inflammatory monocytes and T cells [14]. Circulating monocytes can also be infected and then migrate to peripheral tissues, including the brain [15, 16], lung [17], lymphatic system [18], bone marrow [19, 20], and kidney (reviewed in [21]). Infected monocytes can differentiate into monocyte-derived macrophages (MDMs) and may form a long-lived reservoir for the virus [22–25]. Additionally, MDMs can be infected after differentiation and are more susceptible to new infection in comparison to freshly isolated monocytes due to increased expression of the HIV co-receptor CCR5 [26]; however, this infection is limited, and the production of virus is hindered at many steps which will be discussed. Infected MDMs can seed the periphery with new infectious virus [20], directly transmit virus to T cells [27, 28], release toxic viral proteins [29–31], and produce an altered array of cytokines and effector functions that contribute to HIV pathogenesis [32–35]. Additionally, infected monocyte progenitor cells can harbor virus in the bone marrow and seed the periphery with infected daughter cells. As these cells differentiate in the marrow and periphery, the levels of HIV-1 transcription may increase, resulting in the expression of toxic viral proteins and enhanced replication [36] and Alexaki, Shah, and Wigdahl, unpublished results). These cells can also cross the blood-brain barrier and deliver virus to the central nervous system.

Retroviral gene expression is regulated in a cell type- and differentiation-dependent manner by the binding of both host and viral proteins to the long terminal repeat (LTR), which serves as the viral promoter (reviewed in [37]). Host transcription factors such as the Sp family, nuclear factor kappa B (NF-κB) family, activator protein 1 (AP-1) proteins, nuclear factor of activated T cells (NFAT), and CCAAT enhancer binding protein (C/EBP) family members play key roles in the regulation of retroviral transcription by binding sites in the LTR that display different levels of sequence conservation. Viral proteins such as HIV Vpr and Tat also bind to the LTR to regulate transcription. Many of these host and viral proteins engage in extensive protein-protein interactions, leading to a complex system of transcriptional regulation. Adding to this complexity, the genomes of HIV-1, HIV-2, and simian immunodeficiency virus (SIV) accumulate a significant spectrum of genetic alterations as the virus replicates. When present in the LTR, these sequence alterations affect the ability of host and viral proteins to bind to their cognate binding sites and result in altered transcriptional and replication potential of the virus [38–46].

Regulation of HIV-1 transcription in cells of the monocyte-macrophage lineage varies considerably with the differentiation stage of the cell. Specifically, it has been observed that cyclin T1 expression in monocytes is controlled by differentiation. Cyclin T1 increases as cells of the monocyte-macrophage lineage differentiate [47]. This is important because cyclin T1 is one-half of the positive transcriptional elongation factor b (P-TEFb) complex necessary for the binding of Tat to TAR for the induction of HIV-1 transcription. Unstimulated peripheral monocytes and myeloid progenitor cells support low levels of viral replication and transcription in response to cellular activation [27, 36, 48–54], whereas differentiated MDMs have increased viral replication but either do not respond to [45] or downregulate HIV transcription [48, 55] in response to cellular stimulation. During late-stage disease and AIDS, when CD4+ T cells have largely been depleted, HIV-1-infected MDMs represent a greater component of the total infected cell population, and this pool of virus contributes significantly to the circulating levels of virus in vivo [56, 57].

Lentiviral LTR Structure

Lentiviral LTRs are comprised of U5, R, and U3 regions. The U3 region is further divided into the core promoter, enhancer, and modulatory regions [37]. Lentiviral LTRs, HIV-1, SIV, and HIV-2, have closely related core promoters (Sp binding sites) and enhancer regions (NF-κB binding sites) (Fig. 1). These cis-acting elements allow for efficient replication in a variety of cell types and conditions that result in differential availability and activation state of transcription factors in the nucleus. However, the modulatory region is less closely related between lentiviral LTRs and contributes to the ability of the LTR to regulate transcription in various cell types and under various cellular conditions. These concepts are discussed below.

Figure 1
figure 1

Structure of retroviral LTRs. Retroviral LTRs are divided into the U3, R, and U5 regions, and the U3 region is further divided into the Modulatory, Enhancer (E) and Promoter regions (top bars). HIV-1, HIV-2, and SIV all contain highly conserved promoters containing TATA boxes (yellow) and Sp factor binding sites (red) and enhancers (labeled E in light blue bar) containing NF-κB binding sites (blue). The R region of each contains a trans-acting responsive element (TAR) (orange) that forms an RNA stem loop structure upon transcription that binds to the viral protein Tat. A negative regulatory element (NRE, pink) was identified that was subsequently shown to serve as both activator and repressor by binding NFAT proteins (dark blue), AP-1 proteins (purple), and C/EBP factors (green). The modulatory regions of SIVmac and HIV-2 also contain purine box arrays (PuB, gold) and sites that bind members of the Ets family (teal).

Core promoter and enhancer regions: the interaction of Sp, NF-κB, and NFAT proteins

Sp factors

The core promoters of HIV-1, HIV-2, and SIV all contain a TATA box and multiple binding sites for the Sp family of transcription factors, and their enhancers all contain at least one binding site for NF-κB. The Sp and NF-κB factor binding sites in the core promoter play important cell type-specific roles in regulating transcription and replication. The promoter of HIV-1 contains three binding sites for Sp factors at -46 to -78 relative to the transcriptional start site (Fig. 1) [58]. Sp factors also regulate transcription by binding to positions +271 to +289 [59, 60] and -421 to -451 [61] relative to the transcriptional start site. Sp family members include Sp1-4, as well as M1 and M2, truncated Sp3 proteins that result from alternative translational start sites within the transactivation domain [62–65]. All of the Sp proteins contain zinc finger DNA binding domains, and Sp1, 3, and 4 have similar, though not identical, affinities and specificities for GC-rich (GGGGCGGGGC) DNA [62, 66, 67]. Sp2 binds to GT-rich sequences (GGTGTGGGG) rather than to the GC-rich sequences that constitute the classical Sp binding sites [65]. Sp1 and Sp4 are transcriptional activators, whereas Sp3 has been classified as a repressor of HIV-1 transcription. By itself, Sp3 can weakly activate HIV-1 transcription; however, in the presence of the strong activator Sp1, it competes for binding to the LTR and inhibits activation by Sp1 [66, 68, 69]. In contrast, M1 and M2 have the Sp3 DNA binding domain but lack the transactivation domain and are true repressors of transcription in the absence or presence of other Sp family members [69]. In addition to repressing Sp-mediated transactivation, Sp3 represses LTR activation by the viral protein Tat [66]. Sp4 is expressed predominantly in the brain [62, 70, 71], providing an additional HIV-1 LTR transactivator to drive replication in this compartment. Unlike Sp1, Sp4 does not synergistically activate transcription in the presence of multiple Sp binding sites [71]. Consequently, the loss of one binding site due to genetic variation may have less of an effect in the brain than it would in other tissues, because the loss of function would not synergistically disrupt binding.

Genetic variation within the Sp sites is likely to play a role in HIV-1-associated disease progression. The NF-κB-proximal Sp site (site III) is much less conserved during the course of disease than Sp sites I and II [41] and Kilareski and Wigdahl, unpublished results). A C-to-T change at position 5 of Sp site III has been shown to correlate positively with HIV-1-associated disease progression, both in the periphery and in the brain [41]. This variant greatly reduces the affinity of this site for Sp factors, but greatly increases the response of viral replication to tumor necrosis factor α (TNFα) stimulation in peripheral blood mononuclear cells (Kilareski, Pirrone, and Wigdahl, unpublished observation). This finding is likely due to a loss of steric hindrance leading to an increase in NF-κB binding to its adjacent binding sites (Liu, Banerjee, and Wigdahl, unpublished observations). In the presence of Sp4 in the brain, one could speculate that this effect may be magnified, because Sp4 binding to sites I and II is not affected by the loss of Sp binding to site III, and the resulting stimulated LTR may have high levels of both Sp and NF-κB factors bound to their cognate sites.

Sp factors bound to the HIV-1 core promoter cooperate with the TATA binding protein and TATA binding protein-associated factors 110 and 55 to drive basal transcription [72–75]. They can also recruit P-TEFb to promote phosphorylation of RNA Pol II [76] and play an important role in remodeling chromatin to facilitate or inhibit transcription [77, 78]. Histone deacetylases (HDACs) 1 and 2 are regulated through phosphorylation by protein kinase CK2. Sp1 and Sp3 can bind and recruit the phosphorylated HDACs to the LTR to repress LTR activity [79–81]. The repressor activity of Sp1 and Sp3 is regulated by the expression of CK2 [77].

The three Sp sites in the HIV-1 promoter have different affinities for Sp factors [39, 40, 58, 82], and the affinity of Sp for LTR binding sites correlates with replication kinetics; faster viral replication is achieved when a higher affinity Sp binding site is in the NF-κB proximal site [39]. Interestingly, this might, at first glance, seem to contradict the fact presented above that Sp site III has increased genetic variation with the 5T variant (a low binding affinity site) correlating with disease progression, given traditionally low binding affinity correlates with decreased viral production. However, given that a decreased binding affinity has been shown to promote higher levels of NF-κB binding, this variation may actually provide an opportunity for increased replication (Kilareski and Wigdahl, unpublished observations). This suggests that genetic variations within these sites could have significant effects on the overall viral replication kinetics [41].

Expression patterns of the different Sp isoforms can modulate HIV-1 transcription in different cell types. As cells of the monocyte lineage differentiate, the ratio of Sp1 to Sp3 increases, resulting in increased HIV-1 transcription (McAllister and Wigdahl, unpublished observations). This process allows HIV to replicate at low levels, if at all, in circulating monocytes, and to evade the immune system until the cells are differentiated in peripheral tissues. The importance of the Sp sites also varies depending on the differentiation stage of the cell; in unstimulated monocytes, mutation of the Sp sites reduces LTR activity, whereas in MDMs, transcription of HIV and replication of SIVmac are abolished when these critical binding sites are knocked out [83–86].

DNA binding and transactivation activity of Sp factors are regulated both positively and negatively by phosphorylation and other post-translational modifications (reviewed in [87, 88] and Fig. 2). Phosphorylation at Sp1 Ser131 by DNA-dependent protein kinase increases the affinity of the protein for DNA and also increases the ability of the protein to cooperate with the viral protein Tat to transactivate the LTR [89–92]. In contrast, O-linked N-acetylglucosaminylation (O-GLcNAc) of Sp1 inhibits HIV-1 replication [93]. Therefore, modulating O-GLcNAc of transcription factors may play a role in regulation of HIV-1 latency and activation, and may link glucose metabolism to HIV-1 replication.

Figure 2
figure 2

Important Sp transcription factor signaling in monocyte-macropahges. (a) Activation of HIV transcription by the interaction of viral protein Tat with DNA-dependent protein kinase (DNA-PK) results in the subsequent phosphorylation at Ser131 of Sp1. Phosphorylated Sp1 results in increased transcription of proviral DNA, resulting in an increase in Tat production, perpetuating the cycle. (b) Inhibition of HIV transcription involves O-linked N-acetylglucosamine (O-GlcNAc) transferase (OGT) catalyzing the addition of O-GlcNAC to Sp proteins which blocks their interaction with their binding sites on the LTR, resulting in an inhibition/reduction in HIV transcription.

NF-κB

NF-κB proteins have been shown to be one of the main modulators of the HIV-1 LTR in all cell types and a potential pathway for anti-HIV-1 therapies [94]. NF-κB proteins bind the enhancer at two sites located at nucleotide positions -81 to -91 and -95 to -104 relative to the transcriptional start site [95–97]. NF-κB is composed of heterodimers of five c-rel protein family members: p65/RelA, NF-κB1/p50, c-Rel, RelB, and NF-κB2/p52. Functional NF-κB in T cells is predominantly composed of p65 or c-Rel bound to p50 or p52, whereas in MDMs, Rel B replaces p65 [97–100]. In T cells and immature monocytes, NF-κB shuttles between the cytoplasm and the nucleus in response to cellular stimuli. In the cytoplasm, NF-κB is bound to inhibitor (IκB) proteins [101]. As a result of specific stimuli, IκB is phosphorylated and released from NF-κB; after release from the inhibitory complex, NF-κB translocates to the nucleus where it activates many host and viral genes through the initial recruitment of P-TEFb (Fig. 3) [101–103]. Interestingly, one of the IκB's, IκBα has been shown to play a role in shuttling of NF-κB from the nucleus and cytosol and in the binding NF-κB in the nucleus of T cells, potentially contributing to the lower activation levels of the HIV-1 LTR and possibly promoting viral latency [104]. However, this mechanism has not been explored in cells of the monocyte-macrophage lineage. NF-κB can also function as a repressor of transcription through the recruitment of HDAC1 (Fig. 3) [78, 105].

Figure 3
figure 3

Important NF-κB transcription factor signaling in monocyte-macrophages. (a) Activation of HIV transcription: Translocation of NF-κB from the cytoplasm to the nucleus is controlled by association of IκB with the NF-κB hetero-/homo-dimer. Once IκB is phosphorylated, it relesases NF-κB which then translocates to the nucleus where it can bind the LTR and induce HIV transcription. (b) Inhibition of HIV transcription: In T cells, IκBα has been shown to contribute to lower levels of LTR transcription and potentially contribute to latency. It is postulated that a similar mechanism of action could be in place for cells of the monocyte-macrophage lineage. In addition, NF-κB's association with the histone deacetylase inhibitor HDAC1 results in constriction of the chromatin so that RNA polymerase does not have access to its target DNA.

NF-κB DNA binding activity first occurs in monocytes as they progress from promonocytes to monocytes; however, in mature monocytes and MDMs, NF-κB is constitutively active in the nucleus, and its DNA binding activity is not increased further in response to cellular activation or differentiation [106]. This constitutive pool of NF-κB allows a low level of basal HIV transcription in the absence of cellular stimuli. Binding of NF-κB to the enhancer of the HIV-1 LTR plays a critical role in the response of the LTR to cellular stimuli in both T cells and maturing monocytes [36, 94, 97, 106–109]. Deletion or mutation of the NF-κB sites abolishes LTR activity [97, 109–112] and results in reduced production of infectious virus [98]. Activation of monocytes by LPS, IL-6, or TNF-α (Fig. 3) results in enhanced HIV replication, a process that correlates with activation of NF-κB [27, 49–51, 113]. LPS activation of monocytes leads to the induction of the NF-κB pathway through TNF-α [27, 50]. In contrast, in differentiated primary MDMs, stimulation by LPS results in the downregulation of LTR activity and viral replication [48]. This activity was not affected by mutation of the NF-κB sites, but did map to the enhancer element (position -156 to -121); thus, this effect may involve NFAT proteins (see below) [48]. While this may seem counter-intuitive, one might speculate that stimulation of cells through the NF-κB pathway would enhance LTR activity and viral replication, it should be noted that LPS stimulation of differentiated macrophages could also induce transcription factors that negatively regulate the LTR, however this has not been explored. This would be very interesting as this might provide another reason for macrophages serving as a latent reservoir for HIV-1. In addition to activating transcription by binding the enhancer region, NF-κB activates transcription by binding to sites -1 to +9 and +31 to +40 relative to the transcriptional start site [114, 115].

The NF-κB site(s) located immediately upstream of the Sp sites in the enhancer in HIV and SIV result in Sp-NF-κB protein-protein interactions that further modulate the LTR activity. Sp1 and NF-κB proteins bind the LTR cooperatively and activate transcription synergistically in response to cellular stimulation [66, 82, 109]. This activation is mediated by the binding of the DNA-binding domain of p65 to the DNA-binding domain of Sp1 [108] (Fig. 4). Sp3 and Sp4 are unable to activate transcription cooperatively with NF-κB [66]. In the absence of functional Sp sites (or in the presence of genetic alterations that inactivate the Sp binding sites), binding of NF-κB to the enhancer can restore replication of the virus in T cells [116–118], perhaps by recruiting Sp to the variant sites.

Figure 4
figure 4

Important C/EBP transcription factor signaling in monocyte-macrophages: (a) Activation of HIV transcription: C/EBP, located in the cytoplasm of the cell, can become phosphorylated by the MAP kinase, PKA, or cdk9 through a variety of pathways. Once phosphorylated, C/EBP is translocated into the nucleus where it can transactivate the LTR. In addition, C/EBP associates with histone acetyl transferases such as p300, which when bound to the LTR, make the chromosome accessible for RNA polymerases to bind and transcribe the integrated proviral DNA. Finally, association of C/EBP with APOBEC3G allows for better reverse transcription in the cytoplasm. (b) Inhibition of HIV transcription: The binding of IFNβ to its receptor begins a JAK/STAT signaling cascade that results in increased production of C/EBP3 (LIP). C/EBP3, which does not contain the transactivation domain of full-length C/EBPs, does not interact with histone acetyl transferases and when bound to the LTR, blocks the binding of full-length C/EBPs, thereby leading to a repression of LTR activity.

NFAT (AP-3)

NFAT proteins are part of a family of Rel-related transcription factors that become active early after T cell activation and are constitutively in monocytes. NFAT exists as several isoforms (NFAT1, NFAT2/NFATc, and NFAT3-5) that activate a variety of genes in immune and non-immune cell populations [119–121]. Like NF-κB, NFAT contains a DNA-binding domain that is homologous to rel and shuttles between the cytoplasm and nucleus in response to cellular stimuli [122, 123]. In the cytoplasm, NFAT is dephosphorylated and translocates to the nucleus where it activates transcription of many genes [124–126] (Fig. 4). NFAT can bind DNA as a high affinity dimer or as a lower affinity monomer [127–129]. NFAT proteins frequently cooperate with other transcription factor families when bound to adjacent sites within a promoter.

An NFAT binding site was identified in the HIV-1 LTR at positions -216 and -254, with a footprint extending from -253 to -215 relative to the transcriptional start site [122, 130]. Although this site can bind NFAT in vitro, this site was later shown not to be necessary for NFAT-mediated activation of the HIV-1 LTR [131, 132]. Instead, NFAT binds the NF-κB binding sites in the enhancer in response to cellular activation in T cells and constitutively in monocytes [110, 112, 127, 130, 133]. NFAT activation of genes from κB-like sequences has been documented with a number of host and viral promoters [134, 135] (Fig. 4). In addition to binding to the enhancer, NFAT binding at positions +169 to +181 has been reported to activate transcription [59, 60, 136].

NFAT proteins activate HIV-1 transcription and replication in a variety of cell types. Whereas NFAT1 and NFAT2/NFATc are responsible for the activation of HIV in T cells [110, 133, 137] reviewed in [138, 139]), NFAT5, the most evolutionarily divergent NFAT member, regulates HIV replication in monocyte-MDMs [130]. Terminally differentiated MDMs constitutively express high levels of NFAT5, which is able to bind and activate the enhancer of HIV-1 subtypes B, C, and E, HIV-2, and SIV from multiple primate species [130]. Targeting NFAT5 with siRNAs in primary MDMs modestly reduces viral replication [130]; however, NFAT derived from MDM nuclear extract was unable to compete with NF-κB for binding to the HIV enhancer in vitro [98]. This finding suggests that in vivo, although constitutively expressed NFAT is able to bind the LTR, it is unable to do so in the presence of high levels of NF-κB.

Modulatory region

As its name implies, the modulatory region of the LTR functions to regulate transcription that is driven by the core and/or enhancer regions. A wide array of host and viral proteins bind the modulatory region of the LTR to either enhance or repress transcription [45, 46, 140]. In HIV-1, the loss of both the Sp and NF-κB sites effectively inactivates the LTR. In contrast, the modulatory region of SIVmac and HIV-2 have functional elements that are not present in HIV-1 that can compensate, at least in part, for the loss of the Sp and NF-κB sites [85]. Also, unlike the HIV-1 LTR, the 5' 364 bp of the 517 bp-long U3 region is dispensable for SIV replication [141–143]. Early reports investigating the role of the HIV-1 modulatory region identified bases -423 to -167 as a negative regulatory element (NRE) that repressed LTR activity [144]. Since then, this region has been shown to activate as well as to repress transcription (for review see [140]).

Basic leucine zipper transactivator proteins

C/EBPs, activating transcription factor/cyclic AMP response element binding (ATF/CREB) proteins, and AP-1 factors are members of a large family of basic leucine zipper (bZIP) proteins that play important roles in the regulation of retroviral transcription [145–147]. Dimerization of the bZIP family members occurs in the C-terminal α-helical leucine zipper domain and is necessary for binding to DNA (reviewed in [148, 149]). C/EBP, AP-1, and ATF/CREB proteins each have unique binding sites in the modulatory region of the HIV-1 LTR; however, heterodimerization between C/EBP and ATF/CREB or AP-1 family members has been shown to result in binding to sequences that are different from the consensus sequence for either family of factors [146, 150–155]. These sequences are often composed of half of the recognition sequence for each protein in the heterodimer [146, 150].

C/EBP

The HIV-1 LTR contains three C/EBP binding sites upstream of the transcriptional start site [156, 157] and one binding site downstream of the transcriptional start site, at the 3'-most end of U5 (Liu and Wigdahl, unpublished observations). C/EBPs play a critical role in HIV-1 replication. It has been shown that at least one upstream C/EBP binding site and the presence of C/EBP proteins are necessary for replication in cells of the monocyte-macrophage lineage [157–161]. The two C/EBP binding sites located in the U3 region of the LTR have differing affinities for C/EBP factors, with the upstream site (site II), having a much higher relative affinity than the downstream site (site I) [43]. In addition to activating HIV-1 transcription through direct binding to the LTR, C/EBP factors may inhibit the host cellular antiviral protein APOBEC3G (Fig. 5), allowing more efficient reverse transcription to occur in the cytoplasm [162].

Figure 5
figure 5

Regulation of HIV-1 transcription in circulating monocytes. Transcription of HIV-1 in circulating monocytes is dependent on the ratio of activator to repressor isoforms of transcription factors, the phosphorylation state of transcription factors, and the inducible translocation of NF-κB and NFAT factors from the cytoplasm. NF-κB can be induced to translocate to the nucleus by TNFα-mediated phosphorylation of IκB. NFAT is dephosphorylated in the cytoplasm by calcineurin, which acts in response to calcium levels within the cell. Once it is dephosphorylated, it translocates to the nucleus where it activates transcription by constitutively binding the NF-κB site in the enhancer. Phosphorylation plays a critical role in regulating the activity of C/EBP factors in monocytes. Phosphorylation of C/EBPα by ras-dependent mitogen-activated protein (MAP) kinase, signaled by IL-6 or by cAMP-dependent protein kinase A, results in its nuclear translocation and subsequent transactivation of the LTR. Cyclin-dependent kinase (cdk) 9 specifically phosphorylates C/EBPβ, which then translocates into the nucleus, binds to the LTR, and leads to an increase in HIV-1 gene expression. Once in the nucleus, C/EBP factors then regulate the activity of AP-1 factors. Relatively high levels of C/EBPα dimerize with AP-1 factors to form potent activators of transcription. Lower levels of C/EBPβ balance this activation by binding AP-1 leading to a loss in DNA binding affinity. Sp1 and Sp3 are constitutively expressed in the nucleus. In the presence of Sp1, which is a strong activator, Sp3 competes for binding to the LTR and inhibits activation by Sp1.

The C/EBP family of transcription factors consists of six members, including C/EBP α, β, γ, δ, ε, and ζ [163–169]. C/EBPβ itself has three isoforms that result from the use of internal start codons within a single mRNA [170, 171]. C/EBP-1, the full-length isoform, and C/EBP-2, an isoform that lacks the N-terminal 23 amino acids, both contain three transcriptional activation domains and function as activators of HIV-1 transcription. C/EBP-3, which lacks the N-terminal 198 amino acids that include the activation domains, serves as a repressor of HIV-1 transcription, because it retains the C-terminal DNA-binding domain and competes for binding with the activator isoforms of C/EBP.

C/EBP isoform expression depends on the differentiation and activation state of cells in the monocyte-macrophage lineage. C/EBPα levels are high early in monocyte differentiation and then decrease as cells mature, whereas C/EBPβ and C/EBPδ levels are low early in development and increase as cells mature [172, 173]. C/EBP isoform expression is also regulated by extracellular stimuli. C/EBPβ expression increases upon cellular activation, whereas expression of the other C/EBP isoforms remains constant [172, 174]. Exposure of macrophages to interleukin-1 (IL-1), tumor necrosis factor α (TNFα), or interferon-γ, all of which have been shown to be present at elevated levels during the course of HIV-1 infection, has been shown to induce a reduction in C/EBPα mRNA levels while the levels of C/EBPβ and C/EBPδ expression increase [174]. This results in C/EBPβ and C/EBPδ playing a key role in the regulation of HIV-1 transcription as disease progresses and inflammatory cytokine levels increase (Fig. 4).

An additional level of regulation of C/EBPβ activity resides in two regulatory domains that lie between the activation domains and the DNA binding domain. These domains inhibit C/EBP activity, until phosphorylation results in an increase in DNA binding affinity and transcriptional activation activity [175, 176]. Several signaling cascades regulate the phosphorylation state of C/EBP. Phosphorylation of threonine 235 by a ras-dependent mitogen-activated protein kinase increases transcriptional activation [177]; phosphorylation of serine 288 by cAMP-dependent protein kinase A results in nuclear translocation and subsequent transactivation [178]; and cyclin-dependent kinase 9 (cdk9) phosphorylates C/EBPβ and leads to an increase in HIV-1 gene expression [179] (Fig. 4).

C/EBPs interact with many nuclear proteins to activate transcription. In addition to binding other bZIP proteins, C/EBP recruits chromatin remodeling complexes such as SWI/SNF[180], cAMP response element-binding protein/p300 [181, 182], and p300/CREB-binding protein-associated factor [183] to the HIV-1 LTR. These proteins remodel the chromatin structure and increase transcription of the HIV-1 genome. C/EBP increases the phosphorylation of p300, which in turn alters its nuclear localization and increases its activity [184]. C/EBP can also act synergistically with Sp proteins to activate transcription of the HIV-1 LTR [185].

The importance of C/EBP factors in the regulation of HIV-1 gene expression is underscored by the discovery that a 6G configuration (a T-to-G change at nucleotide position 6) in C/EBP site I increases C/EBP binding, increases LTR activity, and is preferentially encountered in proviral LTRs derived from the brain of HIV-1-infected patients [42, 186]. C/EBP site II was also found to be preferentially conserved in the consensus subtype B configuration or to contain a 6G variation of this site, which are both high affinity sites for C/EBP factors in LTRs present in proviral DNA in cells located in the mid-frontal gyrus of the brain of infected individuals. A high rate of viral replication occurs in this region of the brain. Interestingly, the presence of the 6G configuration of this binding site also correlates with the presence of HIV-1-associated dementia [42, 44]. In contrast, the presence of a 4C C/EBP site II, which is a low-affinity C/EBP site, has been found preferentially in the cerebellum, a region of low viral replication [44]. This observation suggests that high affinity for C/EBP factors may contribute to the maintenance and/or pathogenesis of HIV-1 in the central nervous system, whereas low affinity sites such as 4C may contribute to lower levels of transcription required to maintain a latent reservoir of provirus. We have also identified a 3T configuration (a C-to-T change at position 3) of C/EBP site I that exhibits a low affinity for C/EBP within LTRs in the peripheral blood and brain and has also been shown to correlate with both late stage HIV disease and HIV-1-associated dementia [43], respectively.

ATF/CREB

ATF/CREB binds the HIV-1 LTR at a site immediately upstream of the C/EBP binding site I [38, 187] and at two sites downstream of the transcriptional start site (sites +160 to +167 and +92 to +102) to regulate the LTR [59, 60, 188, 189]. ATF/CREB and C/EBP factors can bind their adjacent upstream sites individually as homodimers, or C/EBP and ATF/CREB can heterodimerize with each other to regulate HIV-1 expression. This heterodimerization results in the recognition of a site composed of the 3' half of the ATF/CREB site and the 5' half of the C/EBP site [146]. As a result, in the presence of genetic variation that results in a low affinity C/EBP site, ATF/CREB is able to recruit C/EBP factors to the site and vice versa [146]. In addition to activating transcription, ATF/CREB can inhibit transcription by binding to Swi6, a component of the remodeling complex SWI/SNF, to promote the formation of heterochromatin [190].

AP-1 (Fos/Jun)

AP-1 proteins exist as homodimers of Jun family members (c-Jun, JunB, and JunD) or as heterodimers of Jun and Fos family members (c-Fos, FosB, Fra-1, and Fra-2) (reviewed in [191]). They bind a palindromic DNA sequence known as the TPA-responsive elements (TRE) at positions -306 to -285 and -242 to -222 of the LTR [59] as well as at positions +95 and +160, downstream of the transcriptional start site [59, 60, 188, 189]. The sequence of these sites has been shown to evolve in a manner that facilitates efficient cell type-specific binding of AP-1 [59, 192]. AP-1 acts as either an activator or repressor of transcription, depending on the components of the dimer [191, 193]. Once bound to the promoter, cFos/cJun heterodimers can recruit the SWI/SNF chromatin remodeling complex to activate transcription, whereas homodimers or heterodimers consisting of other family members lack this ability [194].

AP-1 mRNA is typically absent in quiescent cells; however, it is significantly up-regulated upon cellular stimulation [195]. Jun levels increase during monocytic maturation and become constitutively expressed in MDMs [196–200]. Despite being expressed, AP-1 in MDMs of some tissues, such as the lung, lacks the ability to bind DNA because of the lack of expression of Ref-1, a protein that modulates the oxidation state of Fos [201, 202]. In addition to being regulated by oxidation [201–203], AP-1 protein activity is further controlled post-transcriptionally by sumoylation, which inhibits protein activity [204, 205], and by phosphorylation, which increases activity in response to cellular stimulation [206].

In addition to directly regulating HIV-1 gene expression, AP-1 proteins can modulate the activity of other transcription factors. C/EBPβ dimerization with c-Fos or c-Jun results in C/EBP being unable to bind DNA thus a reduction in C/EBP-mediated transactivation [153, 154]. In contrast, C/EBPα dimerization with c-Jun or c-Fos forms a potent activator of transcription [207]. In response to mitogen or cytokine stimulation, the mitogen-activated protein kinases ERK1/ERK2 phosphorylate AP-1 (reviewed in [208] and [209]). This phosphorylation promotes the interaction of AP-1 with NF-κB and the enhancer element, which leads to the synergistic activation of the LTR [210–213]. This cascade of events is one mechanism by which HIV emerges from latency [210, 214].

Tat

Tat is a virus-encoded transcriptional transactivator that binds to the RNA secondary structure encoded by the transactivation region (TAR) in the repeat segment of the LTR (+1 to +59) [215, 216]. Once bound to the elongating transcript, Tat helps assemble the pre-initiation complex and recruits cdk9 to promote phosphorylation of RNA Pol II [217, 218] and P-TEFb to increase processivity of RNA Pol II [219–223]. Interestingly, mechanistic studies of this complex suggest that one of the functions of Tat is to increase the duration of P-TEFb occupancy at the HIV-1 LTR [224]. Tat also significantly remodels chromatin by recruiting the histone acyltransferases Tip60 [225, 226], human Nucleosome Assembly Protein-1 (hNAP-1) [227], p300/cAMP response element-binding protein [228, 229], and p/CAF [230], as well as the chromatin-remodeling complex SWI/SNF [231]. Tat activity is limited in monocytes due to the lack of sufficient levels of cyclin T1, a component of P-TEFb [54]. Differentiation into macrophages increases Cyclin T1 expression and results in strong Tat activity [54].

Tat regulates the activity of many other transcription factors through direct protein-protein interactions and the modulation of kinase activities. Tat promotes the phosphorylation of Sp1, which in turn increases binding of Sp to the LTR [92]. Conversely, Sp is also necessary to recruit Tat to the LTR [76], and deletion or mutation of the Sp binding sites in the promoter abolishes Tat activity [232, 233]. It is currently unclear whether direct interaction occurs between Sp factors and Tat [234–236]. In addition to regulating Sp1 activity, Tat increases the cooperation between NFAT and AP-1 proteins without altering independent binding of these transcription factors to DNA [137, 237]. It also promotes the interaction of NF-κB and AP-1 factors to synergistically activate transcription [238–240].

Vpr

Vpr is another virus-encoded protein that plays a direct role in the regulation of HIV-1 transcription [241–243]. Vpr is found in the viral particle and plays an important role in early transcriptional activation of the LTR before Tat can be expressed [244–248]. Its importance is highlighted by a recent study that describes alterations in Vpr that provide a significant reduction in Vpr nuclear import and virion incorporation uniquely in a long term non-progressor patient [249]. Vpr also causes cell cycle arrest in the G2 phase, the phase of the cell cycle when the LTR is most active, which results in apoptosis. [250]. It is necessary for viral replication in cells of the monocyte-macrophage lineage [251–256]. Interestingly, Vpr has been shown to interact with the nuclear form of uracil DNA glycosylase (UNG2), a cellular DNA repair enzyme, which helps incorporate this protein into virus particles leading to a decrease in viral mutation rate. Specifically, the lack of UNG in virions during virus replication in primary monocyte-derived macrophages further increases virus mutant frequencies by 18-fold compared with the 4-fold increase measured in actively dividing cells [257]. In addition, Vpr has been shown to concentrate at the nuclear envelope (NE) shortly after infection (4-6 hrs) as part of the pre-integration complex (PIC), supporting an interaction between Vpr and components of the nuclear pore complex [258–261], including the nucleoporin hCG1 [262]. Single-point Vpr mutants within the first α-helix of the protein such as Vpr-L23F and Vpr-K27M fail to associate with hCG1, but are still able to interact with other known relevant host partners of Vpr. In primary human monocyte-derived macrophages, these mutants fail to localize at the NE resulting in a diffuse nucleocytoplasmic distribution, impaired the Vpr-mediated G2-arrest of the cell cycle, and subsequently induced cell death. These observations were obtained in primary macrophages from some but not all donors indicating that the targeting of Vpr to the nuclear pore complex may constitute an early step toward Vpr-induced G2-arrest and subsequent apoptosis. These results also suggest that Vpr targeting to the nuclear pore complex is not absolutely required, but can enhance HIV-1 replication in macrophages [263]. Extracellular Vpr is found in the plasma and the CSF [254, 264] and can enter monocytes and macrophages and behave as if the protein was endogenously expressed [265–267]. Vpr binds the LTR in a sequence-specific manner to activate transcription directly [45, 46] and also interacts with Sp1 [268], TFIIB [269, 270], NF-κB [271], C/EBP [272], and Tat [244, 273] to enhance transcription of the HIV-1 genome. Vpr activates the DNA binding activity of AP-1 by promoting the phosphorylation of cFos and cJun in monocytes and macrophages [267]. It also promotes the translocation of NF-κB p50/p65 to the nucleus by promoting the phosphorylation of IκB [267], which allows an NF-κB- and AP-1-mediated increase in LTR activity.

C/EBP and Vpr interact at the HIV-1 LTR in two ways. Vpr has been shown to increase C/EBPβ DNA binding activity [272]. It has also been shown that Vpr has a high affinity for LTR C/EBP binding site I variants that exhibit a decreased affinity of the site for C/EBP. The presence of these LTR variants correlates with late-stage HIV-associated disease [45, 46]. Thus, as HIV-1-associated disease progresses, viral variants containing this type of LTR C/EBP site I may become more prevalent and function to facilitate a transition from C/EBP-mediated LTR activation to Vpr-mediated transactivation from that site. Alternatively, Vpr and C/EBP may form a complex at that site (Burdo and Wigdahl, unpublished observations). In addition to interacting with cellular proteins, Vpr interacts with Tat and activates transcription in an additive manner [244, 274].

Methylation

HIV proviral DNA that has integrated into the host genome also becomes subject to host factors that regulate chromatin organization and gene transcription. These mechanisms include histone modification, RNA interference/silencing, and DNA methylation. The mechanisms play a role in the control of gene expression, viral activation, and/or latency. DNA methylation of CpG islands within the HIV-1 LTR is one process that results in the downregulation/silencing of the integrated proviral genome [275–278]. This form of transcriptional silencing occurs by specific methyltransferases that are directed to the target DNA by methylation of lysine 9 of histone H3 through histone methyltransferases [279]. In cells of the monocyte-macrophage lineage, methylation of the LTR has been found to result in the transcriptional silencing of the promoter which contributes to limited access of transcription factors to the target DNA [280]. In addition, in the CD4+ T cell line ACH-2, the transcriptional silencing brought about by DNA methylation of the LTR can be reversed through TNF-α treatment of the cells which leads to demethylation of the 5' LTR and the induction of viral gene expression [281] showing that although this modification is inheritable, it is not permanent. The reduction of LTR expression is possibly explained by the binding of methyl-CpG-binding protein 1 complex and methyl-CpG-binding protein 2 to methylated Sp1 transcription factor binding sites, thereby inhibiting the binding of Sp1 transcription factors [282, 283]. In addition, the transcription factors USF and NF-κB lose affinity for their methylated LTR transcription factor binding sites as well [284]. Unfortunately, to date all of these studies have been performed in T cell lines and primary T cells, but not in cells of the monocyte-macrophage lineage.

Cytokines

Cytokines play a critical role in the pathogenesis of HIV-1. IL-6, TNFα, IL-1β, and other proinflammatory cytokine levels are elevated in the blood, bone marrow, and cerebrospinal fluid of HIV-infected patients [285, 286]. IL-6 and TNF-α are induced early after HIV monocytic infection, followed by their continued increased expression [52, 53, 287]. IL-6 is a potent activator of C/EBP, and exposure of monocytes to IL-6 results in increased HIV-1 replication. The increase in C/EBP activity then forms a positive feedback loop for IL-6 expression, because C/EBPβ binds to and activates the IL-6 promoter [288]. C/EBPs can also activate the genes encoding other proinflammatory cytokines such as IL-1β [289] and TNFα [290, 291]. TNFα is one of the most potent activators of NF-κB activity known. It acts by causing a signaling cascade that activates the IκB kinase complex, which then phosphorylates IκB, releasing NF-κB. The free NF-κB translocates to the nucleus and induces the activation of the HIV-1 LTR (Fig. 3 and 4).

In addition to being regulated by cytokines, chemokines contribute to HIV-1 infection and pathogenesis. The HIV-1 Nef protein induces HIV-infected macrophages to secrete at least two chemokines, MIP1α and MIP1β, which recruit and activate resting CD4+ T lymphocytes [292]. These T cells can then become infected and produce high levels of virus.

Summary of important monocytic regulatory pathways regulating the HIV-1 LTR

Regulation of HIV-1 transcription in cells of the monocyte-macrophage lineage varies considerably with the stage of cellular differentiation as well as in comparison to activated T cells. Specifically, it has been observed that cyclin T1 expression in monocytes is controlled by differentiation. Cyclin T1 increases as cells of the monocyte-macrophage lineage differentiate [47]. Unstimulated peripheral blood monocytes and myeloid progenitor cells support low levels of viral replication and activate transcription in response to cellular activation like T cells [27, 36, 48–54] whereas differentiated MDMs have increased viral replication but either do not respond to [45] or downregulate HIV transcription [48, 55] in response to cellular stimulation. As cells of the monocyte lineage differentiate, the ratio of Sp1 to Sp3 increases, resulting in an increase in HIV-1 transcription (McAllister and Wigdahl, unpublished observations). This process may result in low level HIV replication, or viral genomic silence, in circulating monocytes, and evasion of the host immune system until the cells are differentiated in peripheral tissues. The importance of the Sp sites also varies depending on the differentiation stage of the cell; in unstimulated monocytes, mutation of the Sp sites reduces LTR activity, whereas in MDMs, transcription of HIV and replication of SIVmac are abolished when these critical binding sites are knocked out [83–86]. NF-κB regulation of the LTR is also unique in MDMs. In MDMs, NF-κB is composed of Rel B bound to p50 or p52, whereas NF-κB in T cells is predominantly composed of p65 or c-Rel bound to p50 or p52 [97–100]. NF-κB DNA binding activity first occurs in monocytes as they progress from promonocytes to monocytes; however, in mature monocytes and MDMs, NF-κB is constitutively active in the nucleus, and its DNA binding activity is not increased further in response to cellular activation or differentiation [106]. Stimulation of T cells and monocytes by LPS results in enhanced HIV replication, a process that correlates with activation of NF-κB [27, 49–51, 113]. In differentiated primary MDMs, stimulation by LPS results, however, in the downregulation of LTR activity and viral replication [48].

NFAT, C/EBP, Jun and AP-1 transcription factor regulation of LTR activity also have distinct differences in monocyte-macrophages compared to T cells. NFAT binds the NF-κB binding sites in the enhancer in response to cellular activation in T cells but binds constitutively in monocytes [110, 112, 127, 130, 133]. Also, NFAT5, the most evolutionarily divergent NFAT member, regulates HIV replication in monocyte-MDMs [130] but has not been shown to do this in T cells. With regard to C/EBP, it has been shown that at least one upstream C/EBP binding site and the presence of C/EBP proteins are necessary for replication in cells of the monocyte-macrophage lineage but not in T cells [157–161]. Jun levels increase during monocytic maturation and become constitutively expressed in MDMs [196–200]. Despite being expressed, AP-1 in MDMs of some tissues, such as the lung, lacks the ability to bind DNA because of the lack of expression of Ref-1, a protein that modulates the oxidation state of Fos [201, 202].

The viral proteins Tat and Vpr have also been shown to have unique properties with regard to HIV-1 LTR activation in cells of the monocyte-macrophage lineage. Tat activity has been shown to be limited in monocytes due to the lack of sufficient levels of cyclin T1, a component of P-TEFb [54]. Differentiation into macrophages increases Cyclin T1 expression and results in strong Tat activity [54]. Vpr has been shown to be necessary for viral replication in cells of the monocyte-macrophage lineage but not in T cells [251–256]. Vpr has also been shown to specifically play a role in viral mutation rates in cells of the monocyte-macrophage lineage. Specifically, the lack of UNG in virions due to lack of Vpr binding to UNG during viral packaging led to increased virus mutant frequencies as indicated previously (18-fold increase compared to a 4-fold increase) [257]. In addition, genetic variation in Vpr has been shown in primary human monocyte-derived macrophages to fail in Vpr localization at the NE resulting in a diffuse nucleocytoplasmic distribution, impairing the Vpr-mediated G2-arrest of the cell cycle and the subsequent cell death induction, in some but not all donors [263].

Conclusions

Regulation of HIV-1 transcription in cells of the monocyte-macrophage lineage is a complex process involving the interaction of numerous factors that are expressed in a differentiation-dependent manner and whose activity is regulated by both cellular differentiation and extracellular signaling pathways. Although monocytes can be infected, this process is hindered at multiple steps in the viral lifecycle, including transcription. The mechanism behind the block to replication in monocytes has yet to be fully characterized, but it is clear that many factors make contributions. Monocytes express relatively low levels of the HIV co-receptor CCR5 [293, 294] and recently, it has been shown that viral entry is impaired in circulating monocytes [295]. Reverse transcription and integration are also impaired [295, 296]. At the transcriptional level, LTR activity is regulated by the ratio of activator to repressor isoforms of transcription factors, the phosphorylation state of transcription factors, the inducible translocation of NF-κB and NFAT factors from the cytoplasm, and the availability of viral transactivator proteins and their host co-factors (Fig. 4). Members of the AP-1 transcription factor family and relatively equal levels of nuclear Sp1 to Sp3 facilitate a modest level of basal transcription, whereas NF-κB and NFAT proteins remain sequestered in the cytoplasm in the early stages of monocytic differentiation. The presence of Tat has little effect on transcription in monocytes, as cyclin T1 expression is undetectable and other factors required for Tat activation are absent [54]. This lack of Tat activity contributes to replication block observed in unstimulated circulating monocytes.

Although circulating monocytes exhibit low levels of viral replication, replication increases in response to cytokine stimulation. During periods of inflammation caused by HIV-1 infection, co-pathogens, or opportunistic infections, levels of circulating cytokines such as IL-6 and TNFα increase and stimulate HIV-1 replication in monocytic cells. IL-6 increases the activity of C/EBP factors; these factors then activate the LTR and form a positive feedback loop by activating the promoters of cytokines, including TNFα and IL-6. In response to TNFα, NF-κB and NFAT5 translocate to the nucleus, and AP-1 DNA binding activity is stimulated to activate transcription as a result of changes in protein phosphorylation (Fig. 6). NFAT and NF-κB interact at the enhancer to activate transcription synergistically, whereas ATF/CREB, AP-1, and C/EBPα, β, and δ form homo- and heterodimers to regulate LTR activity. Vpr binds to the LTR directly and through interactions with other factors associated with the transcriptional complex in conjunction with AP-1, NF-κB, and C/EBP to activate transcription.

Figure 6
figure 6

Cytokine-regulation of HIV-1 transcription in monocytes. Cytokines play an integral role in regulating the availability and activity of transcription factors that regulate the LTR. TNFα strongly induces the nuclear localization of NF-κB in monocytes. As a result, the subsequently stimulated LTR interfaces with increased levels of Sp and NF-κB factors. Cellular activation increases the expression of C/EBP, particularly activation by IL-6. TNF-α, IL-1, and interferon-γ reduce the expression of C/EBPα and increase expression of both C/EBPβ and C/EBPδ. Stimulation increases the expression of AP-1 in the cell where its interaction with NF-κB at the enhancer element leads to synergistic activation of the LTR. (Black arrows: translocation to nucleus; red arrows: decrease in expression; green arrows: increase in expression).

As monocytes differentiate into macrophages, the permissiveness to viral replication increases dramatically, although MDMs lose the ability to further increase viral replication in response to extracellular stimuli. Cofactors necessary for Tat transactivation of the LTR are expressed, allowing a much greater level of HIV-1 transcription than is possible in monocytes. NF-κB, AP-1, and NFAT proteins are constitutively localized in the nucleus, and the activator Sp1 expression predominates over the repressor Sp3, resulting in greater availability of Sp1 (Fig. 7). The viral protein Nef also activates signaling cascades that result in enhanced binding of AP-1 to the LTR and enhanced cooperation between AP-1 and NF-κB [210, 214].

Figure 7
figure 7

Regulation of HIV-1 transcription in differentiated macrophages. In differentiated macrophages, NF-κB and NFAT are constitutively localized in the nucleus; however, in the presence of large amounts of NF-κB, NFAT is unable to bind the LTR. NF-κB-Sp1 protein-protein interactions bind the LTR cooperatively and activate transcription synergistically in response to cellular stimuli. Sp sites are necessary for viral replication, and the ratio of Sp1 proteins to Sp3 proteins increases, thus increasing transcription of the virus. As the cell matures, C/EBPα levels decrease and C/EBPβ and C/EBPδ levels increase. AP-1 is constitutively expressed but loses its ability to bind to the LTR. Tat binds to the transactivation response region (TAR) structure on the viral RNA and recruits (P-TEFb (the Cyclin dependent kinase 9 (Cdk9) and cyclin T1 (CycT1) complex) through binding to cyclin T1. Recruitment of P-TEFb to TAR induces hyperphosphorylation of CTD by Cdk9, thereby enhancing the transcriptional elongation of HIV-1.

Although macrophages support active viral replication, they are recognized as reservoirs of HIV-1 and quietly harbor the virus during latency. Host proteins that contribute to LTR activation in macrophages during productive viral infection ironically may also contribute to transcriptional silencing during latency. Sp1 proteins have been shown to bind the LTR constitutively, regardless of the level of transcription [297], and, in the latent stage, Sp1, NF-κB, AP1, and ATF/CREB may function as repressors of transcription by recruiting HDACs to the LTR and promoting the formation of heterochromatin. Although AP-1 proteins become constitutively expressed, the level of Ref-1, which is required for the DNA binding activity of AP-1, is significantly reduced in the nucleus of MDMs [201, 202]. This effectively renders nuclear AP-1 proteins inactive. In addition to their inability to transactivate the LTR, the constitutive presence of AP-1 proteins may be sufficient to disrupt the binding of C/EBP to the LTR, because inactive heterodimers of AP-1 and C/EBP may be more likely to form in the presence of excess AP-1 proteins. It is currently unknown what triggers the switch from latency to productive replication, however the presence of factors that can serve as both activators and repressors at the LTR likely contributes to the ability of the virus to resume replication very quickly upon the removal of repressive stimuli such as HAART therapy.

Genetic variation within the LTR also plays a role in HIV-1 transcription as HIV-associated disease progresses. Previous studies have shown that Vpr binds with high affinity to specific configurations of sequences within the HIV-1 LTR C/EBP site I and NF-κB site II, and may directly activate transcription. The HIV-1 LTR C/EBP-NF-κB genotypic configuration that exhibits high affinity for Vpr and low affinity for C/EBPβ is prevalent during late stage HIV/AIDS and in LTRs preferentially encountered in autopsied brain tissue from individuals with HAD at the time of death as compared to that from individuals without HAD. In parallel with these observations, additional studies have identified specific variants of the viral transactivator Tat from HAD brain tissue that are defective with respect to their ability to transactivate the LTR, but still retain the ability to activate promoters of a number of proinflammatory cytokine genes [298]. In some tissues, such as the brain, Tat may become less transcriptionally competent as HIV-associated disease progresses. In these circumstances, it is postulated that Vpr facilitates HIV-1 replication by transactivation of LTR-directed transcription in the absence of a fully active Tat protein.

Future directions

Transcription of the HIV-1 LTR is a highly complex process that involves the interplay of host and viral transcription factors coupled with a wide array of signaling pathways that are activated by extracellular stimuli. Targeting transcriptional pathways in drug discovery recently proved effective in treating certain cancers and may provide an opportunity for additional therapeutic agents in the highly active retroviral therapy (HAART) repertoire. Stat3 has been declared "one of the most important oncogenic transcription factors against which a targeted therapy is needed" [299]. Constitutive Stat3 activity has been observed in many cancers, including prostate [300], squamous cell [301], breast [302, 303], head, and neck cancers [302], and has been associated with a poor prognosis [300]. c-Myc activity has been implicated in prostate cancer, melanoma, and Burkitt's lymphoma, and an anti-myc antisense oligonucleotide has made it to clinical trials for the treatment of prostate cancer [304]. Transcription factors that play critical roles in the regulation of HIV-1, including NF-κB and Sp factors, are also the target of anti-cancer drug development. NF-κB has been implicated in playing a role in tumorigenesis in a variety of cancers [305, 306], including colon [307], prostate, breast, and lung [308, 309]. Small molecule inhibitors that target NF-κB are currently under development for the treatment of cancers [305], and have shown promise in small animal models [310, 311]. Many of these inhibit IκB phosphorylation, resulting in NF-κB being sequestered in the cytoplasm [312]. Bortezomib was recently approved by the FDA for the treatment of multiple myeloma. Developed as a reversible 26S proteasome inhibitor, it is now believed that its antitumor activity may be attributable to its inhibition of NF-κB [313–315]. Tolfenamic acid, a nonsteroidal anti-inflammatory drug approved for the treatment of migraine headaches, has been shown to inhibit pancreatic cancer cell growth in vitro and pancreatic and esophageal tumor growth in vivo by inducting the proteosomal degradation of Sp factors [316–320]. It also has been shown to decrease AP-2 and YY-1 transcription factor expression in breast cancer cells and tumors [320]. P-TEFb has been a target of chemotherapies for the treatment of renal, gastric, and lung cancers, as well as mantle-cell lymphoma, however clinical trials revealed that drugs targeting this factor were not effective as monotherapies but showed some promise when combined with other treatments [321–325]. Drugs targeting P-TEFb have been shown to inhibit HIV-1 transcription and replication in a dose-dependent manner in cell lines with minimal cytotoxicity, however the drugs were less effective and more cytotoxic in primary PBMCs [326]. Further study is necessary to determine the feasibility of applying other chemotherapeutic drugs that target host transcription factors to HAART therapy with important components of the developmental pathway focused on minimizing toxicity.

Vpr and Tat provide obvious candidates for targeted drug therapy directed against HIV. Inhibition of Vpr-mediated nuclear import by the compound hematoxylin has been shown to decrease viral replication [327], and fumagillin has been shown to suppress HIV-1 infection of macrophages by targeting Vpr-mediated growth arrest and transcriptional activity [328]. Peptide analogs of Tat have been shown to inhibit Tat's ability to recruit cdk2 to the LTR, and to decrease transcription in vitro and viral load in a small animal model of HIV-1 infection [329]. Small molecular inhibitors have also been developed that disrupt the Tat-TAR interaction, however these have not developed into clinical trials [330–333]. In addition to targeting individual viral proteins, unique structural motifs created at the interface between these factors and host transcription factors should also be considered in future studies.

Abbreviations

AIDS:

acquired immunodeficiency syndrome

AP-1:

activator protein 1

ATF/CREB:

activating transcription factor/cyclic AMP response element-binding

bZIP:

basic leucine zipper

C/EBP:

CCAAT enhancer binding protein

cdk9:

cyclin-dependent kinase 9

HAD:

HIV-1-associated dementia

HDACs:

histone deacetylases

HIV:

human immunodeficiency virus

HIV-1:

human immunodeficiency virus type 1

HIV-2:

human immunodeficiency virus type 2

IL-1:

interleukin-1

LTR:

long terminal repeat

MDMs:

monocyte-derived macrophages

NFAT:

nuclear factor of activated T cells

NF-κB:

nuclear factor kappa B

NRE:

negative regulatory element

SIV:

simian immunodeficiency virus

Sp:

stimulatory protein

TNFα:

tumor necrosis factor α.

References

  1. HIV/AIDS UJUNPo: AIDS Epidemic Update. 2007

    Google Scholar 

  2. Fusuma EE, Caruso SC, Lopez DF, Costa LJ, Janini LM, De Mendonca JS, Kallas EG, Diaz RS: Duplication of peri-kappaB and NF-kappab sites of the first human immunodeficiency virus type 2 (HIV-2) transmission in Brazil. AIDS Res Hum Retroviruses. 2005, 21: 965-970. 10.1089/aid.2005.21.965.

    CAS  PubMed  Google Scholar 

  3. Hightower M, Kallas EG: Diagnosis, antiretroviral therapy, and emergence of resistance to antiretroviral agents in HIV-2 infection: a review. Braz J Infect Dis. 2003, 7: 7-15. 10.1590/S1413-86702003000100002.

    CAS  PubMed  Google Scholar 

  4. Grant AD, De Cock KM: ABC of AIDS. HIV infection and AIDS in the developing world. Bmj. 2001, 322: 1475-1478. 10.1136/bmj.322.7300.1475.

    PubMed Central  CAS  PubMed  Google Scholar 

  5. Markovitz DM: Infection with the human immunodeficiency virus type 2. Ann Intern Med. 1993, 118: 211-218.

    CAS  PubMed  Google Scholar 

  6. Krebs FC, Ross H, McAllister J, Wigdahl B: HIV-1-associated central nervous system dysfunction. Adv Pharmacol. 2000, 49: 315-385. full_text.

    CAS  PubMed  Google Scholar 

  7. Wigdahl B, Guyton RA, Sarin PS: Human immunodeficiency virus infection of the developing human nervous system. Virology. 1987, 159: 440-445. 10.1016/0042-6822(87)90483-1.

    CAS  PubMed  Google Scholar 

  8. Wigdahl B, Kunsch C: Role of HIV in human nervous system dysfunction. AIDS Res Hum Retroviruses. 1989, 5: 369-374. 10.1089/aid.1989.5.369.

    CAS  PubMed  Google Scholar 

  9. Wigdahl B, Kunsch C: Human immunodeficiency virus infection and neurologic dysfunction. Prog Med Virol. 1990, 37: 1-46.

    CAS  PubMed  Google Scholar 

  10. Gartner S, Markovits P, Markovitz DM, Kaplan MH, Gallo RC, Popovic M: The role of mononuclear phagocytes in HTLV-III/LAV infection. Science. 1986, 233: 215-219. 10.1126/science.3014648.

    CAS  PubMed  Google Scholar 

  11. Gendelman HE, Orenstein JM, Baca LM, Weiser B, Burger H, Kalter DC, Meltzer MS: The macrophage in the persistence and pathogenesis of HIV infection. Aids. 1989, 3: 475-495. 10.1097/00002030-198908000-00001.

    CAS  PubMed  Google Scholar 

  12. Coleman CM, Wu L: HIV interactions with monocytes and dendritic cells: viral latency and reservoirs. Retrovirology. 2009, 6: 51-10.1186/1742-4690-6-51.

    PubMed Central  PubMed  Google Scholar 

  13. Boggiano C, Littman DR: HIV's vagina travelogue. Immunity. 2007, 26: 145-147. 10.1016/j.immuni.2007.02.001.

    CAS  PubMed  Google Scholar 

  14. Milush JM, Kosub D, Marthas M, Schmidt K, Scott F, Wozniakowski A, Brown C, Westmoreland S, Sodora DL: Rapid dissemination of SIV following oral inoculation. AIDS. 2004, 18: 2371-2380.

    PubMed  Google Scholar 

  15. Wiley CA, Schrier RD, Nelson JA, Lampert PW, Oldstone MB: Cellular localization of human immunodeficiency virus infection within the brains of acquired immune deficiency syndrome patients. Proc Natl Acad Sci USA. 1986, 83: 7089-7093. 10.1073/pnas.83.18.7089.

    PubMed Central  CAS  PubMed  Google Scholar 

  16. Ivey NS, MacLean AG, Lackner AA: Acquired immunodeficiency syndrome and the blood-brain barrier. J Neurovirol. 2009, 15: 111-122. 10.1080/13550280902769764.

    PubMed Central  PubMed  Google Scholar 

  17. Chayt KJ, Harper ME, Marselle LM, Lewin EB, Rose RM, Oleske JM, Epstein LG, Wong-Staal F, Gallo RC: Detection of HTLV-III RNA in lungs of patients with AIDS and pulmonary involvement. Jama. 1986, 256: 2356-2359. 10.1001/jama.256.17.2356.

    CAS  PubMed  Google Scholar 

  18. Armstrong JA, Horne R: Follicular dendritic cells and virus-like particles in AIDS-related lymphadenopathy. Lancet. 1984, 2: 370-372. 10.1016/S0140-6736(84)90540-3.

    CAS  PubMed  Google Scholar 

  19. Busch M, Beckstead J, Gantz D, Vyas G: Detection of human immunodeficiency virus infection of myeloid precursors in bone marrow samples from AIDS patients. Blood. 1986, 68 (Suppl 1): 122a-

    Google Scholar 

  20. McElrath MJ, Pruett JE, Cohn ZA: Mononuclear phagocytes of blood and bone marrow: comparative roles as viral reservoirs in human immunodeficiency virus type 1 infections. Proc Natl Acad Sci USA. 1989, 86: 675-679. 10.1073/pnas.86.2.675.

    PubMed Central  CAS  PubMed  Google Scholar 

  21. Atta MG, Lucas GM, Fine DM: HIV-associated nephropathy: epidemiology, pathogenesis, diagnosis and management. Expert Rev Anti Infect Ther. 2008, 6: 365-371. 10.1586/14787210.6.3.365.

    CAS  PubMed  Google Scholar 

  22. Brown A, Zhang H, Lopez P, Pardo CA, Gartner S: In vitro modeling of the HIV-macrophage reservoir. J Leukoc Biol. 2006, 80: 1127-1135. 10.1189/jlb.0206126.

    CAS  PubMed  Google Scholar 

  23. Koenig S, Gendelman HE, Orenstein JM, Dal Canto MC, Pezeshkpour GH, Yungbluth M, Janotta F, Aksamit A, Martin MA, Fauci AS: Detection of AIDS virus in macrophages in brain tissue from AIDS patients with encephalopathy. Science. 1986, 233: 1089-1093. 10.1126/science.3016903.

    CAS  PubMed  Google Scholar 

  24. Schrier RD, McCutchan JA, Venable JC, Nelson JA, Wiley CA: T-cell-induced expression of human immunodeficiency virus in macrophages. J Virol. 1990, 64: 3280-3288.

    PubMed Central  CAS  PubMed  Google Scholar 

  25. Garcia-Blanco MA, Cullen BR: Molecular basis of latency in pathogenic human viruses. Science. 1991, 254: 815-820. 10.1126/science.1658933.

    CAS  PubMed  Google Scholar 

  26. Rich EA, Chen IS, Zack JA, Leonard ML, O'Brien WA: Increased susceptibility of differentiated mononuclear phagocytes to productive infection with human immunodeficiency virus-1 (HIV-1). J Clin Invest. 1992, 89: 176-183. 10.1172/JCI115559.

    PubMed Central  CAS  PubMed  Google Scholar 

  27. Mikovits JA, Raziuddin , Gonda M, Ruta M, Lohrey NC, Kung HF, Ruscetti FW: Negative regulation of human immune deficiency virus replication in monocytes. Distinctions between restricted and latent expression in THP-1 cells. J Exp Med. 1990, 171: 1705-1720. 10.1084/jem.171.5.1705.

    CAS  PubMed  Google Scholar 

  28. Mann DL, Gartner S, Le Sane F, Buchow H, Popovic M: HIV-1 transmission and function of virus-infected monocytes/macrophages. J Immunol. 1990, 144: 2152-2158.

    CAS  PubMed  Google Scholar 

  29. Sabatier JM, Vives E, Mabrouk K, Benjouad A, Rochat H, Duval A, Hue B, Bahraoui E: Evidence for neurotoxic activity of tat from human immunodeficiency virus type 1. J Virol. 1991, 65: 961-967.

    PubMed Central  CAS  PubMed  Google Scholar 

  30. Giulian D, Wendt E, Vaca K, Noonan CA: The envelope glycoprotein of human immunodeficiency virus type 1 stimulates release of neurotoxins from monocytes. Proc Natl Acad Sci USA. 1993, 90: 2769-2773. 10.1073/pnas.90.7.2769.

    PubMed Central  CAS  PubMed  Google Scholar 

  31. Wu P, Price P, Du B, Hatch WC, Terwilliger EF: Direct cytotoxicity of HIV-1 envelope protein gp120 on human NT neurons. Neuroreport. 1996, 7: 1045-1049. 10.1097/00001756-199604100-00018.

    CAS  PubMed  Google Scholar 

  32. Esser R, von Briesen H, Brugger M, Ceska M, Glienke W, Muller S, Rehm A, Rubsamen-Waigmann H, Andreesen R: Secretory repertoire of HIV-infected human monocytes/macrophages. Pathobiology. 1991, 59: 219-222. 10.1159/000163649.

    CAS  PubMed  Google Scholar 

  33. Cox RA, Anders GT, Cappelli PJ, Johnson JE, Blanton HM, Seaworth BJ, Treasure RL: Production of tumor necrosis factor-alpha and interleukin-1 by alveolar macrophages from HIV-1-infected persons. AIDS Res Hum Retroviruses. 1990, 6: 431-441. 10.1089/aid.1990.6.431.

    CAS  PubMed  Google Scholar 

  34. Nakajima K, Martinez-Maza O, Hirano T, Breen EC, Nishanian PG, Salazar-Gonzalez JF, Fahey JL, Kishimoto T: Induction of IL-6 (B cell stimulatory factor-2/IFN-beta 2) production by HIV. J Immunol. 1989, 142: 531-536.

    CAS  PubMed  Google Scholar 

  35. Roy S, Fitz-Gibbon L, Poulin L, Wainberg MA: Infection of human monocytes/macrophages by HIV-1: effect on secretion of IL-1 activity. Immunology. 1988, 64: 233-239.

    PubMed Central  CAS  PubMed  Google Scholar 

  36. Alexaki A, Quiterio SJ, Liu Y, Irish B, Kilareski E, Nonnemacher MR, Wigdahl B: PMA-induced differentiation of a bone marrow progenitor cell line activates HIV-1 LTR-driven transcription. DNA Cell Biol. 2007, 26: 387-394. 10.1089/dna.2006.0542.

    CAS  PubMed  Google Scholar 

  37. Fields BN, Knipe DM, Howley PM: Fields' virology. 2007, Philadelphia: Wolters Kluwer Health/Lippincott Williams & Wilkins, 5

    Google Scholar 

  38. Krebs FC, Mehrens D, Pomeroy S, Goodenow MM, Wigdahl B: Human immunodeficiency virus type 1 long terminal repeat quasispecies differ in basal transcription and nuclear factor recruitment in human glial cells and lymphocytes. J Biomed Sci. 1998, 5: 31-44. 10.1007/BF02253354.

    CAS  PubMed  Google Scholar 

  39. McAllister JJ, Phillips D, Millhouse S, Conner J, Hogan T, Ross HL, Wigdahl B: Analysis of the HIV-1 LTR NF-kappaB-proximal Sp site III: evidence for cell type-specific gene regulation and viral replication. Virology. 2000, 274: 262-277. 10.1006/viro.2000.0476.

    CAS  PubMed  Google Scholar 

  40. Millhouse S, Krebs FC, Yao J, McAllister JJ, Conner J, Ross H, Wigdahl B: Sp1 and related factors fail to interact with the NF-kappaB-proximal G/C box in the LTR of a replication competent, brain-derived strain of HIV-1 (YU-2). J Neurovirol. 1998, 4: 312-323. 10.3109/13550289809114532.

    CAS  PubMed  Google Scholar 

  41. Nonnemacher MR, Irish BP, Liu Y, Mauger D, Wigdahl B: Specific sequence configurations of HIV-1 LTR G/C box array result in altered recruitment of Sp isoforms and correlate with disease progression. J Neuroimmunol. 2004, 157: 39-47. 10.1016/j.jneuroim.2004.08.021.

    CAS  PubMed  Google Scholar 

  42. Ross HL, Gartner S, McArthur JC, Corboy JR, McAllister JJ, Millhouse S, Wigdahl B: HIV-1 LTR C/EBP binding site sequence configurations preferentially encountered in brain lead to enhanced C/EBP factor binding and increased LTR-specific activity. J Neurovirol. 2001, 7: 235-249. 10.1080/13550280152403281.

    CAS  PubMed  Google Scholar 

  43. Hogan TH, Stauff DL, Krebs FC, Gartner S, Quiterio SJ, Wigdahl B: Structural and functional evolution of human immunodeficiency virus type 1 long terminal repeat CCAAT/enhancer binding protein sites and their use as molecular markers for central nervous system disease progression. J Neurovirol. 2003, 9: 55-68.

    CAS  PubMed  Google Scholar 

  44. Burdo TH, Gartner S, Mauger D, Wigdahl B: Region-specific distribution of human immunodeficiency virus type 1 long terminal repeats containing specific configurations of CCAAT/enhancer-binding protein site II in brains derived from demented and nondemented patients. J Neurovirol. 2004, 10 (Suppl 1): 7-14.

    CAS  PubMed  Google Scholar 

  45. Burdo TH, Nonnemacher M, Irish BP, Choi CH, Krebs FC, Gartner S, Wigdahl B: High-affinity interaction between HIV-1 Vpr and specific sequences that span the C/EBP and adjacent NF-kappaB sites within the HIV-1 LTR correlate with HIV-1-associated dementia. DNA Cell Biol. 2004, 23: 261-269. 10.1089/104454904773819842.

    CAS  PubMed  Google Scholar 

  46. Hogan TH, Nonnemacher MR, Krebs FC, Henderson A, Wigdahl B: HIV-1 Vpr binding to HIV-1 LTR C/EBP cis-acting elements and adjacent regions is sequence-specific. Biomed Pharmacother. 2003, 57: 41-48. 10.1016/S0753-3322(02)00333-5.

    CAS  PubMed  Google Scholar 

  47. Yu W, Wang Y, Shaw CA, Qin XF, Rice AP: Induction of the HIV-1 Tat co-factor cyclin T1 during monocyte differentiation is required for the regulated expression of a large portion of cellular mRNAs. Retrovirology. 2006, 3: 32-10.1186/1742-4690-3-32.

    PubMed Central  PubMed  Google Scholar 

  48. Bernstein MS, Tong-Starksen SE, Locksley RM: Activation of human monocyte--derived macrophages with lipopolysaccharide decreases human immunodeficiency virus replication in vitro at the level of gene expression. J Clin Invest. 1991, 88: 540-545. 10.1172/JCI115337.

    PubMed Central  CAS  PubMed  Google Scholar 

  49. Molina JM, Scadden DT, Byrn R, Dinarello CA, Groopman JE: Production of tumor necrosis factor alpha and interleukin 1 beta by monocytic cells infected with human immunodeficiency virus. J Clin Invest. 1989, 84: 733-737. 10.1172/JCI114230.

    PubMed Central  CAS  PubMed  Google Scholar 

  50. Pomerantz RJ, Feinberg MB, Trono D, Baltimore D: Lipopolysaccharide is a potent monocyte/macrophage-specific stimulator of human immunodeficiency virus type 1 expression. J Exp Med. 1990, 172: 253-261. 10.1084/jem.172.1.253.

    CAS  PubMed  Google Scholar 

  51. Poli G, Kinter A, Justement JS, Kehrl JH, Bressler P, Stanley S, Fauci AS: Tumor necrosis factor alpha functions in an autocrine manner in the induction of human immunodeficiency virus expression. Proc Natl Acad Sci USA. 1990, 87: 782-785. 10.1073/pnas.87.2.782.

    PubMed Central  CAS  PubMed  Google Scholar 

  52. Koyanagi Y, O'Brien WA, Zhao JQ, Golde DW, Gasson JC, Chen IS: Cytokines alter production of HIV-1 from primary mononuclear phagocytes. Science. 1988, 241: 1673-1675. 10.1126/science.3047875.

    CAS  PubMed  Google Scholar 

  53. Folks TM, Kessler SW, Orenstein JM, Justement JS, Jaffe ES, Fauci AS: Infection and replication of HIV-1 in purified progenitor cells of normal human bone marrow. Science. 1988, 242: 919-922. 10.1126/science.2460922.

    CAS  PubMed  Google Scholar 

  54. Dong C, Kwas C, Wu L: Transcriptional restriction of human immunodeficiency virus type 1 gene expression in undifferentiated primary monocytes. J Virol. 2009, 83: 3518-3527. 10.1128/JVI.02665-08.

    PubMed Central  CAS  PubMed  Google Scholar 

  55. Sundaravaradan V, Saxena SK, Ramakrishnan R, Yedavalli VR, Harris DT, Ahmad N: Differential HIV-1 replication in neonatal and adult blood mononuclear cells is influenced at the level of HIV-1 gene expression. Proc Natl Acad Sci USA. 2006, 103: 11701-11706. 10.1073/pnas.0602185103.

    PubMed Central  CAS  PubMed  Google Scholar 

  56. Orenstein JM, Fox C, Wahl SM: Macrophages as a source of HIV during opportunistic infections. Science. 1997, 276: 1857-1861. 10.1126/science.276.5320.1857.

    CAS  PubMed  Google Scholar 

  57. Igarashi T, Brown CR, Endo Y, Buckler-White A, Plishka R, Bischofberger N, Hirsch V, Martin MA: Macrophage are the principal reservoir and sustain high virus loads in rhesus macaques after the depletion of CD4+ T cells by a highly pathogenic simian immunodeficiency virus/HIV type 1 chimera (SHIV): Implications for HIV-1 infections of humans. Proc Natl Acad Sci USA. 2001, 98: 658-663. 10.1073/pnas.021551798.

    PubMed Central  CAS  PubMed  Google Scholar 

  58. Jones KA, Kadonaga JT, Luciw PA, Tjian R: Activation of the AIDS retrovirus promoter by the cellular transcription factor, Sp1. Science. 1986, 232: 755-759. 10.1126/science.3008338.

    CAS  PubMed  Google Scholar 

  59. Canonne-Hergaux F, Aunis D, Schaeffer E: Interactions of the transcription factor AP-1 with the long terminal repeat of different human immunodeficiency virus type 1 strains in Jurkat, glial, and neuronal cells. J Virol. 1995, 69: 6634-6642.

    PubMed Central  CAS  PubMed  Google Scholar 

  60. Van Lint C, Amella CA, Emiliani S, John M, Jie T, Verdin E: Transcription factor binding sites downstream of the human immunodeficiency virus type 1 transcription start site are important for virus infectivity. J Virol. 1997, 71: 6113-6127.

    PubMed Central  CAS  PubMed  Google Scholar 

  61. Peeters A, Lambert PF, Deacon NJ: A fourth Sp1 site in the human immunodeficiency virus type 1 long terminal repeat is essential for negative-sense transcription. J Virol. 1996, 70: 6665-6672.

    PubMed Central  CAS  PubMed  Google Scholar 

  62. Hagen G, Muller S, Beato M, Suske G: Cloning by recognition site screening of two novel GT box binding proteins: a family of Sp1 related genes. Nucleic Acids Res. 1992, 20: 5519-5525. 10.1093/nar/20.21.5519.

    PubMed Central  CAS  PubMed  Google Scholar 

  63. Dynan WS, Tjian R: Isolation of transcription factors that discriminate between different promoters recognized by RNA polymerase II. Cell. 1983, 32: 669-680. 10.1016/0092-8674(83)90053-3.

    CAS  PubMed  Google Scholar 

  64. Kadonaga JT, Carner KR, Masiarz FR, Tjian R: Isolation of cDNA encoding transcription factor Sp1 and functional analysis of the DNA binding domain. Cell. 1987, 51: 1079-1090. 10.1016/0092-8674(87)90594-0.

    CAS  PubMed  Google Scholar 

  65. Kingsley C, Winoto A: Cloning of GT box-binding proteins: a novel Sp1 multigene family regulating T-cell receptor gene expression. Mol Cell Biol. 1992, 12: 4251-4261.

    PubMed Central  CAS  PubMed  Google Scholar 

  66. Majello B, De Luca P, Hagen G, Suske G, Lania L: Different members of the Sp1 multigene family exert opposite transcriptional regulation of the long terminal repeat of HIV-1. Nucleic Acids Res. 1994, 22: 4914-4921. 10.1093/nar/22.23.4914.

    PubMed Central  CAS  PubMed  Google Scholar 

  67. Hagen G, Muller S, Beato M, Suske G: Sp1-mediated transcriptional activation is repressed by Sp3. Embo J. 1994, 13: 3843-3851.

    PubMed Central  CAS  PubMed  Google Scholar 

  68. Udvadia AJ, Templeton DJ, Horowitz JM: Functional interactions between the retinoblastoma (Rb) protein and Sp-family members: superactivation by Rb requires amino acids necessary for growth suppression. Proc Natl Acad Sci USA. 1995, 92: 3953-3957. 10.1073/pnas.92.9.3953.

    PubMed Central  CAS  PubMed  Google Scholar 

  69. Kennett SB, Udvadia AJ, Horowitz JM: Sp3 encodes multiple proteins that differ in their capacity to stimulate or repress transcription. Nucleic Acids Res. 1997, 25: 3110-3117. 10.1093/nar/25.15.3110.

    PubMed Central  CAS  PubMed  Google Scholar 

  70. Supp DM, Witte DP, Branford WW, Smith EP, Potter SS: Sp4, a member of the Sp1-family of zinc finger transcription factors, is required for normal murine growth, viability, and male fertility. Dev Biol. 1996, 176: 284-299. 10.1006/dbio.1996.0134.

    CAS  PubMed  Google Scholar 

  71. Hagen G, Dennig J, Preiss A, Beato M, Suske G: Functional analyses of the transcription factor Sp4 reveal properties distinct from Sp1 and Sp3. J Biol Chem. 1995, 270: 24989-24994. 10.1074/jbc.270.42.24989.

    CAS  PubMed  Google Scholar 

  72. Gill G, Pascal E, Tseng ZH, Tjian R: A glutamine-rich hydrophobic patch in transcription factor Sp1 contacts the dTAFII110 component of the Drosophila TFIID complex and mediates transcriptional activation. Proc Natl Acad Sci USA. 1994, 91: 192-196. 10.1073/pnas.91.1.192.

    PubMed Central  CAS  PubMed  Google Scholar 

  73. Emili A, Greenblatt J, Ingles CJ: Species-specific interaction of the glutamine-rich activation domains of Sp1 with the TATA box-binding protein. Mol Cell Biol. 1994, 14: 1582-1593.

    PubMed Central  CAS  PubMed  Google Scholar 

  74. Chiang CM, Roeder RG: Cloning of an intrinsic human TFIID subunit that interacts with multiple transcriptional activators. Science. 1995, 267: 531-536. 10.1126/science.7824954.

    CAS  PubMed  Google Scholar 

  75. Huang LM, Jeang KT: Increased spacing between Sp1 and TATAA renders human immunodeficiency virus type 1 replication defective: implication for Tat function. J Virol. 1993, 67: 6937-6944.

    PubMed Central  CAS  PubMed  Google Scholar 

  76. Yedavalli VS, Benkirane M, Jeang KT: Tat and trans-activation-responsive (TAR) RNA-independent induction of HIV-1 long terminal repeat by human and murine cyclin T1 requires Sp1. J Biol Chem. 2003, 278: 6404-6410. 10.1074/jbc.M209162200.

    CAS  PubMed  Google Scholar 

  77. Widlak P, Gaynor RB, Garrard WT: In vitro chromatin assembly of the HIV-1 promoter. ATP-dependent polar repositioning of nucleosomes by Sp1 and NFkappaB. J Biol Chem. 1997, 272: 17654-17661. 10.1074/jbc.272.28.17654.

    CAS  PubMed  Google Scholar 

  78. Pazin MJ, Sheridan PL, Cannon K, Cao Z, Keck JG, Kadonaga JT, Jones KA: NF-kappa B-mediated chromatin reconfiguration and transcriptional activation of the HIV-1 enhancer in vitro. Genes Dev. 1996, 10: 37-49. 10.1101/gad.10.1.37.

    CAS  PubMed  Google Scholar 

  79. Won J, Yim J, Kim TK: Sp1 and Sp3 recruit histone deacetylase to repress transcription of human telomerase reverse transcriptase (hTERT) promoter in normal human somatic cells. J Biol Chem. 2002, 277: 38230-38238. 10.1074/jbc.M206064200.

    CAS  PubMed  Google Scholar 

  80. Doetzlhofer A, Rotheneder H, Lagger G, Koranda M, Kurtev V, Brosch G, Wintersberger E, Seiser C: Histone deacetylase 1 can repress transcription by binding to Sp1. Mol Cell Biol. 1999, 19: 5504-5511.

    PubMed Central  CAS  PubMed  Google Scholar 

  81. Sun JM, Chen HY, Moniwa M, Litchfield DW, Seto E, Davie JR: The transcriptional repressor Sp3 is associated with CK2-phosphorylated histone deacetylase 2. J Biol Chem. 2002, 277: 35783-35786. 10.1074/jbc.C200378200.

    CAS  PubMed  Google Scholar 

  82. Li Y, Mak G, Franza BR: In vitro study of functional involvement of Sp1, NF-kappa B/Rel, and AP1 in phorbol 12-myristate 13-acetate-mediated HIV-1 long terminal repeat activation. J Biol Chem. 1994, 269: 30616-30619.

    CAS  PubMed  Google Scholar 

  83. Moses AV, Ibanez C, Gaynor R, Ghazal P, Nelson JA: 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.

    PubMed Central  CAS  PubMed  Google Scholar 

  84. Du Z, Ilyinskii PO, Sasseville VG, Newstein M, Lackner AA, Desrosiers RC: Requirements for lymphocyte activation by unusual strains of simian immunodeficiency virus. J Virol. 1996, 70: 4157-4161.

    PubMed Central  CAS  PubMed  Google Scholar 

  85. Ilyinskii PO, Desrosiers RC: Efficient transcription and replication of simian immunodeficiency virus in the absence of NF-kappaB and Sp1 binding elements. J Virol. 1996, 70: 3118-3126.

    PubMed Central  CAS  PubMed  Google Scholar 

  86. Ilyinskii PO, Simon MA, Czajak SC, Lackner AA, Desrosiers RC: Induction of AIDS by simian immunodeficiency virus lacking NF-kappaB and SP1 binding elements. J Virol. 1997, 71: 1880-1887.

    PubMed Central  CAS  PubMed  Google Scholar 

  87. Resendes KK, Rosmarin AG: Sp1 control of gene expression in myeloid cells. Crit Rev Eukaryot Gene Expr. 2004, 14: 171-181. 10.1615/CritRevEukaryotGeneExpr.v14.i3.20.

    CAS  PubMed  Google Scholar 

  88. Chu S, Ferro TJ: Sp1: regulation of gene expression by phosphorylation. Gene. 2005, 348: 1-11. 10.1016/j.gene.2005.01.013.

    CAS  PubMed  Google Scholar 

  89. Jackson S, Gottlieb T, Hartley K: Phosphorylation of transcription factor Sp1 by the DNA-dependent protein kinase. Adv Second Messenger Phosphoprotein Res. 1993, 28: 279-286.

    CAS  PubMed  Google Scholar 

  90. Jackson SP, MacDonald JJ, Lees-Miller S, Tjian R: GC box binding induces phosphorylation of Sp1 by a DNA-dependent protein kinase. Cell. 1990, 63: 155-165. 10.1016/0092-8674(90)90296-Q.

    CAS  PubMed  Google Scholar 

  91. Vlach J, Garcia A, Jacque JM, Rodriguez MS, Michelson S, Virelizier JL: Induction of Sp1 phosphorylation and NF-kappa B-independent HIV promoter domain activity in T lymphocytes stimulated by okadaic acid. Virology. 1995, 208: 753-761. 10.1006/viro.1995.1207.

    CAS  PubMed  Google Scholar 

  92. Chun RF, Semmes OJ, Neuveut C, Jeang KT: Modulation of Sp1 phosphorylation by human immunodeficiency virus type 1 Tat. J Virol. 1998, 72: 2615-2629.

    PubMed Central  CAS  PubMed  Google Scholar 

  93. Jochmann R, Thurau M, Jung S, Hofmann C, Naschberger E, Kremmer E, Harrer T, Miller M, Schaft N, Sturzl M: O-linked N-acetylglucosaminylation of Sp1 inhibits the human immunodeficiency virus type 1 promoter. J Virol. 2009, 83: 3704-3718. 10.1128/JVI.01384-08.

    PubMed Central  CAS  PubMed  Google Scholar 

  94. Mingyan Y, Xinyong L, De Clercq E: NF-kappaB: the inducible factors of HIV-1 transcription and their inhibitors. Mini Rev Med Chem. 2009, 9: 60-69. 10.2174/138955709787001677.

    PubMed  Google Scholar 

  95. Roulston A, Lin R, Beauparlant P, Wainberg MA, Hiscott J: Regulation of human immunodeficiency virus type 1 and cytokine gene expression in myeloid cells by NF-kappa B/Rel transcription factors. Microbiol Rev. 1995, 59: 481-505.

    PubMed Central  CAS  PubMed  Google Scholar 

  96. Phares W, Franza BR, Herr W: The kappa B enhancer motifs in human immunodeficiency virus type 1 and simian virus 40 recognize different binding activities in human Jurkat and H9 T cells: evidence for NF-kappa B-independent activation of the kappa B motif. J Virol. 1992, 66: 7490-7498.

    PubMed Central  CAS  PubMed  Google Scholar 

  97. 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.

    CAS  PubMed  Google Scholar 

  98. Asin S, Bren GD, Carmona EM, Solan NJ, Paya CV: NF-kappaB cis-acting motifs of the human immunodeficiency virus (HIV) long terminal repeat regulate HIV transcription in human macrophages. J Virol. 2001, 75: 11408-11416. 10.1128/JVI.75.23.11408-11416.2001.

    PubMed Central  CAS  PubMed  Google Scholar 

  99. Asin S, Taylor JA, Trushin S, Bren G, Paya CV: Ikappakappa mediates NF-kappaB activation in human immunodeficiency virus-infected cells. J Virol. 1999, 73: 3893-3903.

    PubMed Central  CAS  PubMed  Google Scholar 

  100. Neumann M, Fries H, Scheicher C, Keikavoussi P, Kolb-Maurer A, Brocker E, Serfling E, Kampgen E: Differential expression of Rel/NF-kappaB and octamer factors is a hallmark of the generation and maturation of dendritic cells. Blood. 2000, 95: 277-285.

    CAS  PubMed  Google Scholar 

  101. Baeuerle P, Baltimore D: Activation of DNA-Binding Activity in an Apparently Cytoplasmic Precursor of the NF-KB Transcription Factor. Cell. 1988, 53: 211-217. 10.1016/0092-8674(88)90382-0.

    CAS  PubMed  Google Scholar 

  102. Beg AA, Finco TS, Nantermet PV, Baldwin AS: Tumor necrosis factor and interleukin-1 lead to phosphorylation and loss of I kappa B alpha: a mechanism for NF-kappa B activation. Mol Cell Biol. 1993, 13: 3301-3310.

    PubMed Central  CAS  PubMed  Google Scholar 

  103. Barboric M, Nissen RM, Kanazawa S, Jabrane-Ferrat N, Peterlin BM: NF-kappaB binds P-TEFb to stimulate transcriptional elongation by RNA polymerase II. Mol Cell. 2001, 8: 327-337. 10.1016/S1097-2765(01)00314-8.

    CAS  PubMed  Google Scholar 

  104. Coiras M, Lopez-Huertas MR, Rullas J, Mittelbrunn M, Alcami J: Basal shuttle of NF-kappaB/I kappaB alpha in resting T lymphocytes regulates HIV-1 LTR dependent expression. Retrovirology. 2007, 4: 56-10.1186/1742-4690-4-56.

    PubMed Central  PubMed  Google Scholar 

  105. Ashburner BP, Westerheide SD, Baldwin AS: The p65 (RelA) subunit of NF-kappaB interacts with the histone deacetylase (HDAC) corepressors HDAC1 and HDAC2 to negatively regulate gene expression. Mol Cell Biol. 2001, 21: 7065-7077. 10.1128/MCB.21.20.7065-7077.2001.

    PubMed Central  CAS  PubMed  Google Scholar 

  106. Griffin GE, Leung K, Folks TM, Kunkel S, Nabel GJ: Activation of HIV gene expression during monocyte differentiation by induction of NF-kappa B. Nature. 1989, 339: 70-73. 10.1038/339070a0.

    CAS  PubMed  Google Scholar 

  107. Roulston A, Beauparlant P, Rice N, Hiscott J: Chronic human immunodeficiency virus type 1 infection stimulates distinct NF-kappa B/rel DNA binding activities in myelomonoblastic cells. J Virol. 1993, 67: 5235-5246.

    PubMed Central  CAS  PubMed  Google Scholar 

  108. Perkins ND, Agranoff AB, Pascal E, Nabel GJ: An interaction between the DNA-binding domains of RelA(p65) and Sp1 mediates human immunodeficiency virus gene activation. Mol Cell Biol. 1994, 14: 6570-6583.

    PubMed Central  CAS  PubMed  Google Scholar 

  109. Perkins ND, Edwards NL, Duckett CS, Agranoff AB, Schmid RM, Nabel GJ: A cooperative interaction between NF-kappa B and Sp1 is required for HIV-1 enhancer activation. Embo J. 1993, 12: 3551-3558.

    PubMed Central  CAS  PubMed  Google Scholar 

  110. Cron RQ, Bartz SR, Clausell A, Bort SJ, Klebanoff SJ, Lewis DB: NFAT1 enhances HIV-1 gene expression in primary human CD4 T cells. Clin Immunol. 2000, 94: 179-191. 10.1006/clim.1999.4831.

    CAS  PubMed  Google Scholar 

  111. Kawakami K, Scheidereit C, Roeder RG: Identification and purification of a human immunoglobulin-enhancer-binding protein (NF-kappa B) that activates transcription from a human immunodeficiency virus type 1 promoter in vitro. Proc Natl Acad Sci USA. 1988, 85: 4700-4704. 10.1073/pnas.85.13.4700.

    PubMed Central  CAS  PubMed  Google Scholar 

  112. Kinoshita S, Su L, Amano M, Timmerman LA, Kaneshima H, Nolan GP: The T cell activation factor NF-ATc positively regulates HIV-1 replication and gene expression in T cells. Immunity. 1997, 6: 235-244. 10.1016/S1074-7613(00)80326-X.

    CAS  PubMed  Google Scholar 

  113. Berg RS, Aggerholm A, Bertelsen LS, Ostergaard L, Paludan SR: Role of mitogen-activated protein kinases, nuclear factor-kappaB, and interferon regulatory factor 3 in Toll-like receptor 4-mediated activation of HIV long terminal repeat. APMIS. 2009, 117: 124-132. 10.1111/j.1600-0463.2008.00024.x.

    PubMed  Google Scholar 

  114. Mallardo M, Dragonetti E, Baldassarre F, Ambrosino C, Scala G, Quinto I: An NF-kappaB site in the 5'-untranslated leader region of the human immunodeficiency virus type 1 enhances the viral expression in response to NF-kappaB-activating stimuli. J Biol Chem. 1996, 271: 20820-20827. 10.1074/jbc.271.34.20820.

    CAS  PubMed  Google Scholar 

  115. Montano MA, Kripke K, Norina CD, Achacoso P, Herzenberg LA, Roy AL, Nolan GP: NF-kappa B homodimer binding within the HIV-1 initiator region and interactions with TFII-I. Proc Natl Acad Sci USA. 1996, 93: 12376-12381. 10.1073/pnas.93.22.12376.

    PubMed Central  CAS  PubMed  Google Scholar 

  116. Parrott C, Seidner T, Duh E, Leonard J, Theodore TS, Buckler-White A, Martin MA, Rabson AB: Variable role of the long terminal repeat Sp1-binding sites in human immunodeficiency virus replication in T lymphocytes. J Virol. 1991, 65: 1414-1419.

    PubMed Central  CAS  PubMed  Google Scholar 

  117. Ross EK, Buckler-White AJ, Rabson AB, Englund G, Martin MA: Contribution of NF-kappa B and Sp1 binding motifs to the replicative capacity of human immunodeficiency virus type 1: distinct patterns of viral growth are determined by T-cell types. J Virol. 1991, 65: 4350-4358.

    PubMed Central  CAS  PubMed  Google Scholar 

  118. Leonard J, Parrott C, Buckler-White AJ, Turner W, Ross EK, Martin MA, Rabson AB: The NF-kappa B binding sites in the human immunodeficiency virus type 1 long terminal repeat are not required for virus infectivity. J Virol. 1989, 63: 4919-4924.

    PubMed Central  CAS  PubMed  Google Scholar 

  119. Crabtree GR, Olson EN: NFAT signaling: choreographing the social lives of cells. Cell. 2002, 109 (Suppl): S67-79. 10.1016/S0092-8674(02)00699-2.

    CAS  PubMed  Google Scholar 

  120. Hogan PG, Chen L, Nardone J, Rao A: Transcriptional regulation by calcium, calcineurin, and NFAT. Genes Dev. 2003, 17: 2205-2232. 10.1101/gad.1102703.

    CAS  PubMed  Google Scholar 

  121. Lopez-Rodriguez C, Aramburu J, Rakeman AS, Rao A: NFAT5, a constitutively nuclear NFAT protein that does not cooperate with Fos and Jun. Proc Natl Acad Sci USA. 1999, 96: 7214-7219. 10.1073/pnas.96.13.7214.

    PubMed Central  CAS  PubMed  Google Scholar 

  122. Shaw JP, Utz PJ, Durand DB, Toole JJ, Emmel EA, Crabtree GR: Identification of a putative regulator of early T cell activation genes. Science. 1988, 241: 202-205. 10.1126/science.3260404.

    CAS  PubMed  Google Scholar 

  123. Serfling E, Berberich-Siebelt F, Avots A, Chuvpilo S, Klein-Hessling S, Jha MK, Kondo E, Pagel P, Schulze-Luehrmann J, Palmetshofer A: NFAT and NF-kappaB factors-the distant relatives. Int J Biochem Cell Biol. 2004, 36: 1166-1170. 10.1016/j.biocel.2003.07.002.

    CAS  PubMed  Google Scholar 

  124. Fortin JF, Barbeau B, Robichaud GA, Pare ME, Lemieux AM, Tremblay MJ: Regulation of nuclear factor of activated T cells by phosphotyrosyl-specific phosphatase activity: a positive effect on HIV-1 long terminal repeat-driven transcription and a possible implication of SHP-1. Blood. 2001, 97: 2390-2400. 10.1182/blood.V97.8.2390.

    CAS  PubMed  Google Scholar 

  125. Crabtree GR: Generic signals and specific outcomes: signaling through Ca2+, calcineurin, and NF-AT. Cell. 1999, 96: 611-614. 10.1016/S0092-8674(00)80571-1.

    CAS  PubMed  Google Scholar 

  126. Stankunas K, Graef IA, Neilson JR, Park SH, Crabtree GR: Signaling through calcium, calcineurin, and NF-AT in lymphocyte activation and development. Cold Spring Harb Symp Quant Biol. 1999, 64: 505-516. 10.1101/sqb.1999.64.505.

    CAS  PubMed  Google Scholar 

  127. Giffin MJ, Stroud JC, Bates DL, von Koenig KD, Hardin J, Chen L: Structure of NFAT1 bound as a dimer to the HIV-1 LTR kappa B element. Nat Struct Biol. 2003, 10: 800-806. 10.1038/nsb981.

    CAS  PubMed  Google Scholar 

  128. Bates DL, Barthel KK, Wu Y, Kalhor R, Stroud JC, Giffin MJ, Chen L: Crystal Structure of NFAT Bound to the HIV-1 LTR Tandem kappaB Enhancer Element. Structure. 2008, 16: 684-694. 10.1016/j.str.2008.01.020.

    PubMed Central  CAS  PubMed  Google Scholar 

  129. de Lumley M, Hart DJ, Cooper MA, Symeonides S, Blackburn JM: A biophysical characterisation of factors controlling dimerisation and selectivity in the NF-kappaB and NFAT families. J Mol Biol. 2004, 339: 1059-1075. 10.1016/j.jmb.2004.03.083.

    CAS  PubMed  Google Scholar 

  130. Ranjbar S, Tsytsykova AV, Lee SK, Rajsbaum R, Falvo JV, Lieberman J, Shankar P, Goldfeld AE: NFAT5 regulates HIV-1 in primary monocytes via a highly conserved long terminal repeat site. PLoS Pathog. 2006, 2: e130-10.1371/journal.ppat.0020130.

    PubMed Central  PubMed  Google Scholar 

  131. Markovitz DM, Hannibal MC, Smith MJ, Cossman R, Nabel GJ: Activation of the human immunodeficiency virus type 1 enhancer is not dependent on NFAT-1. J Virol. 1992, 66: 3961-3965.

    PubMed Central  CAS  PubMed  Google Scholar 

  132. Lu YC, Touzjian N, Stenzel M, Dorfman T, Sodroski JG, Haseltine WA: Identification of cis-acting repressive sequences within the negative regulatory element of human immunodeficiency virus type 1. J Virol. 1990, 64: 5226-5229.

    PubMed Central  CAS  PubMed  Google Scholar 

  133. Kinoshita S, Chen BK, Kaneshima H, Nolan GP: Host control of HIV-1 parasitism in T cells by the nuclear factor of activated T cells. Cell. 1998, 95: 595-604. 10.1016/S0092-8674(00)81630-X.

    CAS  PubMed  Google Scholar 

  134. Manley K, O'Hara BA, Gee GV, Simkevich CP, Sedivy JM, Atwood WJ: NFAT4 is required for JC virus infection of glial cells. J Virol. 2006, 80: 12079-12085. 10.1128/JVI.01456-06.

    PubMed Central  CAS  PubMed  Google Scholar 

  135. Sica A, Dorman L, Viggiano V, Cippitelli M, Ghosh P, Rice N, Young HA: Interaction of NF-kappaB and NFAT with the interferon-gamma promoter. J Biol Chem. 1997, 272: 30412-30420. 10.1074/jbc.272.48.30412.

    CAS  PubMed  Google Scholar 

  136. Romanchikova N, Ivanova V, Scheller C, Jankevics E, Jassoy C, Serfling E: NFAT transcription factors control HIV-1 expression through a binding site downstream of TAR region. Immunobiology. 2003, 208: 361-365. 10.1078/0171-2985-00283.

    CAS  PubMed  Google Scholar 

  137. Macian F, Rao A: Reciprocal modulatory interaction between human immunodeficiency virus type 1 Tat and transcription factor NFAT1. Mol Cell Biol. 1999, 19: 3645-3653.

    PubMed Central  CAS  PubMed  Google Scholar 

  138. Pessler F, Cron RQ: Reciprocal regulation of the nuclear factor of activated T cells and HIV-1. Genes Immun. 2004, 5: 158-167. 10.1038/sj.gene.6364047.

    CAS  PubMed  Google Scholar 

  139. Cron RQ: HIV-1, NFAT, and cyclosporin: immunosuppression for the immunosuppressed?. DNA Cell Biol. 2001, 20: 761-767. 10.1089/104454901753438570.

    CAS  PubMed  Google Scholar 

  140. Pereira LA, Bentley K, Peeters A, Churchill MJ, Deacon NJ: A 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.

    PubMed Central  CAS  PubMed  Google Scholar 

  141. Ilyinskii PO, Daniel MD, Simon MA, Lackner AA, Desrosiers RC: The role of upstream U3 sequences in the pathogenesis of simian immunodeficiency virus-induced AIDS in rhesus monkeys. J Virol. 1994, 68: 5933-5944.

    PubMed Central  CAS  PubMed  Google Scholar 

  142. Kirchhoff F, Kestler HW, Desrosiers RC: Upstream U3 sequences in simian immunodeficiency virus are selectively deleted in vivo in the absence of an intact nef gene. J Virol. 1994, 68: 2031-2037.

    PubMed Central  CAS  PubMed  Google Scholar 

  143. Pohlmann S, Floss S, Ilyinskii PO, Stamminger T, Kirchhoff F: Sequences just upstream of the simian immunodeficiency virus core enhancer allow efficient replication in the absence of NF-kappaB and Sp1 binding elements. J Virol. 1998, 72: 5589-5598.

    PubMed Central  CAS  PubMed  Google Scholar 

  144. Rosen CA, Sodroski JG, Haseltine WA: The location of cis-acting regulatory sequences in the human T cell lymphotropic virus type III (HTLV-III/LAV) long terminal repeat. Cell. 1985, 41: 813-823. 10.1016/S0092-8674(85)80062-3.

    CAS  PubMed  Google Scholar 

  145. Nonnemacher MR, Hogan TH, Quiterio S, Wigdahl B, Henderson A, Krebs FC: Identification of binding sites for members of the CCAAT/enhancer binding protein transcription factor family in the simian immunodeficiency virus long terminal repeat. Biomed Pharmacother. 2003, 57: 34-40. 10.1016/S0753-3322(02)00334-7.

    CAS  PubMed  Google Scholar 

  146. Ross HL, Nonnemacher MR, Hogan TH, Quiterio SJ, Henderson A, McAllister JJ, Krebs FC, Wigdahl B: Interaction between CCAAT/enhancer binding protein and cyclic AMP response element binding protein 1 regulates human immunodeficiency virus type 1 transcription in cells of the monocyte/macrophage lineage. J Virol. 2001, 75: 1842-1856. 10.1128/JVI.75.4.1842-1856.2001.

    PubMed Central  CAS  PubMed  Google Scholar 

  147. Quiterio S, Grant C, Hogan TH, Krebs FC, Wigdahl B: C/EBP- and Tat-mediated activation of the HIV-1 LTR in CD34+ hematopoietic progenitor cells. Biomed Pharmacother. 2003, 57: 49-56. 10.1016/S0753-3322(02)00332-3.

    CAS  PubMed  Google Scholar 

  148. Deppmann CD, Alvania RS, Taparowsky EJ: Cross-species annotation of basic leucine zipper factor interactions: Insight into the evolution of closed interaction networks. Mol Biol Evol. 2006, 23: 1480-1492. 10.1093/molbev/msl022.

    CAS  PubMed  Google Scholar 

  149. Landschulz WH, Johnson PF, McKnight SL: The leucine zipper: a hypothetical structure common to a new class of DNA binding proteins. Science. 1988, 240: 1759-1764. 10.1126/science.3289117.

    CAS  PubMed  Google Scholar 

  150. Cai DH, Wang D, Keefer J, Yeamans C, Hensley K, Friedman AD: C/EBP alpha:AP-1 leucine zipper heterodimers bind novel DNA elements, activate the PU.1 promoter and direct monocyte lineage commitment more potently than C/EBP alpha homodimers or AP-1. Oncogene. 2008, 27: 2772-2779. 10.1038/sj.onc.1210940.

    PubMed Central  CAS  PubMed  Google Scholar 

  151. Hai T, Curran T: Cross-family dimerization of transcription factors Fos/Jun and ATF/CREB alters DNA binding specificity. Proc Natl Acad Sci USA. 1991, 88: 3720-3724. 10.1073/pnas.88.9.3720.

    PubMed Central  CAS  PubMed  Google Scholar 

  152. Gombart AF, Grewal J, Koeffler HP: ATF4 differentially regulates transcriptional activation of myeloid-specific genes by C/EBPepsilon and C/EBPalpha. J Leukoc Biol. 2007, 81: 1535-1547. 10.1189/jlb.0806516.

    CAS  PubMed  Google Scholar 

  153. Hsu W, Kerppola TK, Chen PL, Curran T, Chen-Kiang S: Fos and Jun repress transcription activation by NF-IL6 through association at the basic zipper region. Mol Cell Biol. 1994, 14: 268-276.

    PubMed Central  CAS  PubMed  Google Scholar 

  154. Rangatia J, Vangala RK, Treiber N, Zhang P, Radomska H, Tenen DG, Hiddemann W, Behre G: Downregulation of c-Jun expression by transcription factor C/EBPalpha is critical for granulocytic lineage commitment. Mol Cell Biol. 2002, 22: 8681-8694. 10.1128/MCB.22.24.8681-8694.2002.

    PubMed Central  CAS  PubMed  Google Scholar 

  155. Nolan GP: NF-AT-AP-1 and Rel-bZIP: hybrid vigor and binding under the influence. Cell. 1994, 77: 795-798. 10.1016/0092-8674(94)90126-0.

    CAS  PubMed  Google Scholar 

  156. Mondal D, Alam J, Prakash O: NF-kappa B site-mediated negative regulation of the HIV-1 promoter by CCAAT/enhancer binding proteins in brain-derived cells. J Mol Neurosci. 1994, 5: 241-258. 10.1007/BF02736725.

    CAS  PubMed  Google Scholar 

  157. Tesmer VM, Rajadhyaksha A, Babin J, Bina M: NF-IL6-mediated transcriptional activation of the long terminal repeat of the human immunodeficiency virus type 1. Proc Natl Acad Sci USA. 1993, 90: 7298-7302. 10.1073/pnas.90.15.7298.

    PubMed Central  CAS  PubMed  Google Scholar 

  158. Henderson AJ, Zou X, Calame KL: C/EBP proteins activate transcription from the human immunodeficiency virus type 1 long terminal repeat in macrophages/monocytes. J Virol. 1995, 69: 5337-5344.

    PubMed Central  CAS  PubMed  Google Scholar 

  159. Henderson AJ, Connor RI, Calame KL: C/EBP activators are required for HIV-1 replication and proviral induction in monocytic cell lines. Immunity. 1996, 5: 91-101. 10.1016/S1074-7613(00)80313-1.

    CAS  PubMed  Google Scholar 

  160. Yang Y, Pares-Matos EI, Tesmer VM, Dai C, Ashworth S, Huai J, Bina M: Organization of the promoter region of the human NF-IL6 gene. Biochim Biophys Acta. 2002, 1577: 102-108.

    CAS  PubMed  Google Scholar 

  161. Tesmer VM, Bina M: Regulation of HIV-1 gene expression by NF-IL6. J Mol Biol. 1996, 262: 327-335. 10.1006/jmbi.1996.0516.

    CAS  PubMed  Google Scholar 

  162. Kinoshita SM, Taguchi S: NF-IL6 (C/EBPbeta) induces HIV-1 replication by inhibiting cytidine deaminase APOBEC3G. Proc Natl Acad Sci USA. 2008, 105: 15022-15027. 10.1073/pnas.0807269105.

    PubMed Central  CAS  PubMed  Google Scholar 

  163. Landschulz WH, Johnson PF, Adashi EY, Graves BJ, McKnight SL: Isolation of a recombinant copy of the gene encoding C/EBP. Genes Dev. 1988, 2: 786-800. 10.1101/gad.2.7.786.

    CAS  PubMed  Google Scholar 

  164. Akira S, Isshiki H, Sugita T, Tanabe O, Kinoshita S, Nishio Y, Nakajima T, Hirano T, Kishimoto T: A nuclear factor for IL-6 expression (NF-IL6) is a member of a C/EBP family. EMBO J. 1990, 9: 1897-1906.

    PubMed Central  CAS  PubMed  Google Scholar 

  165. Chang CJ, Chen TT, Lei HY, Chen DS, Lee SC: Molecular cloning of a transcription factor, AGP/EBP, that belongs to members of the C/EBP family. Mol Cell Biol. 1990, 10: 6642-6653.

    PubMed Central  CAS  PubMed  Google Scholar 

  166. Antonson P, Stellan B, Yamanaka R, Xanthopoulos KG: A novel human CCAAT/enhancer binding protein gene, C/EBPepsilon, is expressed in cells of lymphoid and myeloid lineages and is localized on chromosome 14q11.2 close to the T-cell receptor alpha/delta locus. Genomics. 1996, 35: 30-38. 10.1006/geno.1996.0319.

    CAS  PubMed  Google Scholar 

  167. Ron D, Habener JF: CHOP, a novel developmentally regulated nuclear protein that dimerizes with transcription factors C/EBP and LAP and functions as a dominant-negative inhibitor of gene transcription. Genes Dev. 1992, 6: 439-453. 10.1101/gad.6.3.439.

    CAS  PubMed  Google Scholar 

  168. Cooper C, Henderson A, Artandi S, Avitahl N, Calame K: Ig/EBP (C/EBP gamma) is a transdominant negative inhibitor of C/EBP family transcriptional activators. Nucleic Acids Res. 1995, 23: 4371-4377. 10.1093/nar/23.21.4371.

    PubMed Central  CAS  PubMed  Google Scholar 

  169. Poli V, Mancini FP, Cortese R: IL-6DBP, a nuclear protein involved in interleukin-6 signal transduction, defines a new family of leucine zipper proteins related to C/EBP. Cell. 1990, 63: 643-653. 10.1016/0092-8674(90)90459-R.

    CAS  PubMed  Google Scholar 

  170. Descombes P, Schibler U: A liver-enriched transcriptional activator protein, LAP, and a transcriptional inhibitory protein, LIP, are translated from the same mRNA. Cell. 1991, 67: 569-579. 10.1016/0092-8674(91)90531-3.

    CAS  PubMed  Google Scholar 

  171. Eaton E, Hanlon M, Bundy L, Sealy L: Characterization of C/EBPbeta isoforms in normal versus neoplastic mammary epithelial cells. J Cell Physiol. 2001, 189 (1): 91-105. 10.1002/jcp.1139.

    CAS  PubMed  Google Scholar 

  172. Scott LM, Civin CI, Rorth P, Friedman AD: A novel temporal expression pattern of three C/EBP family members in differentiating myelomonocytic cells. Blood. 1992, 80: 1725-1735.

    CAS  PubMed  Google Scholar 

  173. Natsuka S, Akira S, Nishio Y, Hashimoto S, Sugita T, Isshiki H, Kishimoto T: Macrophage differentiation-specific expression of NF-IL6, a transcription factor for interleukin-6. Blood. 1992, 79: 460-466.

    CAS  PubMed  Google Scholar 

  174. Tengku-Muhammad TS, Hughes TR, Ranki H, Cryer A, Ramji DP: Differential regulation of macrophage CCAAT-enhancer binding protein isoforms by lipopolysaccharide and cytokines. Cytokine. 2000, 12: 1430-1436. 10.1006/cyto.2000.0711.

    CAS  PubMed  Google Scholar 

  175. Williams SC, Baer M, Dillner AJ, Johnson PF: CRP2 (C/EBP beta) contains a bipartite regulatory domain that controls transcriptional activation, DNA binding and cell specificity. EMBO J. 1995, 14: 3170-3183.

    PubMed Central  CAS  PubMed  Google Scholar 

  176. Kowenz-Leutz E, Twamley G, Ansieau S, Leutz A: Novel mechanism of C/EBP beta (NF-M) transcriptional control: activation through derepression. Genes Dev. 1994, 8: 2781-2791. 10.1101/gad.8.22.2781.

    CAS  PubMed  Google Scholar 

  177. Nakajima T, Kinoshita S, Sasagawa T, Sasaki K, Naruto M, Kishimoto T, Akira S: Phosphorylation at threonine-235 by a ras-dependent mitogen-activated protein kinase cascade is essential for transcription factor NF-IL6. Proc Natl Acad Sci USA. 1993, 90: 2207-2211. 10.1073/pnas.90.6.2207.

    PubMed Central  CAS  PubMed  Google Scholar 

  178. Chinery R, Brockman JA, Dransfield DT, Coffey RJ: Antioxidant-induced nuclear translocation of CCAAT/enhancer-binding protein beta. A critical role for protein kinase A-mediated phosphorylation of Ser299. J Biol Chem. 1997, 272: 30356-30361. 10.1074/jbc.272.48.30356.

    CAS  PubMed  Google Scholar 

  179. Mameli G, Deshmane SL, Ghafouri M, Cui J, Simbiri K, Khalili K, Mukerjee R, Dolei A, Amini S, Sawaya BE: C/EBPbeta regulates human immunodeficiency virus 1 gene expression through its association with cdk9. J Gen Virol. 2007, 88: 631-640. 10.1099/vir.0.82487-0.

    CAS  PubMed  Google Scholar 

  180. Kowenz-Leutz E, Leutz A: A C/EBP beta isoform recruits the SWI/SNF complex to activate myeloid genes. Mol Cell. 1999, 4: 735-743. 10.1016/S1097-2765(00)80384-6.

    CAS  PubMed  Google Scholar 

  181. Lee ES, Sarma D, Zhou H, Henderson AJ: CCAAT/enhancer binding proteins are not required for HIV-1 entry but regulate proviral transcription by recruiting coactivators to the long-terminal repeat in monocytic cells. Virology. 2002, 299: 20-31. 10.1006/viro.2002.1500.

    CAS  PubMed  Google Scholar 

  182. Wang H, Larris B, Peiris TH, Zhang L, Le Lay J, Gao Y, Greenbaum LE: C/EBPbeta activates E2F-regulated genes in vivo via recruitment of the coactivator CREB-binding protein/P300. J Biol Chem. 2007, 282: 24679-24688. 10.1074/jbc.M705066200.

    CAS  PubMed  Google Scholar 

  183. Mink S, Haenig B, Klempnauer KH: Interaction and functional collaboration of p300 and C/EBPbeta. Mol Cell Biol. 1997, 17: 6609-6617.

    PubMed Central  CAS  PubMed  Google Scholar 

  184. Schwartz C, Beck K, Mink S, Schmolke M, Budde B, Wenning D, Klempnauer KH: Recruitment of p300 by C/EBPbeta triggers phosphorylation of p300 and modulates coactivator activity. EMBO J. 2003, 22: 882-892. 10.1093/emboj/cdg076.

    PubMed Central  CAS  PubMed  Google Scholar 

  185. Schwartz C, Catez P, Rohr O, Lecestre D, Aunis D, Schaeffer E: Functional interactions between C/EBP, Sp1, and COUP-TF regulate human immunodeficiency virus type 1 gene transcription in human brain cells. J Virol. 2000, 74: 65-73. 10.1128/JVI.74.1.65-73.2000.

    PubMed Central  CAS  PubMed  Google Scholar 

  186. Hogan TH, Krebs FC, Wigdahl B: Regulation of human immunodeficiency virus type 1 gene expression and pathogenesis by CCAAT/enhancer binding proteins in cells of the monocyte/macrophage lineage. J Neurovirol. 2002, 8 (Suppl 2): 21-26. 10.1080/13550280290167911.

    CAS  PubMed  Google Scholar 

  187. Krebs FC, Goodenow MM, Wigdahl B: Neuroglial ATF/CREB factors interact with the human immunodeficiency virus type 1 long terminal repeat. J Neurovirol. 1997, 3 (Suppl 1): S28-32.

    PubMed  Google Scholar 

  188. Rabbi MF, Saifuddin M, Gu DS, Kagnoff MF, Roebuck KA: U5 region of the human immunodeficiency virus type 1 long terminal repeat contains TRE-like cAMP-responsive elements that bind both AP-1 and CREB/ATF proteins. Virology. 1997, 233: 235-245. 10.1006/viro.1997.8602.

    CAS  PubMed  Google Scholar 

  189. Roebuck KA, Brenner DA, Kagnoff MF: Identification of c-fos-responsive elements downstream of TAR in the long terminal repeat of human immunodeficiency virus type-1. J Clin Invest. 1993, 92: 1336-1348. 10.1172/JCI116707.

    PubMed Central  CAS  PubMed  Google Scholar 

  190. Jia S, Noma K, Grewal SI: RNAi-independent heterochromatin nucleation by the stress-activated ATF/CREB family proteins. Science. 2004, 304: 1971-1976. 10.1126/science.1099035.

    CAS  PubMed  Google Scholar 

  191. Hess J, Angel P, Schorpp-Kistner M: AP-1 subunits: quarrel and harmony among siblings. J Cell Sci. 2004, 117: 5965-5973. 10.1242/jcs.01589.

    CAS  PubMed  Google Scholar 

  192. Chen P, Flory E, Avots A, Jordan BW, Kirchhoff F, Ludwig S, Rapp UR: Transactivation of naturally occurring HIV-1 long terminal repeats by the JNK signaling pathway. The most frequent naturally occurring length polymorphism sequence introduces a novel binding site for AP-1 factors. J Biol Chem. 2000, 275: 20382-20390. 10.1074/jbc.M001149200.

    CAS  PubMed  Google Scholar 

  193. Eferl R, Wagner EF: AP-1: a double-edged sword in tumorigenesis. Nat Rev Cancer. 2003, 3: 859-868. 10.1038/nrc1209.

    CAS  PubMed  Google Scholar 

  194. Ito T, Yamauchi M, Nishina M, Yamamichi N, Mizutani T, Ui M, Murakami M, Iba H: Identification of SWI.SNF complex subunit BAF60a as a determinant of the transactivation potential of Fos/Jun dimers. J Biol Chem. 2001, 276: 2852-2857. 10.1074/jbc.M009633200.

    CAS  PubMed  Google Scholar 

  195. Herschman HR: Primary response genes induced by growth factors and tumor promoters. Annu Rev Biochem. 1991, 60: 281-319. 10.1146/annurev.bi.60.070191.001433.

    CAS  PubMed  Google Scholar 

  196. Datta R, Sherman ML, Stone RM, Kufe D: Expression of the jun-B gene during induction of monocytic differentiation. Cell Growth Differ. 1991, 2: 43-49.

    CAS  PubMed  Google Scholar 

  197. Sherman ML, Stone RM, Datta R, Bernstein SH, Kufe DW: Transcriptional and post-transcriptional regulation of c-jun expression during monocytic differentiation of human myeloid leukemic cells. J Biol Chem. 1990, 265: 3320-3323.

    CAS  PubMed  Google Scholar 

  198. Lord KA, Abdollahi A, Hoffman-Liebermann B, Liebermann DA: Proto-oncogenes of the fos/jun family of transcription factors are positive regulators of myeloid differentiation. Mol Cell Biol. 1993, 13: 841-851.

    PubMed Central  CAS  PubMed  Google Scholar 

  199. Szabo E, Preis LH, Birrer MJ: Constitutive cJun expression induces partial macrophage differentiation in U-937 cells. Cell Growth Differ. 1994, 5: 439-446.

    CAS  PubMed  Google Scholar 

  200. Li J, King I, Sartorelli AC: Differentiation of WEHI-3B D+ myelomonocytic leukemia cells induced by ectopic expression of the protooncogene c-jun. Cell Growth Differ. 1994, 5: 743-751.

    CAS  PubMed  Google Scholar 

  201. Xanthoudakis S, Miao G, Wang F, Pan YC, Curran T: Redox activation of Fos-Jun DNA binding activity is mediated by a DNA repair enzyme. Embo J. 1992, 11: 3323-3335.

    PubMed Central  CAS  PubMed  Google Scholar 

  202. Monick MM, Carter AB, Hunninghake GW: Human alveolar macrophages are markedly deficient in REF-1 and AP-1 DNA binding activity. J Biol Chem. 1999, 274: 18075-18080. 10.1074/jbc.274.25.18075.

    CAS  PubMed  Google Scholar 

  203. Hirota K, Matsui M, Iwata S, Nishiyama A, Mori K, Yodoi J: AP-1 transcriptional activity is regulated by a direct association between thioredoxin and Ref-1. Proc Natl Acad Sci USA. 1997, 94: 3633-3638. 10.1073/pnas.94.8.3633.

    PubMed Central  CAS  PubMed  Google Scholar 

  204. Bossis G, Malnou CE, Farras R, Andermarcher E, Hipskind R, Rodriguez M, Schmidt D, Muller S, Jariel-Encontre I, Piechaczyk M: Down-regulation of c-Fos/c-Jun AP-1 dimer activity by sumoylation. Mol Cell Biol. 2005, 25: 6964-6979. 10.1128/MCB.25.16.6964-6979.2005.

    PubMed Central  CAS  PubMed  Google Scholar 

  205. Garaude J, Farras R, Bossis G, Charni S, Piechaczyk M, Hipskind RA, Villalba M: SUMOylation regulates the transcriptional activity of JunB in T lymphocytes. J Immunol. 2008, 180: 5983-5990.

    CAS  PubMed  Google Scholar 

  206. Boyle WJ, Smeal T, Defize LH, Angel P, Woodgett JR, Karin M, Hunter T: Activation of protein kinase C decreases phosphorylation of c-Jun at sites that negatively regulate its DNA-binding activity. Cell. 1991, 64: 573-584. 10.1016/0092-8674(91)90241-P.

    CAS  PubMed  Google Scholar 

  207. Friedman AD: C/EBPalpha induces PU.1 and interacts with AP-1 and NF-kappaB to regulate myeloid development. Blood Cells Mol Dis. 2007, 39: 340-343. 10.1016/j.bcmd.2007.06.010.

    PubMed Central  CAS  PubMed  Google Scholar 

  208. Davis RJ: The mitogen-activated protein kinase signal transduction pathway. J Biol Chem. 1993, 268: 14553-14556.

    CAS  PubMed  Google Scholar 

  209. Seger R, Krebs EG: The MAPK signaling cascade. Faseb J. 1995, 9: 726-735.

    CAS  PubMed  Google Scholar 

  210. Yang X, Chen Y, Gabuzda D: ERK MAP kinase links cytokine signals to activation of latent HIV-1 infection by stimulating a cooperative interaction of AP-1 and NF-kappaB. J Biol Chem. 1999, 274: 27981-27988. 10.1074/jbc.274.39.27981.

    CAS  PubMed  Google Scholar 

  211. Leppa S, Saffrich R, Ansorge W, Bohmann D: Differential regulation of c-Jun by ERK and JNK during PC12 cell differentiation. Embo J. 1998, 17: 4404-4413. 10.1093/emboj/17.15.4404.

    PubMed Central  CAS  PubMed  Google Scholar 

  212. Karin M: The regulation of AP-1 activity by mitogen-activated protein kinases. J Biol Chem. 1995, 270: 16483-16486.

    CAS  PubMed  Google Scholar 

  213. Stein B, Baldwin AS, Ballard DW, Greene WC, Angel P, Herrlich P: Cross-coupling of the NF-kappa B p65 and Fos/Jun transcription factors produces potentiated biological function. Embo J. 1993, 12: 3879-3891.

    PubMed Central  CAS  PubMed  Google Scholar 

  214. Biggs TE, Cooke SJ, Barton CH, Harris MP, Saksela K, Mann DA: Induction of activator protein 1 (AP-1) in macrophages by human immunodeficiency virus type-1 NEF is a cell-type-specific response that requires both hck and MAPK signaling events. J Mol Biol. 1999, 290: 21-35. 10.1006/jmbi.1999.2849.

    CAS  PubMed  Google Scholar 

  215. Berkhout B, Jeang KT: trans activation of human immunodeficiency virus type 1 is sequence specific for both the single-stranded bulge and loop of the trans-acting-responsive hairpin: a quantitative analysis. J Virol. 1989, 63: 5501-5504.

    PubMed Central  CAS  PubMed  Google Scholar 

  216. Feng S, Holland EC: HIV-1 tat trans-activation requires the loop sequence within tar. Nature. 1988, 334: 165-167. 10.1038/334165a0.

    CAS  PubMed  Google Scholar 

  217. Herrmann CH, Rice AP: Lentivirus Tat proteins specifically associate with a cellular protein kinase, TAK, that hyperphosphorylates the carboxyl-terminal domain of the large subunit of RNA polymerase II: candidate for a Tat cofactor. J Virol. 1995, 69: 1612-1620.

    PubMed Central  CAS  PubMed  Google Scholar 

  218. Yang X, Gold MO, Tang DN, Lewis DE, Aguilar-Cordova E, Rice AP, Herrmann CH: TAK, an HIV Tat-associated kinase, is a member of the cyclin-dependent family of protein kinases and is induced by activation of peripheral blood lymphocytes and differentiation of promonocytic cell lines. Proc Natl Acad Sci USA. 1997, 94: 12331-12336. 10.1073/pnas.94.23.12331.

    PubMed Central  CAS  PubMed  Google Scholar 

  219. Kao SY, Calman 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.

    CAS  PubMed  Google Scholar 

  220. Laspia MF, Rice AP, Mathews MB: HIV-1 Tat protein increases transcriptional initiation and stabilizes elongation. Cell. 1989, 59: 283-292. 10.1016/0092-8674(89)90290-0.

    CAS  PubMed  Google Scholar 

  221. Raha T, Cheng SW, Green MR: HIV-1 Tat stimulates transcription complex assembly through recruitment of TBP in the absence of TAFs. PLoS Biol. 2005, 3: e44-10.1371/journal.pbio.0030044.

    PubMed Central  PubMed  Google Scholar 

  222. Wang Y, Liu XY, De Clercq E: Role of the HIV-1 positive elongation factor P-TEFb and inhibitors thereof. Mini Rev Med Chem. 2009, 9: 379-385. 10.2174/138955709789055207.

    CAS  PubMed  Google Scholar 

  223. D'Orso I, Frankel AD: Tat acetylation modulates assembly of a viral-host RNA-protein transcription complex. Proc Natl Acad Sci USA. 2009, 106: 3101-3106. 10.1073/pnas.0900012106.

    PubMed Central  PubMed  Google Scholar 

  224. Molle D, Maiuri P, Boireau S, Bertrand E, Knezevich A, Marcello A, Basyuk E: A real-time view of the TAR:Tat:P-TEFb complex at HIV-1 transcription sites. Retrovirology. 2007, 4: 36-10.1186/1742-4690-4-36.

    PubMed Central  PubMed  Google Scholar 

  225. Kimura A, Horikoshi M: Tip60 acetylates six lysines of a specific class in core histones in vitro. Genes Cells. 1998, 3: 789-800. 10.1046/j.1365-2443.1998.00229.x.

    CAS  PubMed  Google Scholar 

  226. 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.

    CAS  PubMed  Google Scholar 

  227. Vardabasso C, Manganaro L, Lusic M, Marcello A, Giacca M: The histone chaperone protein Nucleosome Assembly Protein-1 (hNAP-1) binds HIV-1 Tat and promotes viral transcription. Retrovirology. 2008, 5: 8-10.1186/1742-4690-5-8.

    PubMed Central  PubMed  Google Scholar 

  228. 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.

    PubMed Central  CAS  PubMed  Google Scholar 

  229. Deng L, de la Fuente C, Fu P, Wang L, Donnelly R, Wade JD, Lambert P, Li H, Lee CG, Kashanchi F: Acetylation of HIV-1 Tat by CBP/P300 increases transcription of integrated HIV-1 genome and enhances binding to core histones. Virology. 2000, 277: 278-295. 10.1006/viro.2000.0593.

    CAS  PubMed  Google Scholar 

  230. 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.

    CAS  PubMed  Google Scholar 

  231. Treand C, du Chene I, Bres V, Kiernan R, Benarous R, Benkirane M, Emiliani S: Requirement for SWI/SNF chromatin-remodeling complex in Tat-mediated activation of the HIV-1 promoter. Embo J. 2006, 25: 1690-1699. 10.1038/sj.emboj.7601074.

    PubMed Central  CAS  PubMed  Google Scholar 

  232. Harrich D, Garcia J, Wu F, Mitsuyasu R, Gonazalez J, Gaynor R: Role of SP1-binding domains in in vivo transcriptional regulation of the human immunodeficiency virus type 1 long terminal repeat. J Virol. 1989, 63: 2585-2591.

    PubMed Central  CAS  PubMed  Google Scholar 

  233. Berkhout B, Jeang KT: Functional roles for the TATA promoter and enhancers in basal and Tat-induced expression of the human immunodeficiency virus type 1 long terminal repeat. J Virol. 1992, 66: 139-149.

    PubMed Central  CAS  PubMed  Google Scholar 

  234. Jeang KT, Chun R, Lin NH, Gatignol A, Glabe CG, Fan H: In vitro and in vivo binding of human immunodeficiency virus type 1 Tat protein and Sp1 transcription factor. J Virol. 1993, 67: 6224-6233.

    PubMed Central  CAS  PubMed  Google Scholar 

  235. Loregian A, Bortolozzo K, Boso S, Caputo A, Palu G: Interaction of Sp1 transcription factor with HIV-1 Tat protein: looking for cellular partners. FEBS Lett. 2003, 543: 61-65. 10.1016/S0014-5793(03)00399-5.

    CAS  PubMed  Google Scholar 

  236. Loregian A, Bortolozzo K, Boso S, Sapino B, Betti M, Biasolo MA, Caputo A, Palu G: The Sp1 transcription factor does not directly interact with the HIV-1 Tat protein. J Cell Physiol. 2003, 196: 251-257. 10.1002/jcp.10271.

    CAS  PubMed  Google Scholar 

  237. Hidalgo-Estevez AM, Gonzalez E, Punzon C, Fresno M: Human immunodeficiency virus type 1 Tat increases cooperation between AP-1 and NFAT transcription factors in T cells. J Gen Virol. 2006, 87: 1603-1612. 10.1099/vir.0.81637-0.

    CAS  PubMed  Google Scholar 

  238. Biswas DK, Salas TR, Wang F, Ahlers CM, Dezube BJ, Pardee AB: A Tat-induced auto-up-regulatory loop for superactivation of the human immunodeficiency virus type 1 promoter. J Virol. 1995, 69: 7437-7444.

    PubMed Central  CAS  PubMed  Google Scholar 

  239. Demarchi F, Gutierrez MI, Giacca M: Human immunodeficiency virus type 1 tat protein activates transcription factor NF-kappaB through the cellular interferon-inducible, double-stranded RNA-dependent protein kinase, PKR. J Virol. 1999, 73: 7080-7086.

    PubMed Central  CAS  PubMed  Google Scholar 

  240. Kumar A, Manna SK, Dhawan S, Aggarwal BB: HIV-Tat protein activates c-Jun N-terminal kinase and activator protein-1. J Immunol. 1998, 161: 776-781.

    CAS  PubMed  Google Scholar 

  241. Subbramanian RA, Kessous-Elbaz A, Lodge R, Forget J, Yao XJ, Bergeron D, Cohen EA: Human immunodeficiency virus type 1 Vpr is a positive regulator of viral transcription and infectivity in primary human macrophages. J Exp Med. 1998, 187: 1103-1111. 10.1084/jem.187.7.1103.

    PubMed Central  CAS  PubMed  Google Scholar 

  242. Cohen EA, Terwilliger EF, Jalinoos Y, Proulx J, Sodroski JG, Haseltine WA: Identification of HIV-1 vpr product and function. J Acquir Immune Defic Syndr. 1990, 3: 11-18.

    CAS  PubMed  Google Scholar 

  243. Philippon V, Matsuda Z, Essex M: Transactivation is a conserved function among primate lentivirus Vpr proteins but is not shared by Vpx. J Hum Virol. 1999, 2: 167-174.

    CAS  PubMed  Google Scholar 

  244. Sawaya BE, Khalili K, Gordon J, Taube R, Amini S: Cooperative interaction between HIV-1 regulatory proteins Tat and Vpr modulates transcription of the viral genome. J Biol Chem. 2000, 275: 35209-35214. 10.1074/jbc.M005197200.

    CAS  PubMed  Google Scholar 

  245. Vanitharani R, Mahalingam S, Rafaeli Y, Singh SP, Srinivasan A, Weiner DB, Ayyavoo V: HIV-1 Vpr transactivates LTR-directed expression through sequences present within -278 to -176 and increases virus replication in vitro. Virology. 2001, 289: 334-342. 10.1006/viro.2001.1153.

    CAS  PubMed  Google Scholar 

  246. Zhao LJ, Mukherjee S, Narayan O: Biochemical mechanism of HIV-I Vpr function. Specific interaction with a cellular protein. J Biol Chem. 1994, 269: 15577-15582.

    CAS  PubMed  Google Scholar 

  247. Yu XF, Matsuda M, Essex M, Lee TH: Open reading frame vpr of simian immunodeficiency virus encodes a virion-associated protein. J Virol. 1990, 64: 5688-5693.

    PubMed Central  CAS  PubMed  Google Scholar 

  248. Cohen EA, Dehni G, Sodroski JG, Haseltine WA: Human immunodeficiency virus vpr product is a virion-associated regulatory protein. J Virol. 1990, 64: 3097-3099.

    PubMed Central  CAS  PubMed  Google Scholar 

  249. Caly L, Saksena NK, Piller SC, Jans DA: Impaired nuclear import and viral incorporation of Vpr derived from a HIV long-term non-progressor. Retrovirology. 2008, 5: 67-10.1186/1742-4690-5-67.

    PubMed Central  PubMed  Google Scholar 

  250. Emerman M: HIV-1, Vpr and the cell cycle. Curr Biol. 1996, 6: 1096-1103. 10.1016/S0960-9822(02)00676-0.

    CAS  PubMed  Google Scholar 

  251. Balliet JW, Kolson DL, Eiger G, Kim FM, McGann KA, Srinivasan A, Collman R: Distinct effects in primary macrophages and lymphocytes of the human immunodeficiency virus type 1 accessory genes vpr, vpu, and nef: mutational analysis of a primary HIV-1 isolate. Virology. 1994, 200: 623-631. 10.1006/viro.1994.1225.

    CAS  PubMed  Google Scholar 

  252. Balotta C, Lusso P, Crowley R, Gallo RC, Franchini G: Antisense phosphorothioate oligodeoxynucleotides targeted to the vpr gene inhibit human immunodeficiency virus type 1 replication in primary human macrophages. J Virol. 1993, 67: 4409-4414.

    PubMed Central  CAS  PubMed  Google Scholar 

  253. Connor RI, Chen BK, Choe S, Landau NR: Vpr is required for efficient replication of human immunodeficiency virus type-1 in mononuclear phagocytes. Virology. 1995, 206: 935-944. 10.1006/viro.1995.1016.

    CAS  PubMed  Google Scholar 

  254. Levy DN, Refaeli Y, Weiner DB: Extracellular Vpr protein increases cellular permissiveness to human immunodeficiency virus replication and reactivates virus from latency. J Virol. 1995, 69: 1243-1252.

    PubMed Central  CAS  PubMed  Google Scholar 

  255. Hattori N, Michaels F, Fargnoli K, Marcon L, Gallo RC, Franchini G: The human immunodeficiency virus type 2 vpr gene is essential for productive infection of human macrophages. Proc Natl Acad Sci USA. 1990, 87: 8080-8084. 10.1073/pnas.87.20.8080.

    PubMed Central  CAS  PubMed  Google Scholar 

  256. Westervelt P, Henkel T, Trowbridge DB, Orenstein J, Heuser J, Gendelman HE, Ratner L: Dual regulation of silent and productive infection in monocytes by distinct human immunodeficiency virus type 1 determinants. J Virol. 1992, 66: 3925-3931.

    PubMed Central  CAS  PubMed  Google Scholar 

  257. Chen R, Le Rouzic E, Kearney JA, Mansky LM, Benichou S: Vpr-mediated incorporation of UNG2 into HIV-1 particles is required to modulate the virus mutation rate and for replication in macrophages. J Biol Chem. 2004, 279: 28419-28425. 10.1074/jbc.M403875200.

    CAS  PubMed  Google Scholar 

  258. Fouchier RA, Meyer BE, Simon JH, Fischer U, Albright AV, Gonzalez-Scarano F, Malim MH: Interaction of the human immunodeficiency virus type 1 Vpr protein with the nuclear pore complex. J Virol. 1998, 72: 6004-6013.

    PubMed Central  CAS  PubMed  Google Scholar 

  259. Popov S, Rexach M, Ratner L, Blobel G, Bukrinsky M: Viral protein R regulates docking of the HIV-1 preintegration complex to the nuclear pore complex. J Biol Chem. 1998, 273: 13347-13352. 10.1074/jbc.273.21.13347.

    CAS  PubMed  Google Scholar 

  260. Popov S, Rexach M, Zybarth G, Reiling N, Lee MA, Ratner L, Lane CM, Moore MS, Blobel G, Bukrinsky M: Viral protein R regulates nuclear import of the HIV-1 pre-integration complex. EMBO J. 1998, 17: 909-917. 10.1093/emboj/17.4.909.

    PubMed Central  CAS  PubMed  Google Scholar 

  261. Vodicka MA, Koepp DM, Silver PA, Emerman M: HIV-1 Vpr interacts with the nuclear transport pathway to promote macrophage infection. Genes Dev. 1998, 12: 175-185. 10.1101/gad.12.2.175.

    PubMed Central  CAS  PubMed  Google Scholar 

  262. Le Rouzic E, Mousnier A, Rustum C, Stutz F, Hallberg E, Dargemont C, Benichou S: Docking of HIV-1 Vpr to the nuclear envelope is mediated by the interaction with the nucleoporin hCG1. J Biol Chem. 2002, 277: 45091-45098. 10.1074/jbc.M207439200.

    CAS  PubMed  Google Scholar 

  263. Jacquot G, Le Rouzic E, David A, Mazzolini J, Bouchet J, Bouaziz S, Niedergang F, Pancino G, Benichou S: Localization of HIV-1 Vpr to the nuclear envelope: impact on Vpr functions and virus replication in macrophages. Retrovirology. 2007, 4: 84-10.1186/1742-4690-4-84.

    PubMed Central  PubMed  Google Scholar 

  264. Levy DN, Refaeli Y, MacGregor RR, Weiner DB: Serum Vpr regulates productive infection and latency of human immunodeficiency virus type 1. Proc Natl Acad Sci USA. 1994, 91: 10873-10877. 10.1073/pnas.91.23.10873.

    PubMed Central  CAS  PubMed  Google Scholar 

  265. Henklein P, Bruns K, Sherman MP, Tessmer U, Licha K, Kopp J, de Noronha CM, Greene WC, Wray V, Schubert U: Functional and structural characterization of synthetic HIV-1 Vpr that transduces cells, localizes to the nucleus, and induces G2 cell cycle arrest. J Biol Chem. 2000, 275: 32016-32026. 10.1074/jbc.M004044200.

    CAS  PubMed  Google Scholar 

  266. Sherman MP, Schubert U, Williams SA, de Noronha CM, Kreisberg JF, Henklein P, Greene WC: HIV-1 Vpr displays natural protein-transducing properties: implications for viral pathogenesis. Virology. 2002, 302: 95-105. 10.1006/viro.2002.1576.

    CAS  PubMed  Google Scholar 

  267. Varin A, Decrion AZ, Sabbah E, Quivy V, Sire J, Van Lint C, Roques BP, Aggarwal BB, Herbein G: Synthetic Vpr protein activates activator protein-1, c-Jun N-terminal kinase, and NF-kappaB and stimulates HIV-1 transcription in promonocytic cells and primary macrophages. J Biol Chem. 2005, 280: 42557-42567. 10.1074/jbc.M502211200.

    CAS  PubMed  Google Scholar 

  268. Wang L, Mukherjee S, Jia F, Narayan O, Zhao LJ: Interaction of virion protein Vpr of human immunodeficiency virus type 1 with cellular transcription factor Sp1 and trans-activation of viral long terminal repeat. J Biol Chem. 1995, 270: 25564-25569. 10.1074/jbc.270.43.25564.

    CAS  PubMed  Google Scholar 

  269. Agostini I, Navarro JM, Bouhamdan M, Willetts K, Rey F, Spire B, Vigne R, Pomerantz R, Sire J: The HIV-1 Vpr co-activator induces a conformational change in TFIIB. FEBS Lett. 1999, 450: 235-239. 10.1016/S0014-5793(99)00501-3.

    CAS  PubMed  Google Scholar 

  270. Agostini I, Navarro JM, Rey F, Bouhamdan M, Spire B, Vigne R, Sire J: The human immunodeficiency virus type 1 Vpr transactivator: cooperation with promoter-bound activator domains and binding to TFIIB. J Mol Biol. 1996, 261: 599-606. 10.1006/jmbi.1996.0485.

    CAS  PubMed  Google Scholar 

  271. Ayyavoo V, Mahboubi A, Mahalingam S, Ramalingam R, Kudchodkar S, Williams WV, Green DR, Weiner DB: HIV-1 Vpr suppresses immune activation and apoptosis through regulation of nuclear factor kappa B. Nat Med. 1997, 3: 1117-1123. 10.1038/nm1097-1117.

    CAS  PubMed  Google Scholar 

  272. Roux P, Alfieri C, Hrimech M, Cohen EA, Tanner JE: Activation of transcription factors NF-kappaB and NF-IL-6 by human immunodeficiency virus type 1 protein R (Vpr) induces interleukin-8 expression. J Virol. 2000, 74: 4658-4665. 10.1128/JVI.74.10.4658-4665.2000.

    PubMed Central  CAS  PubMed  Google Scholar 

  273. Forget J, Yao XJ, Mercier J, Cohen EA: Human immunodeficiency virus type 1 vpr protein transactivation function: mechanism and identification of domains involved. J Mol Biol. 1998, 284: 915-923. 10.1006/jmbi.1998.2206.

    CAS  PubMed  Google Scholar 

  274. Hrimech M, Yao XJ, Bachand F, Rougeau N, Cohen EA: Human immunodeficiency virus type 1 (HIV-1) Vpr functions as an immediate-early protein during HIV-1 infection. J Virol. 1999, 73: 4101-4109.

    PubMed Central  CAS  PubMed  Google Scholar 

  275. Bednarik DP, Mosca JD, Raj NB: Methylation as a modulator of expression of human immunodeficiency virus. J Virol. 1987, 61: 1253-1257.

    PubMed Central  CAS  PubMed  Google Scholar 

  276. Gutekunst KA, Kashanchi F, Brady JN, Bednarik DP: Transcription of the HIV-1 LTR is regulated by the density of DNA CpG methylation. J Acquir Immune Defic Syndr. 1993, 6: 541-549.

    CAS  PubMed  Google Scholar 

  277. Schulze-Forster K, Gotz F, Wagner H, Kroger H, Simon D: Transcription of HIV1 is inhibited by DNA methylation. Biochem Biophys Res Commun. 1990, 168: 141-147. 10.1016/0006-291X(90)91685-L.

    CAS  PubMed  Google Scholar 

  278. 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 

  279. Marcello A: Latency: the hidden HIV-1 challenge. Retrovirology. 2006, 3: 7-10.1186/1742-4690-3-7.

    PubMed Central  PubMed  Google Scholar 

  280. Singh MK, Pauza CD: Extrachromosomal human immunodeficiency virus type 1 sequences are methylated in latently infected U937 cells. Virology. 1992, 188: 451-458. 10.1016/0042-6822(92)90498-E.

    CAS  PubMed  Google Scholar 

  281. Ishida T, Hamano A, Koiwa T, Watanabe T: 5' long terminal repeat (LTR)-selective methylation of latently infected HIV-1 provirus that is demethylated by reactivation signals. Retrovirology. 2006, 3: 69-10.1186/1742-4690-3-69.

    PubMed Central  PubMed  Google Scholar 

  282. Joel P, Shao W, Pratt K: A nuclear protein with enhanced binding to methylated Sp1 sites in the AIDS virus promoter. Nucleic Acids Res. 1993, 21: 5786-5793. 10.1093/nar/21.24.5786.

    PubMed Central  CAS  PubMed  Google Scholar 

  283. Shao W: Characterization of HMBP-2, a DNA-Binding Protein That Binds to HIV-1 LTR When only One of the Three Sp1 Sites Is Methylated. J Biomed Sci. 1997, 4: 39-46. 10.1007/BF02255592.

    CAS  PubMed  Google Scholar 

  284. 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.

    PubMed Central  CAS  PubMed  Google Scholar 

  285. Ciardi M, Sharief MK, Thompson EJ, Salotti A, Vullo V, Sorice F, Cirelli A: High cerebrospinal fluid and serum levels of tumor necrosis factor-alpha in asymptomatic HIV-1 seropositive individuals. Correlation with interleukin-2 and soluble IL-2 receptor. J Neurol Sci. 1994, 125: 175-179. 10.1016/0022-510X(94)90031-0.

    CAS  PubMed  Google Scholar 

  286. Ownby RL, Kumar AM, Benny Fernandez J, Moleon-Borodowsky I, Gonzalez L, Eisdorfer S, Waldrop-Valverde D, Kumar M: Tumor Necrosis Factor-alpha Levels in HIV-1 Seropositive Injecting Drug Users. J Neuroimmune Pharmacol. 2009

    Google Scholar 

  287. Esser R, Glienke W, von Briesen H, Rubsamen-Waigmann H, Andreesen R: Differential regulation of proinflammatory and hematopoietic cytokines in human macrophages after infection with human immunodeficiency virus. Blood. 1996, 88: 3474-3481.

    CAS  PubMed  Google Scholar 

  288. Hungness ES, Luo GJ, Pritts TA, Sun X, Robb BW, Hershko D, Hasselgren PO: Transcription factors C/EBP-beta and -delta regulate IL-6 production in IL-1beta-stimulated human enterocytes. J Cell Physiol. 2002, 192: 64-70. 10.1002/jcp.10116.

    CAS  PubMed  Google Scholar 

  289. Yang Z, Wara-Aswapati N, Chen C, Tsukada J, Auron PE: NF-IL6 (C/EBPbeta) vigorously activates il1b gene expression via a Spi-1 (PU.1) protein-protein tether. J Biol Chem. 2000, 275: 21272-21277. 10.1074/jbc.M000145200.

    CAS  PubMed  Google Scholar 

  290. Pope R, Mungre S, Liu H, Thimmapaya B: Regulation of TNF-alpha expression in normal macrophages: the role of C/EBPbeta. Cytokine. 2000, 12: 1171-1181. 10.1006/cyto.2000.0691.

    CAS  PubMed  Google Scholar 

  291. Pope RM, Leutz A, Ness SA: C/EBP beta regulation of the tumor necrosis factor alpha gene. J Clin Invest. 1994, 94: 1449-1455. 10.1172/JCI117482.

    PubMed Central  CAS  PubMed  Google Scholar 

  292. Swingler S, Mann A, Jacque J, Brichacek B, Sasseville VG, Williams K, Lackner AA, Janoff EN, Wang R, Fisher D, Stevenson M: HIV-1 Nef mediates lymphocyte chemotaxis and activation by infected macrophages. Nat Med. 1999, 5: 997-103. 10.1038/12433.

    CAS  PubMed  Google Scholar 

  293. Zylla D, Li Y, Bergenstal E, Merrill JD, Douglas SD, Mooney K, Guo CJ, Song L, Ho WZ: CCR5 expression and beta-chemokine production during placental neonatal monocyte differentiation. Pediatr Res. 2003, 53: 853-858. 10.1203/01.PDR.0000059749.82140.4A.

    PubMed Central  CAS  PubMed  Google Scholar 

  294. Naif HM, Li S, Alali M, Sloane A, Wu L, Kelly M, Lynch G, Lloyd A, Cunningham AL: CCR5 expression correlates with susceptibility of maturing monocytes to human immunodeficiency virus type 1 infection. J Virol. 1998, 72: 830-836.

    PubMed Central  CAS  PubMed  Google Scholar 

  295. Arfi V, Riviere L, Jarrosson-Wuilleme L, Goujon C, Rigal D, Darlix JL, Cimarelli A: Characterization of the early steps of infection of primary blood monocytes by human immunodeficiency virus type 1. J Virol. 2008, 82: 6557-6565. 10.1128/JVI.02321-07.

    PubMed Central  CAS  PubMed  Google Scholar 

  296. Triques K, Stevenson M: Characterization of restrictions to human immunodeficiency virus type 1 infection of monocytes. J Virol. 2004, 78: 5523-5527. 10.1128/JVI.78.10.5523-5527.2004.

    PubMed Central  CAS  PubMed  Google Scholar 

  297. Demarchi F, D'Agaro P, Falaschi A, Giacca M: In vivo footprinting analysis of constitutive and inducible protein-DNA interactions at the long terminal repeat of human immunodeficiency virus type 1. J Virol. 1993, 67: 7450-7460.

    PubMed Central  CAS  PubMed  Google Scholar 

  298. Siddappa NB, Venkatramanan M, Venkatesh P, Janki MV, Jayasuryan N, Desai A, Ravi V, Ranga U: Transactivation and signaling functions of Tat are not correlated: biological and immunological characterization of HIV-1 subtype-C Tat protein. Retrovirology. 2006, 3: 53-10.1186/1742-4690-3-53.

    PubMed Central  PubMed  Google Scholar 

  299. Redell MS, Tweardy DJ: Targeting transcription factors for cancer therapy. Curr Pharm Des. 2005, 11: 2873-2887. 10.2174/1381612054546699.

    CAS  PubMed  Google Scholar 

  300. Mora LB, Buettner R, Seigne J, Diaz J, Ahmad N, Garcia R, Bowman T, Falcone R, Fairclough R, Cantor A, Muro-Cacho C, Livingston S, Karras J, Pow-Sang J, Jove R: Constitutive activation of Stat3 in human prostate tumors and cell lines: direct inhibition of Stat3 signaling induces apoptosis of prostate cancer cells. Cancer Res. 2002, 62: 6659-6666.

    CAS  PubMed  Google Scholar 

  301. Grandis JR, Drenning SD, Zeng Q, Watkins SC, Melhem MF, Endo S, Johnson DE, Huang L, He Y, Kim JD: Constitutive activation of Stat3 signaling abrogates apoptosis in squamous cell carcinogenesis in vivo. Proc Natl Acad Sci USA. 2000, 97: 4227-4232. 10.1073/pnas.97.8.4227.

    PubMed Central  CAS  PubMed  Google Scholar 

  302. Jing N, Zhu Q, Yuan P, Li Y, Mao L, Tweardy DJ: Targeting signal transducer and activator of transcription 3 with G-quartet oligonucleotides: a potential novel therapy for head and neck cancer. Mol Cancer Ther. 2006, 5: 279-286. 10.1158/1535-7163.MCT-05-0302.

    CAS  PubMed  Google Scholar 

  303. Jing N, Li Y, Xiong W, Sha W, Jing L, Tweardy DJ: G-quartet oligonucleotides: a new class of signal transducer and activator of transcription 3 inhibitors that suppresses growth of prostate and breast tumors through induction of apoptosis. Cancer Res. 2004, 64: 6603-6609. 10.1158/0008-5472.CAN-03-4041.

    CAS  PubMed  Google Scholar 

  304. Iversen PL, Arora V, Acker AJ, Mason DH, Devi GR: Efficacy of antisense morpholino oligomer targeted to c-myc in prostate cancer xenograft murine model and a Phase I safety study in humans. Clin Cancer Res. 2003, 9: 2510-2519.

    CAS  PubMed  Google Scholar 

  305. Baud V, Karin M: Is NF-kappaB a good target for cancer therapy? Hopes and pitfalls. Nat Rev Drug Discov. 2009, 8: 33-40. 10.1038/nrd2781.

    PubMed Central  CAS  PubMed  Google Scholar 

  306. Karin M: Nuclear factor-kappaB in cancer development and progression. Nature. 2006, 441: 431-436. 10.1038/nature04870.

    CAS  PubMed  Google Scholar 

  307. Guo J, Verma UN, Gaynor RB, Frenkel EP, Becerra CR: Enhanced chemosensitivity to irinotecan by RNA interference-mediated down-regulation of the nuclear factor-kappaB p65 subunit. Clin Cancer Res. 2004, 10: 3333-3341. 10.1158/1078-0432.CCR-03-0366.

    CAS  PubMed  Google Scholar 

  308. Huang YT, Pan SL, Guh JH, Chang YL, Lee FY, Kuo SC, Teng CM: YC-1 suppresses constitutive nuclear factor-kappaB activation and induces apoptosis in human prostate cancer cells. Mol Cancer Ther. 2005, 4: 1628-1635. 10.1158/1535-7163.MCT-05-0090.

    CAS  PubMed  Google Scholar 

  309. Matsumoto G, Namekawa J, Muta M, Nakamura T, Bando H, Tohyama K, Toi M, Umezawa K: Targeting of nuclear factor kappaB Pathways by dehydroxymethylepoxyquinomicin, a novel inhibitor of breast carcinomas: antitumor and antiangiogenic potential in vivo. Clin Cancer Res. 2005, 11: 1287-1293.

    CAS  PubMed  Google Scholar 

  310. Schon M, Wienrich BG, Kneitz S, Sennefelder H, Amschler K, Vohringer V, Weber O, Stiewe T, Ziegelbauer K, Schon MP: KINK-1, a novel small-molecule inhibitor of IKKbeta, and the susceptibility of melanoma cells to antitumoral treatment. J Natl Cancer Inst. 2008, 100: 862-875. 10.1093/jnci/djn174.

    PubMed  Google Scholar 

  311. Cho SJ, Kim JS, Kim JM, Lee JY, Jung HC, Song IS: Simvastatin induces apoptosis in human colon cancer cells and in tumor xenografts, and attenuates colitis-associated colon cancer in mice. Int J Cancer. 2008, 123: 951-957. 10.1002/ijc.23593.

    CAS  PubMed  Google Scholar 

  312. Karin M, Yamamoto Y, Wang QM: The IKK NF-kappa B system: a treasure trove for drug development. Nat Rev Drug Discov. 2004, 3: 17-26. 10.1038/nrd1279.

    CAS  PubMed  Google Scholar 

  313. Hideshima T, Chauhan D, Schlossman R, Richardson P, Anderson KC: The role of tumor necrosis factor alpha in the pathophysiology of human multiple myeloma: therapeutic applications. Oncogene. 2001, 20: 4519-4527. 10.1038/sj.onc.1204623.

    CAS  PubMed  Google Scholar 

  314. Hideshima T, Richardson P, Chauhan D, Palombella VJ, Elliott PJ, Adams J, Anderson KC: The proteasome inhibitor PS-341 inhibits growth, induces apoptosis, and overcomes drug resistance in human multiple myeloma cells. Cancer Res. 2001, 61: 3071-3076.

    CAS  PubMed  Google Scholar 

  315. Mitsiades N, Mitsiades CS, Richardson PG, Poulaki V, Tai YT, Chauhan D, Fanourakis G, Gu X, Bailey C, Joseph M, Libermann TA, Schlossman R, Munshi NC, Hideshima T, Anderson KC: The proteasome inhibitor PS-341 potentiates sensitivity of multiple myeloma cells to conventional chemotherapeutic agents: therapeutic applications. Blood. 2003, 101: 2377-2380. 10.1182/blood-2002-06-1768.

    CAS  PubMed  Google Scholar 

  316. Papineni S, Chintharlapalli S, Abdelrahim M, Lee SO, Burghardt R, Abudayyeh A, Baker C, Herrera L, Safe S: Tolfenamic Acid Inhibits Esophageal Cancer Through Repression of Specificity Proteins and c-Met. Carcinogenesis. 2009

    Google Scholar 

  317. Abdelrahim M, Smith R, Burghardt R, Safe S: Role of Sp proteins in regulation of vascular endothelial growth factor expression and proliferation of pancreatic cancer cells. Cancer Res. 2004, 64: 6740-6749. 10.1158/0008-5472.CAN-04-0713.

    CAS  PubMed  Google Scholar 

  318. Abdelrahim M, Baker CH, Abbruzzese JL, Sheikh-Hamad D, Liu S, Cho SD, Yoon K, Safe S: Regulation of vascular endothelial growth factor receptor-1 expression by specificity proteins 1, 3, and 4 in pancreatic cancer cells. Cancer Res. 2007, 67: 3286-3294. 10.1158/0008-5472.CAN-06-3831.

    CAS  PubMed  Google Scholar 

  319. Abdelrahim M, Baker CH, Abbruzzese JL, Safe S: Tolfenamic acid and pancreatic cancer growth, angiogenesis, and Sp protein degradation. J Natl Cancer Inst. 2006, 98: 855-868. 10.1093/jnci/djj232.

    CAS  PubMed  Google Scholar 

  320. Liu X, Abdelrahim M, Abudayyeh A, Lei P, Safe S: The nonsteroidal anti-inflammatory drug tolfenamic acid inhibits BT474 and SKBR3 breast cancer cell and tumor growth by repressing erbB2 expression. Mol Cancer Ther. 2009

    Google Scholar 

  321. Stadler WM, Vogelzang NJ, Amato R, Sosman J, Taber D, Liebowitz D, Vokes EE: Flavopiridol, a novel cyclin-dependent kinase inhibitor, in metastatic renal cancer: a University of Chicago Phase II Consortium study. J Clin Oncol. 2000, 18: 371-375.

    CAS  PubMed  Google Scholar 

  322. Schwartz GK, Ilson D, Saltz L, O'Reilly E, Tong W, Maslak P, Werner J, Perkins P, Stoltz M, Kelsen D: Phase II study of the cyclin-dependent kinase inhibitor flavopiridol administered to patients with advanced gastric carcinoma. J Clin Oncol. 2001, 19: 1985-1992.

    CAS  PubMed  Google Scholar 

  323. Kouroukis CT, Belch A, Crump M, Eisenhauer E, Gascoyne RD, Meyer R, Lohmann R, Lopez P, Powers J, Turner R, Connors JM: Flavopiridol in untreated or relapsed mantle-cell lymphoma: results of a phase II study of the National Cancer Institute of Canada Clinical Trials Group. J Clin Oncol. 2003, 21: 1740-1745. 10.1200/jco.2003.09.057.

    CAS  PubMed  Google Scholar 

  324. Kim JC, Saha D, Cao Q, Choy H: Enhancement of radiation effects by combined docetaxel and flavopiridol treatment in lung cancer cells. Radiother Oncol. 2004, 71: 213-221. 10.1016/j.radonc.2004.03.006.

    CAS  PubMed  Google Scholar 

  325. Bible KC, Lensing JL, Nelson SA, Lee YK, Reid JM, Ames MM, Isham CR, Piens J, Rubin SL, Rubin J, Kaufmann SH, Atherton PJ, Sloan JA, Daiss MK, Adjei AA, Erlichman C: Phase 1 trial of flavopiridol combined with cisplatin or carboplatin in patients with advanced malignancies with the assessment of pharmacokinetic and pharmacodynamic end points. Clin Cancer Res. 2005, 11: 5935-5941. 10.1158/1078-0432.CCR-04-2566.

    CAS  PubMed  Google Scholar 

  326. Biglione S, Byers SA, Price JP, Nguyen VT, Bensaude O, Price DH, Maury W: Inhibition of HIV-1 replication by P-TEFb inhibitors DRB, seliciclib and flavopiridol correlates with release of free P-TEFb from the large, inactive form of the complex. Retrovirology. 2007, 4: 47-10.1186/1742-4690-4-47.

    PubMed Central  PubMed  Google Scholar 

  327. Suzuki T, Yamamoto N, Nonaka M, Hashimoto Y, Matsuda G, Takeshima SN, Matsuyama M, Igarashi T, Miura T, Tanaka R, Kato S, Aida Y: Inhibition of human immunodeficiency virus type 1 (HIV-1) nuclear import via Vpr-Importin alpha interactions as a novel HIV-1 therapy. Biochem Biophys Res Commun. 2009, 380: 838-843. 10.1016/j.bbrc.2009.01.180.

    CAS  PubMed  Google Scholar 

  328. Watanabe N, Nishihara Y, Yamaguchi T, Koito A, Miyoshi H, Kakeya H, Osada H: Fumagillin suppresses HIV-1 infection of macrophages through the inhibition of Vpr activity. FEBS Lett. 2006, 580: 2598-2602. 10.1016/j.febslet.2006.04.007.

    CAS  PubMed  Google Scholar 

  329. Van Duyne R, Cardenas J, Easley R, Wu W, Kehn-Hall K, Klase Z, Mendez S, Zeng C, Chen H, Saifuddin M, Kashanchi F: Effect of transcription peptide inhibitors on HIV-1 replication. Virology. 2008, 376: 308-322. 10.1016/j.virol.2008.02.036.

    CAS  PubMed  Google Scholar 

  330. Iguchi T, Ishikawa H, Matumoto H, Mizuno M, Goto K, Hamasaki K: Amino disaccharides having an alpha-(1-->4) or a beta-(1-->4) linkage, their synthesis and evaluation as a potential inhibitor for HIV-1 TAR-Tat. Nucleic Acids Symp Ser (Oxf). 2005, 169-170. 10.1093/nass/49.1.169.

    Google Scholar 

  331. Lapidot A, Berchanski A, Borkow G: Insight into the mechanisms of aminoglycoside derivatives interaction with HIV-1 entry steps and viral gene transcription. FEBS J. 2008, 275: 5236-5257. 10.1111/j.1742-4658.2008.06657.x.

    CAS  PubMed  Google Scholar 

  332. Riguet E, Desire J, Boden O, Ludwig V, Gobel M, Bailly C, Decout JL: Neamine dimers targeting the HIV-1 TAR RNA. Bioorg Med Chem Lett. 2005, 15: 4651-4655. 10.1016/j.bmcl.2005.07.082.

    CAS  PubMed  Google Scholar 

  333. Hwang S, Tamilarasu N, Kibler K, Cao H, Ali A, Ping YH, Jeang KT, Rana TM: Discovery of a small molecule Tat-trans-activation-responsive RNA antagonist that potently inhibits human immunodeficiency virus-1 replication. J Biol Chem. 2003, 278: 39092-39103. 10.1074/jbc.M301749200.

    CAS  PubMed  Google Scholar 

Download references

Acknowledgements

These studies were funded in part by the Public Health Service, National Institutes of Health through grants (B. Wigdahl, Principal Investigator) from the National Institute of Neurological Disorders and Stroke, NS32092 and NS46263, and the National Institute of Drug Abuse, DA19807.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Brian Wigdahl.

Additional information

Competing interests

The authors declare that they have no competing interests.

Authors' contributions

EK was responsible for drafting and revising the manuscript as well as organizing the content. SS created Figures 1, 2, 3, 4, 5, 6, 7 and their legends and proofread the final version of the manuscript for content and consistency. MN drafted portions of the manuscript, assisted in the conceptualization of the figures, and proofread and edited the final version of the manuscript. BW assisted in all aspects of each phase of development from initial concept, through revisions to final approval of the version to be published.

Authors’ original submitted files for images

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

Kilareski, E.M., Shah, S., Nonnemacher, M.R. et al. Regulation of HIV-1 transcription in cells of the monocyte-macrophage lineage. Retrovirology 6, 118 (2009). https://doi.org/10.1186/1742-4690-6-118

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/1742-4690-6-118

Keywords