Skip to main content
Advertisement
  • Loading metrics

Chemokine Coreceptor Signaling in HIV-1 Infection and Pathogenesis

  • Yuntao Wu ,

    ywu8@gmu.edu

    Affiliation Department of Molecular and Microbiology, George Mason University, Manassas, Virginia, United States of America

  • Alyson Yoder

    Affiliation Department of Molecular and Microbiology, George Mason University, Manassas, Virginia, United States of America

Abstract

Binding of the HIV-1 envelope to its chemokine coreceptors mediates two major biological events: membrane fusion and signaling transduction. The fusion process has been well studied, yet the role of chemokine coreceptor signaling in viral infection has remained elusive through the past decade. With the recent demonstration of the signaling requirement for HIV latent infection of resting CD4 T cells, the issue of coreceptor signaling needs to be thoroughly revisited. It is likely that virus-mediated signaling events may facilitate infection in various immunologic settings in vivo where cellular conditions need to be primed; in other words, HIV may exploit the chemokine signaling network shared among immune cells to gain access to downstream cellular components, which can then serve as effective tools to break cellular barriers. This virus-hijacked aberrant signaling process may in turn facilitate pathogenesis. In this review, we summarize past and present studies on HIV coreceptor signaling. We also discuss possible roles of coreceptor signaling in facilitating viral infection and pathogenesis.

Introduction

Infection by the human immunodeficiency virus (HIV) causes severe depletion of the CD4 T cell population, which eventually leads to acquired immunodeficiency syndrome (AIDS). The selective nature of the infection immediately prompted speculation and the subsequent identification of the CD4 molecule as the main surface receptor for HIV entry [1],[2]. Nevertheless, it soon became apparent that CD4 alone did not seem to be sufficient to permit entry [3],[4]. A hunt for the elusive coreceptors ensued. In 1996, Berger's group first identified a G-protein-coupled receptor designated “fusin” as the elusive cofactor for HIV-1 entry [5]. “Fusin” was later renamed CXCR4 after the realization that its natural ligand is the CXC chemokine stromal cell derived factor 1 (SDF1) [6]. Before the identification of “fusin”, chemokines such as the CC chemokines RANTES, MIP-1α, and MIP-1β had been reported to suppress infection by the macrophage tropic (M-tropic) HIV-1 [7]. Thus, with the identification of CXCR4, it took little time to confirm that CCR5, the receptor for the aforementioned CC chemokines [8], was indeed also a coreceptor for the entry of M-tropic HIV-1 [9],[10],[11],[12],[13]. Following the discoveries of CXCR4 and CCR5, several other G-protein-coupled receptors have also been identified [14]. Nevertheless, the in vivo importance of these other coreceptors in viral infection and pathogenesis is less studied than that of CXCR4 and CCR5. This review focuses mainly on CXCR4 and CCR5 signaling.

The Signaling Diversity of the Chemokine Coreceptors

The discovery of the HIV chemokine coreceptors opened up a new avenue for AIDS research. The fact that both CXCR4 and CCR5 are chemokine receptors raised interesting questions regarding the role of chemokine receptor signaling in viral infection and pathogenesis [5]. Early studies by Fauci's group demonstrated that the envelope from M-tropic but not T-tropic viruses can trigger calcium flux that was inhibited by pertussis toxin (PTX) or MIP-1β, which is suggestive of viral envelope-induced signaling transduction through CCR5 [15]. Subsequently, Davis et al. demonstrated that similar to SDF-1 and RANTES, both the T-tropic and M-tropic envelopes can induce rapid tyrosine phosphorylation of the protein tyrosine kinase Pyk2 through binding to CXCR4 or CCR5 [16]. Pyk2 phosphorylation is frequently associated with G protein signaling and calcium flux. These results provided early evidence that binding of the viral envelope to its chemokine corecptors, both CXCR4 and CCR5, not only mediates entry but also activates multiple intracellular signaling cascades, a process mimicking chemokine signaling through binding to their cognate receptors.

Chemokine receptor signaling is known to be diverse and is coupled to distinct signaling pathways that mediate cell migration, transcriptional activation, and cell growth and differentiation (Figure 1). For example, SDF-1 binding to the G-protein-coupled receptor CXCR4 activates heterotrimeric G-proteins (Gα and Gβγ). There are numerous classes of Gα (Gαs, Gαi, Gαq, Gα12/13), and CXCR4 seems to be specifically coupled to Gαi and Gαq. Gαq proteins activate phosphatidylinositol-specific phospholipases such as phospholipase C-γ (PLC-γ), which hydrolyzes phosphatidylinositol-4,5-biphosphate (PIP2) to generate inositol triphosphate (IP3) and diacylglyerol (DAG). These events lead to calcium flux and the activation of several PKC isoforms that have been shown to be important for SDF-1-induced chemotaxis [17],[18] (Figure 1A). Another pathway activated from SDF-1 binding to CXCR4 is through Gαi protein activation, which inhibits adenylyl cyclases that in turn lead to a reduction in cAMP levels as well as the activation of phospholipases and phosphodiesterases. These events result in the activation of the lipid kinase PI3K via Gαi-coupled Src-family kinases as well as PI3Kγ through direct binding of Gβγ to the regulatory subunits of PI3Kγ [19],[20],[21]. Protein kinase B (PKB/Akt) and mitogen/extracellular signal-regulated kinase (MEK-1) and extracellular signal-regulated kinase (ERK1/2) are downstream of PI3K and function in cell survival and proliferation [21],[22],[23]. PI3K also stimulates the tyrosine phosphorylation of focal adhesion complex components such as proline-rich tyrosine kinase (Pyk2) [16], paxillin, and Crk [21],[24],[25], all of which are important for cell migration and cell adhesion. PI3K is also upstream of the critical nuclear transcription activator NF-κB, which regulates gene expression in response to inflammation and activates HIV proviral gene expression [21],[26] (Figure 1B).

thumbnail
Figure 1. Chemokine receptor signaling pathways.

SDF-1 binding to CXCR4 or RANTES/MIP-1α/MIP-1β binding to CCR5 activates G proteins (Gα particularly Gαi, Gαq, and Gβγ) and multiple downstream pathways. (A) Gαq activates phospholipases such as phospholipase C-γ (PLC-γ), which hydrolyzes phosphatidylinositol-4,5-biphosphate (PIP2) to generate inositol triphosphate (IP3) and diacylglyerol (DAG), triggering calcium influx and the activation of kinases such as protein kinase C (PKC). (B) Gαi activates phospholipases, phosphodiesterases, and the lipid kinase PI3K via Src-family kinases. Gβγ also activates PI3Kγ. PI3K activation stimulates downstream targets such as protein kinase B (PKB/Akt), NF-κB, mitogen/extracellular signal-regulated kinase (MEK-1), and extracellular signal-regulated kinase (ERK1/2). PI3K also triggers the tyrosine phosphorylation of focal adhesion complex components such as proline-rich tyrosine kinase (Pyk2), paxillin, Crk, and p130Cas. (C) GTP-bound Gβγ stimulates guanine nucleotide exchange factors (GEFs) such as TIAM1 and PREX1 specific for the Rho family GTPases (Rac/CDC42/RhoA). These GTPases activate pathways regulating cytoskeleton: Rac activates p21-activated kinase (PAK), which then activates LIM kinase (LIMK), leading to cofilin phosporylation and actin polymerization. CDC42 promotes actin assembly through the Wiskott-Aldrich Syndrome family protein (WASP) and actin-nucleating protein Arp2/3. RhoA activates Rho kinase (ROCK) , leading to myosin light-chain (MLC) phosporylation and microtubule rearrangement. (D) SDF-1 may also trigger Gαi-independent activation of the JAK-STAT pathways.

https://doi.org/10.1371/journal.ppat.1000520.g001

SDF-1 binding to CXCR4 also results in GTP-bound Gβγ, which can stimulate guanine nucleotide exchange factors (GEFs) such as TIAM1 and PREX1 specific for Rho family GTPases such as Rac, Rho [27],[28],[29],[30], Cdc42 [31], and Ral [32]. These GTPases activate well-characterized downstream effector pathways, all of which modulate the cytoskeleton: Rac activates its downstream effector p21-activated kinase (PAK), which then activates LIM kinase (LIMK) [33], leading to the alteration of cofilin activity, which regulates actin turnover. Cdc42 activation promotes actin assembly through the Wiskott-Aldrich syndrome family protein's (WASP) activation of the actin nucleating protein Arp2/3 [34]. RhoA activation can lead to microtubule rearrangement through the Rho kinase (ROCK) activation of myosin light-chain phosphatase (MLCP) (Figure 1C). Finally, SDF-1 has been shown to activate perhaps some JAKs/STATs [35],[36],[37],[38], linking SDF-1-induced signaling to cytokine and growth factor–driven pathways regulating cell proliferation and differentiation. SDF-1 also activates Cbl/Cbl-b, which functions as an E3 ligase regulating cell signaling through ubiquitination [39]. In summary, most of these signaling molecules are components of the signaling transduction pathways mediating chemotactic responses for cytoskeleton rearrangement, cell polarization, and migration [19],[20],[30], as well as transcriptional activation, cell survival, and proliferation [21],[22],[40].

The Signaling Diversity of the HIV-1 Envelope–Coreceptor Interaction

Consistent with the signaling diversity of chemokine–coreceptor interaction, binding of HIV-1 gp120 to CCR5 or CXCR4 has also been shown to trigger the activation of Pyk2 [16], PI3K, Akt [41],[42], Erk-1/2 [42], and CD4/CXCR4-dependent NFAT (nuclear factor of activated T cells) nuclear translocation [43] (Figure 2). Recently, gp120 was demonstrated to mediate chemotaxis, actin cytoskeleton rearrangement [42],[44], and the activation of an actin depolymerization factor, cofilin, to increase the cortical actin dynamics in resting CD4 T cells [44].

thumbnail
Figure 2. Components of the chemokine coreceptor signaling pathways activated by HIV-1 envelope.

HIV-1 gp120 binding to CXCR4 or CCR5 activates a number of signaling molecules common to chemokine-mediated signaling pathways, including (A) PLC-γ-dependent calcium flux and NFAT nuclear translocation; (B) PI3K-dependent activation of FAK, PyK2, AKT, and ERK1/2; (C) the downstream targets of the Rho family GTPases such as LIMK1 and cofilin for actin rearrangement.

https://doi.org/10.1371/journal.ppat.1000520.g002

Given this array of targets that can be activated by gp120 in vitro, one question is whether these signaling events are physiological at the low gp120 dosages present in vivo. The presumption based on plasma viral load is that the physiological levels of gp120 are likely to be significantly lower than most of those used in in vitro experiments [45]. However, the in vivo concentrations of gp120, particularly the local gp120 concentrations in tissues, are difficult to measure, and a large percentage of the virus in the body is present in lymphoid tissues. Therefore, the compartmentalized viral concentrations could be very high regardless of the plasma viral loads [46]. Thus, the judgment of whether or not an experimental gp120 dosage is physiological is largely arbitrary at best. On the other hand, a recent experiment suggested that a few HIV particles might be sufficient to trigger signaling through CXCR4 [47].

In our recent studies of gp120 signaling in peripheral resting CD4 T cells, we did notice dosage- and conformation-dependent differences in gp120 signaling [44],[48]. At high gp120 dosages, gp120 acts more like SDF-1, triggering rapid cofilin phosphorylation and actin polymerization that was followed by cofilin dephosphorylation and actin depolymerization [42],[44]. At lower dosages, gp120 was incapable of triggering such rapid changes, but was able to mediate gradual cofilin dephosphorylation and actin depolymerization [44],[48]. These differences may play different roles in various settings during the course of HIV infection.

Roles of HIV-1 Envelope–Coreceptor Signaling in HIV Infection

The direct involvement of chemokine coreceptor signaling in HIV infection has been speculated right from the identification of the chemokine coreceptors [5]. Within a year of the initial discovery, multiple groups started to test the requirement of coreceptor signaling in HIV entry and subsequent replication steps. Cocchi and colleagues were the first to demonstrate that inhibition of coreceptor signaling through PTX neither inhibited HIV-1 replication nor the ability of chemokines such as RANTES to block HIV-1 entry into PM1 cells [49]. Similarly, Farzan et al. further created three CCR5 mutants that abolished its signaling ability to mobilize calcium but detected minimal effects on viral entry and replication in Hela-CD4 cells [50]. These results were corroborated by several other groups using cell lines transfected with either CCR5- or CXCR4-signaling-defective mutants [51],[52],[53],[54],[55]; the results largely suggested that the signaling function of the chemokine coreceptors is an independent function not required for viral entry or replication. Nevertheless, in the following decade, a significant number of studies have been published describing modulation of cellular functions by HIV gp120 signaling, ranging from causing neurotoxicity to promoting apoptosis. Therefore, the relevance of chemokine coreceptor signaling to viral infection itself remained an open question. With this significant issue unresolved, there were also sporadic findings implying that chemokine coreceptor signaling might be important for viral replication. For example, prestimulation of macrophages or CD4 T cells with CC-chemokines was found to enhance HIV replication [56],[57]. Similar stimulation with viral gp120 can even induce viral replication in cultures of resting CD4 T cells of infected patients [58]. These positive effects of coreceptor signaling were also reflected in facilitating HIV infection of non-natural targets; the replication of some HIV-1 isolates in macaque cells was blocked at a step after entry and reverse transcription but prior to the nuclear import of the preintegration complex, and this block could be relieved by the expression of human coreceptor CCR5 or CXCR4 in macaque cells [59]. These results suggested a possible involvement of coreceptor signaling in a post entry step. In agreement with this hypothesis, Mori and colleagues demonstrated that three to nine amino acid changes in the envelope of SIVmac293 conferred upon the virus the ability to replicate 100–1,000 times more efficiently in macrophages. Importantly, the amino acid changes in the envelope did not enhance virus entry, but rather affected some post entry steps [60]. Furthermore, Arthos and colleagues [61] demonstrated that there was a direct correlation between the capacity of viral envelopes to initiate signaling and the ability of the same viruses to infect cells. They also demonstrated that a signaling-deficient R5 HIV-1, 92MW959, entered macrophages but failed to replicate, and the block was at a post entry step. In parallel, the density of CCR5 on the cell surface also appeared to correlate with the capacity of the coreceptor to transduce signals, and this capacity directly impacted viral post entry processes such as reverse transcription and integration [62]. Recently, Grainger and Lever [63] used a family of new chemokine inhibitors to block chemokine receptor signaling without affecting receptor binding. When tested in HIV-infected cells, these inhibitors inhibit HIV-1 infection without blocking gp120/CCR5 interaction [63], and presumably, these inhibitors also inhibited viral replication at steps post binding and entry.

Despite growing evidence supporting a possible role for coreceptor signaling, contradictory findings remained. Amara et al. compared R5 HIV infection of CCR5 wild-type and signaling-deficient T cells derived from CCR5Δ32 individuals [55]. In this study, signaling-deficient CCR5 supported R5 HIV replication to an extent similar to the wild-type CCR5, suggesting again that G protein signaling through CCR5 plays no role in R5 HIV-1 replication. An important aspect of this study was the use of PHA to activate primary T cells for transduction. In contrast, Lin et al. used unstimulated primary T cells in a similar experiment and showed that the signaling-deficient CCR5 T cells were impaired in their support of HIV-1 infection [64]. Importantly, in order to observe the inhibition, cells in this experiment had to be rested for ten days after transduction before HIV-1 infection. These data are consistent with an earlier observation that the B-oligomer of PTX, although not directly inhibiting coreceptor signaling thourgh Gαi, diminishes the R5 virus-mediated receptor capping and reverse transcription [65]. These inhibitory effects can only be observed in primary monocyte-depleted PBMCs but not in transformed PM1 cells [65]. It is likely that much of the seemingly contradictory observations in different studies may arise from the use of cells in different activation states. When studies used activated T cells or transformed cell lines, they largely observed that HIV-1 envelope–coreceptor signaling was dispensable. Since a majority of T cells and macrophages in vivo are in resting and noncycling states, it is critical that the study of HIV-initiated signaling is investigated using similar cells.

Recently, we examined the requirement for CXCR4 signaling in HIV-1 latent infection of human resting CD4 T cells using an in vitro system to mimic in vivo latent infection of resting T cells [44]. Using this system, we observed an absolute requirement of CXCR4 signaling for HIV-1 latent infection of resting CD4 T cells. This requirement was even maintained in resting T cells exposed to certain cytokines such as IL-2 and IL-7, suggesting that CXCR4 signaling could potentiate infectivity even in certain cytokine-enriched microenvironments, such as lymphoid tissues. We also identified the molecular mechanism of this signaling requirement, and demonstrated that the static cortical actin in noncycling, resting CD4 T cells represents a unique barrier for viral post entry migration. To overcome this restriction, HIV-1 relies on viral envelope and the Gαi-dependent signaling from the chemokine coreceptor CXCR4 to activate a cellular actin-depolymerizing factor, cofilin, to increase the cortical actin dynamics (Figure 3). This unique requirement for coreceptor signaling can only be observed in noncycling, resting CD4 T cells because in transformed or activated T cells cofilin is constitutively active to facilitate the cell cycle–driven actin remodeling that renders CXCR4 signaling unnecessary [44].

thumbnail
Figure 3. HIV-1 envelope-CXCR4 signaling triggers cofilin activation to promote cortical actin dynamics and HIV nuclear migration.

(A) HIV-1 envelope binding to CD4 and the chemokine coreceptor, CXCR4, triggers membrane fusion and signaling transduction. The initial viral contact with CD4 and then CXCR4 may trigger rapid actin polymerization to facilitate CD4/CXCR4 cocapping for fusion and entry. Following fusion, the viral preintegration complex (PIC) may be directly anchored onto F-actin to facilitate reverse transcription. Subsequent actin activity mediated by cofilin activation through CXCR4 promotes viral nuclear migration. (B) Model of HIV PIC migration along the cortical actin filaments. It is possible that cofilin activation increases actin treadmilling, which promotes the movement of the viral PIC across the cortical actin barrier, allowing PIC to gain access to the perinuclear or nuclear region. The number is arbitrarily assigned to an actin monomer to demonstrate the actin movement during treadmilling. (Figure 3 is modified from [44] with permission).

https://doi.org/10.1371/journal.ppat.1000520.g003

With the identification of the molecular mechanism of CXCR4 signaling, the remaining issues are to confirm the requirement for CCR5 signaling, and to determine the molecular basis for this requirement. There appears to be a clear distinction in the requirement for signaling between R5 and X4 viruses. For example, the B-oligomer of PTX inhibits R5 virus at entry and reverse transcription, but inhibits the X4 virus at multiple steps post reverse transcription [65],[66]. The R5 virus predominates at the early time of HIV infection, largely infecting macrophages and active memory CD4 T cells [67] in the gastrointestinal (GI) tract and lymph nodes [68]. Both types of cells support productive viral replication and do not have the restrictions usually seen in resting CD4 T cells. It is unknown whether HIV-mediated cofilin activation through CCR5 may also occur in active memory T cells and macrophages, since cells that are actively cycling and migrating normally disassemble actin cytoskeleton themselves, leaving the cells naturally susceptible to HIV-1 infection. If required, the CCR5 signaling might be involved in several early processes: initial fusion and entry, uncoating, viral DNA synthesis, or subsequent viral gene transcription. Chemokine receptor signaling also triggers rapid actin polymerization [42],[44], which may be important for the cocapping of CD4 with CCR5 or CXCR4 [69] and for the efficient synthesis of viral DNA [44],[70]. This early actin polymerization appears to be involved in assembling high concentrations of CD4 and CCR5 or CXCR4 at the plasma membrane, which facilitate gp120 binding and viral entry (for a review, please see [71]). Nevertheless, it is unlikely that this immediate actin activity is triggered through CCR5 or CXCR4 exclusively. Rather, it may be mediated through the initial gp120 contact with CD4 and then facilitated through further gp120 contact with CCR5 or CXCR4 [72] (Figure 3). Also, sufficient actin contact with the core following fusion is probably critical for rapid uncoating, or a subsequent particle conformational change that is necessary for optimal reverse transcription. Indeed, excessive actin depolymerization inhibits viral replication, and artificially increasing the cortical actin density through small interfering RNA (siRNA)-mediated cofilin knockdown enhances viral DNA synthesis [44]. In transformed cell lines, the requirement for CCR5 signaling in viral replication was not observed, likely because the cortical actin itself is dynamic and minimally affected by inhibiting CCR5 signaling [44]. In addition, the presence of CD4 signaling may also compensate some of the early requirements for actin activity. Interestingly, when the density of CCR5 was artificially increased on transformed HOS-CD4 cells, a direct correlation between CCR5 density and viral DNA synthesis was observed, and this correlation was dependent on PTX-sensitive Gαi signaling [62].

CCR5 signaling has also been known to trigger distinctive signaling cascades that activate kinases and transcription factors associated with cell activation [43],[73],[74]. For example, R5 envelopes can induce the expression of genes belonging to MAPK signal transduction pathways and genes regulating the cell cycle [74]. R5 envelopes can also activate NFATs and induce their translocation into the nucleus [43]. Because the HIV long terminal repeat (LTR) encodes NFAT recognition sites, NFAT activation likely enhances viral transcription directly from the LTR promoter, especially at early time points when cellular conditions are limiting. For example, the microenvironment in the GI tract where the R5 viruses infect active memory T cells is filled with inhibitory cytokines such as TGF-β [75], which can reorganize cytoskeletal structure [76] and inhibit the activation of major transcription factors such as NFκB [77]. The contribution of CCR5 signaling to viral transcription could probably be observed in suboptimally activated cells, but may not be seen in highly active transformed cell lines.

Possible Roles of HIV-1 Envelope–Receptor Signaling in Viral Pathogenesis

It has long been suggested that HIV-1 envelope plays a central role in HIV-mediated CD4 T cell depletion and pathogenesis. Prior to the identification of the chemokine coreceptors, gp120 was proposed to directly trigger activation-dependent T cell apoptosis through its binding to surface CD4 [78]. Later, when the chemokine coreceptors were discovered, gp120 binding to CCR5 or CXCR4 was also proposed to trigger the apoptosis of CD4 cells, CD8 T cells [79],[80], and neurons [81]. There was considerable complexity and discrepancy regarding possible mechanisms of gp120-mediated apoptosis in T cells, in terms of possible involvements of Gαi-dependent pathways and caspases [80],[82], as well as the involvement of autophagy [83]. Envelope-primed apoptosis has been implicated as directly contributing to the depletion of either infected or uninfected CD4 T cells during disease progression [78],[84],[85]. Nevertheless, latently infected resting T cells were frequently detected persisting in patients [86], and the half life of these cells can be as long as 3 to 4 years [87]. These in vivo findings suggested that although gp120 can trigger T cell apoptosis in certain circumstances, there must be mechanisms that may prevent this process from occurring during the initial virus–cell engagement [88].

HIV envelope–chemokine receptor signaling has been shown to mediate chemotaxis in both CD4 and CD8 T cells [15],[42],[44],[89],[90], a property directly related to the role of chemokine receptor signaling. The viral envelope has also been suggested to do the opposite by repelling T cells [89], a process resembling chemofugetaxis, in which high concentrations of chemokines such as SDF-1 drive away T cells [91]. It has been proposed that this capacity of gp120 repels antigen-specific cytotoxic T lymphocytes from viral infection sites to evade immune effector mechanisms [89]. Similar functions of gp120 in modulating the immune system have also been suggested. For example, persistent gp120 stimulation induces CXCR4-dependent T cell anergy [92], as well as CCR5- and CXCR4-dependent macrophage activation and the secretion of proinflammatory cytokines [93]. Many of these gp120-induced phenotypes resemble the pathogenic features of HIV-induced T cell dysfunction and chronic immune activation [92],[94],[95],[96],[97]. However, the molecular targets and signaling mechanisms directly responsible for these phenotypes remain largely uncharacterized. Recently, we identified cofilin as one of the primary downstream targets of gp120–CXCR4 signaling in human resting CD4 T cells [44]. We also found that in the resting CD4 T cells of infected patients, cofilin activity is aberrantly upregulated [48]. HIV-mediated cofilin activation likely results from a bystander effect since a majority of resting CD4 T cells in the peripheral blood of infected patients contain no virus (0.2–16.4 cells latently infected with HIV per 106 resting CD4 T cells [98]). Although not directly infected, these residual CD4 T cells in patients are also known to carry numerous functional abnormalities such as loss of T helper function [94], anergy [92],[95], increased T cell proliferation [99], and abnormal T cell homing and migration [89],[100].

In the human immune system, cofilin is involved in two hallmark activities of T cells, namely chemotaxis and T cell activation [101]. In chemotaxis, cofilin is the driving force for promoting the cortical actin dynamics central to cell migration [33]. In antigen-specific T cell activation, cofilin is activated through CD2/CD28 co-stimulation, and plays a critical role in actin reorganization and the formation of the immunological synapse that is required for sustaining T cell activation [102]. It is possible that cofilin dysregulation could result in CD4 T cell abnormities that may contribute to T cell depletion and immune deficiencies. Severe effects of altered actin dynamics on the human immune system have been well documented. For instance, a genetic defect of WASP that affects actin dynamics causes immunodeficiency [103]. HIV-mediated aberrant activation of cofilin strikingly resembles the activation of cofilin detected in migratory T lymphoma cells [104]. This resemblance may indicate a similar abnormal migratory behavior that could result in the eventual destruction of peripheral CD4 T cells in lymph nodes or tissues of HIV patients.

The utilization of different chemokine coreceptors, either CCR5 or CXCR4, for entry largely differentiates HIV into two distinctive phenotypes with either M- or T-tropism. The high pathogenic potential of the late emerging T-tropic viruses in causing rapid CD4 depletion is clearly a demonstration of the pathogenic significance of the CXCR4-engaging viruses [105],[106],[107]. Nevertheless, a remaining question for the importance of CXCR4 signaling in HIV pathogenesis is the lack of X4 viruses in some patients. There are only approximately 50% patients who experience the conversion from R5 to X4 viruses at later stages of disease. It is possible that in these patients, a complete viral switch from M- to T-tropism may never occur, but constant viral mutation may generate intermediate viruses that can engage and signal through CXCR4. It has been known that one or two amino acid changes in the V3 loop of the viral envelope may confer CXCR4 binding [108], although successful fusion and entry often require more than two mutations or mutations even outside of the V3 loop [109]. It is also possible that these hypothetical X4-signaling viruses may trigger signaling transduction and cause CD4 T cell dysfunction without actually replicating in them (for further reading on this hypothesis, please see [110]). In comparison with the highly pathogenic T-tropic viruses, these M-tropic, X4-signaling viruses may cause slower CD4 depletion and disease progression.

Conclusions and Perspectives

Understanding HIV envelope–coreceptor interaction holds the key to unlocking the mystery of HIV-mediated CD4 depletion and pathogenesis. The signaling capacity of the viral envelope, along with its pathogenic potential, has been a subject of great interest and intense investigation. Yet, the critical role of the coreceptor signaling in HIV disease had not been clearly defined. This lack of appreciation has been largely attributed to the complexity of the chemokine signaling network itself, as well as the complication of many in vitro systems used, in which cellular states are derailed from the genuine signaling circuitries of natural viral targets in the body. Given these complexities, a critical function of the coreceptor signaling in promoting viral infection has only recently been demonstrated in the latent infection of human resting CD4 T cells. The deciphering of the molecular mechanisms involved starts to offer exciting fresh perspectives and opens a new avenue for examining HIV pathogenesis. However, a great deal is yet to be tested regarding the in vivo importance of coreceptor signaling in mediating CD4 T cell dysfunction and pathogenesis.

Gene Accession Numbers

Human CXCR4 – Entrez Gene ID # 7852; Human CCR5 – Entrez Gene ID # 1234; Human non-muscle Cofilin 1 – Entrez Gene ID # 1072.

References

  1. 1. Klatzmann D, Champagne E, Chamaret S, Gruest J, Guetard D, et al. (1984) T-lymphocyte T4 molecule behaves as the receptor for human retrovirus LAV. Nature 312: 767–768.
  2. 2. Dalgleish AG, Beverley PC, Clapham PR, Crawford DH, Greaves MF, et al. (1984) The CD4 (T4) antigen is an essential component of the receptor for the AIDS retrovirus. Nature 312: 763–767.
  3. 3. Ashorn PA, Berger EA, Moss B (1990) Human immunodeficiency virus envelope glycoprotein/CD4-mediated fusion of nonprimate cells with human cells. J Virol 64: 2149–2156.
  4. 4. Weiner DB, Huebner K, Williams WV, Greene MI (1991) Human genes other than CD4 facilitate HIV-1 infection of murine cells. Pathobiology 59: 361–371.
  5. 5. Feng Y, Broder CC, Kennedy PE, Berger EA (1996) HIV-1 entry cofactor: functional cDNA cloning of a seven-transmembrane, G protein-coupled receptor. Science 272: 872–877.
  6. 6. Oberlin E, Amara A, Bachelerie F, Bessia C, Virelizier JL, et al. (1996) The CXC chemokine SDF-1 is the ligand for LESTR/fusin and prevents infection by T-cell-line-adapted HIV-1. Nature 382: 833–835.
  7. 7. Cocchi F, DeVico AL, Garzino-Demo A, Arya SK, Gallo RC, et al. (1995) Identification of RANTES, MIP-1 alpha, and MIP-1 beta as the major HIV-suppressive factors produced by CD8+ T cells. Science 270: 1811–1815.
  8. 8. Combadiere C, Ahuja SK, Tiffany HL, Murphy PM (1996) Cloning and functional expression of CC CKR5, a human monocyte CC chemokine receptor selective for MIP-1(alpha), MIP-1(beta), and RANTES. J Leukoc Biol 60: 147–152.
  9. 9. Alkhatib G, Combadiere C, Broder CC, Feng Y, Kennedy PE, et al. (1996) CC CKR5: a RANTES, MIP-1alpha, MIP-1beta receptor as a fusion cofactor for macrophage-tropic HIV-1. Science 272: 1955–1958.
  10. 10. Choe H, Farzan M, Sun Y, Sullivan N, Rollins B, et al. (1996) The beta-chemokine receptors CCR3 and CCR5 facilitate infection by primary HIV-1 isolates. Cell 85: 1135–1148.
  11. 11. Deng H, Liu R, Ellmeier W, Choe S, Unutmaz D, et al. (1996) Identification of a major co-receptor for primary isolates of HIV-1. Nature 381: 661–666.
  12. 12. Doranz BJ, Rucker J, Yi Y, Smyth RJ, Samson M, et al. (1996) A dual-tropic primary HIV-1 isolate that uses fusin and the beta-chemokine receptors CKR-5, CKR-3, and CKR-2b as fusion cofactors. Cell 85: 1149–1158.
  13. 13. Dragic T, Litwin V, Allaway GP, Martin SR, Huang Y, et al. (1996) HIV-1 entry into CD4+ cells is mediated by the chemokine receptor CC-CKR-5. Nature 381: 667–673.
  14. 14. Berger EA, Murphy PM, Farber JM (1999) Chemokine receptors as HIV-1 coreceptors: roles in viral entry, tropism, and disease. Annu Rev Immunol 17: 657–700.
  15. 15. Weissman D, Rabin RL, Arthos J, Rubbert A, Dybul M, et al. (1997) Macrophage-tropic HIV and SIV envelope proteins induce a signal through the CCR5 chemokine receptor. Nature 389: 981–985.
  16. 16. Davis CB, Dikic I, Unutmaz D, Hill CM, Arthos J, et al. (1997) Signal transduction due to HIV-1 envelope interactions with chemokine receptors CXCR4 or CCR5. Journal of Experimental Medicine 186: 1793–1798.
  17. 17. Shahabi NA, McAllen K, Sharp BM (2008) Stromal cell-derived factor 1-alpha (SDF)-induced human T cell chemotaxis becomes phosphoinositide 3-kinase (PI3K)-independent: role of PKC-theta. J Leukoc Biol 83: 663–671.
  18. 18. Petit I, Goichberg P, Spiegel A, Peled A, Brodie C, et al. (2005) Atypical PKC-zeta regulates SDF-1-mediated migration and development of human CD34+ progenitor cells. J Clin Invest 115: 168–176.
  19. 19. Sotsios Y, Whittaker GC, Westwick J, Ward SG (1999) The CXC chemokine stromal cell-derived factor activates a Gi-coupled phosphoinositide 3-kinase in T lymphocytes. Journal of Immunology 163: 5954–5963.
  20. 20. Vicente-Manzanares M, Rey M, Jones DR, Sancho D, Mellado M, et al. (1999) Involvement of phosphatidylinositol 3-kinase in stromal cell-derived factor-1 alpha-induced lymphocyte polarization and chemotaxis. Journal of Immunology 163: 4001–4012.
  21. 21. Ganju RK, Brubaker SA, Meyer J, Dutt P, Yang Y, et al. (1998) The alpha-chemokine, stromal cell-derived factor-1alpha, binds to the transmembrane G-protein-coupled CXCR-4 receptor and activates multiple signal transduction pathways. J Biol Chem 273: 23169–23175.
  22. 22. Barbero S, Bonavia R, Bajetto A, Porcile C, Pirani P, et al. (2003) Stromal cell-derived factor 1alpha stimulates human glioblastoma cell growth through the activation of both extracellular signal-regulated kinases 1/2 and Akt. Cancer Res 63: 1969–1974.
  23. 23. Chinni SR, Sivalogan S, Dong Z, Filho JC, Deng X, et al. (2006) CXCL12/CXCR4 signaling activates Akt-1 and MMP-9 expression in prostate cancer cells: the role of bone microenvironment-associated CXCL12. Prostate 66: 32–48.
  24. 24. Wang JF, Park IW, Groopman JE (2000) Stromal cell-derived factor-1alpha stimulates tyrosine phosphorylation of multiple focal adhesion proteins and induces migration of hematopoietic progenitor cells: roles of phosphoinositide-3 kinase and protein kinase C. Blood 95: 2505–2513.
  25. 25. Cicala C, Arthos J, Ruiz M, Vaccarezza M, Rubbert A, et al. (1999) Induction of phosphorylation and intracellular association of CC chemokine receptor 5 and focal adhesion kinase in primary human CD4+ T cells by macrophage-tropic HIV envelope. J Immunol 163: 420–426.
  26. 26. Nabel G, Baltimore D (1987) An inducible transcription factor activates expression of human immunodeficiency virus in T cells. Nature 326: 711–713.
  27. 27. Bartolome RA, Galvez BG, Longo N, Baleux F, Van Muijen GN, et al. (2004) Stromal cell-derived factor-1alpha promotes melanoma cell invasion across basement membranes involving stimulation of membrane-type 1 matrix metalloproteinase and Rho GTPase activities. Cancer Research 64: 2534–2543.
  28. 28. Tan W, Martin D, Gutkind JS (2006) The Galpha13-Rho signaling axis is required for SDF-1-induced migration through CXCR4. J Biol Chem 281: 39542–39549.
  29. 29. Henschler R, Piiper A, Bistrian R, Mobest D (2003) SDF-1alpha-induced intracellular calcium transient involves Rho GTPase signalling and is required for migration of hematopoietic progenitor cells. Biochem Biophys Res Commun 311: 1067–1071.
  30. 30. Vicente-Manzanares M, Cabrero JR, Rey M, Perez-Martinez M, Ursa A, et al. (2002) A role for the Rho-p160 Rho coiled-coil kinase axis in the chemokine stromal cell-derived factor-1alpha-induced lymphocyte actomyosin and microtubular organization and chemotaxis. J Immunol 168: 400–410.
  31. 31. Haddad E, Zugaza JL, Louache F, Debili N, Crouin C, et al. (2001) The interaction between Cdc42 and WASP is required for SDF-1-induced T-lymphocyte chemotaxis. Blood 97: 33–38.
  32. 32. de Gorter DJ, Reijmers RM, Beuling EA, Naber HP, Kuil A, et al. (2008) The small GTPase Ral mediates SDF-1-induced migration of B cells and multiple myeloma cells. Blood 111: 3364–3372.
  33. 33. Nishita M, Aizawa H, Mizuno K (2002) Stromal cell-derived factor 1alpha activates LIM kinase 1 and induces cofilin phosphorylation for T-cell chemotaxis. Molecular & Cellular Biology 22: 774–783.
  34. 34. Okabe S, Fukuda S, Broxmeyer HE (2002) Activation of Wiskott-Aldrich syndrome protein and its association with other proteins by stromal cell-derived factor-1alpha is associated with cell migration in a T-lymphocyte line. Exp Hematol 30: 761–766.
  35. 35. Vila-Coro AJ, Rodriguez-Frade JM, Martin De Ana A, Moreno-Ortiz MC, Martinez AC, et al. (1999) The chemokine SDF-1alpha triggers CXCR4 receptor dimerization and activates the JAK/STAT pathway. Faseb J 13: 1699–1710.
  36. 36. Ahr B, Denizot M, Robert-Hebmann V, Brelot A, Biard-Piechaczyk M (2005) Identification of the cytoplasmic domains of CXCR4 involved in Jak2 and STAT3 phosphorylation. J Biol Chem 280: 6692–6700.
  37. 37. Moriguchi M, Hissong BD, Gadina M, Yamaoka K, Tiffany HL, et al. (2005) CXCL12 signaling is independent of Jak2 and Jak3. J Biol Chem 280: 17408–17414.
  38. 38. Mowafi F, Cagigi A, Matskova L, Bjork O, Chiodi F, et al. (2008) Chemokine CXCL12 enhances proliferation in pre-B-ALL via STAT5 activation. Pediatr Blood Cancer 50: 812–817.
  39. 39. Okabe S, Tauchi T, Ohyashiki K, Broxmeyer HE (2006) Stromal-cell-derived factor-1/CXCL12-induced chemotaxis of a T cell line involves intracellular signaling through Cbl and Cbl-b and their regulation by Src kinases and CD45. Blood Cells Mol Dis 36: 308–314.
  40. 40. Porcile C, Bajetto A, Barbieri F, Barbero S, Bonavia R, et al. (2005) Stromal cell-derived factor-1alpha (SDF-1alpha/CXCL12) stimulates ovarian cancer cell growth through the EGF receptor transactivation. Exp Cell Res 308: 241–253.
  41. 41. Francois F, Klotman ME (2003) Phosphatidylinositol 3-kinase regulates human immunodeficiency virus type 1 replication following viral entry in primary CD4+ T lymphocytes and macrophages. J Virol 77: 2539–2549.
  42. 42. Balabanian K, Harriague J, Decrion C, Lagane B, Shorte S, et al. (2004) CXCR4-tropic HIV-1 envelope glycoprotein functions as a viral chemokine in unstimulated primary CD4+ T lymphocytes. J Immunol 173: 7150–7160.
  43. 43. Cicala C, Arthos J, Censoplano N, Cruz C, Chung E, et al. (2006) HIV-1 gp120 induces NFAT nuclear translocation in resting CD4+ T-cells. Virology 345: 105–114.
  44. 44. Yoder A, Yu D, Dong L, Iyer SR, Xu X, et al. (2008) HIV envelope-CXCR4 signaling activates cofilin to overcome cortical actin restriction in resting CD4 T cells. Cell 134: 782–792.
  45. 45. Klasse PJ, Moore JP (2004) Is there enough gp120 in the body fluids of HIV-1-infected individuals to have biologically significant effects? Virology 323: 1–8.
  46. 46. Hockett RD, Kilby JM, Derdeyn CA, Saag MS, Sillers M, et al. (1999) Constant mean viral copy number per infected cell in tissues regardless of high, low, or undetectable plasma HIV RNA. J Exp Med 189: 1545–1554.
  47. 47. Melar M, Ott DE, Hope TJ (2007) Physiological levels of virion-associated human immunodeficiency virus type 1 envelope induce coreceptor-dependent calcium flux. J Virol 81: 1773–1785.
  48. 48. Wu Y, Yoder A, Yu D, Wang W, Liu J, et al. (2008) Cofilin activation in peripheral CD4 T cells of HIV-1 infected patients: a pilot study. Retrovirology 5: 95.
  49. 49. Cocchi F, DeVico AL, Garzino-Demo A, Cara A, Gallo RC, et al. (1996) The V3 domain of the HIV-1 gp120 envelope glycoprotein is critical for chemokine-mediated blockade of infection. Nat Med 2: 1244–1247.
  50. 50. Farzan M, Choe H, Martin KA, Sun Y, Sidelko M, et al. (1997) HIV-1 entry and macrophage inflammatory protein-1beta-mediated signaling are independent functions of the chemokine receptor CCR5. Journal of Biological Chemistry 272: 6854–6857.
  51. 51. Alkhatib G, Locati M, Kennedy PE, Murphy PM, Berger EA (1997) HIV-1 coreceptor activity of CCR5 and its inhibition by chemokines: independence from G protein signaling and importance of coreceptor downmodulation. Virology 234: 340–348.
  52. 52. Gosling J, Monteclaro FS, Atchison RE, Arai H, Tsou CL, et al. (1997) Molecular uncoupling of C-C chemokine receptor 5-induced chemotaxis and signal transduction from HIV-1 coreceptor activity. Proc Natl Acad Sci U S A 94: 5061–5066.
  53. 53. Aramori I, Ferguson SS, Bieniasz PD, Zhang J, Cullen B, et al. (1997) Molecular mechanism of desensitization of the chemokine receptor CCR-5: receptor signaling and internalization are dissociable from its role as an HIV-1 co-receptor. Embo J 16: 4606–4616.
  54. 54. Amara A, Gall SL, Schwartz O, Salamero J, Montes M, et al. (1997) HIV coreceptor downregulation as antiviral principle: SDF-1alpha-dependent internalization of the chemokine receptor CXCR4 contributes to inhibition of HIV replication. J Exp Med 186: 139–146.
  55. 55. Amara A, Vidy A, Boulla G, Mollier K, Garcia-Perez J, et al. (2003) G protein-dependent CCR5 signaling is not required for efficient infection of primary T lymphocytes and macrophages by R5 human immunodeficiency virus type 1 isolates. Journal of Virology 77: 2550–2558.
  56. 56. Kelly MD, Naif HM, Adams SL, Cunningham AL, Lloyd AR (1998) Dichotomous effects of beta-chemokines on HIV replication in monocytes and monocyte-derived macrophages. J Immunol 160: 3091–3095.
  57. 57. Kinter A, Catanzaro A, Monaco J, Ruiz M, Justement J, et al. (1998) CC-chemokines enhance the replication of T-tropic strains of HIV-1 in CD4(+) T cells: role of signal transduction. Proc Natl Acad Sci U S A 95: 11880–11885.
  58. 58. Kinter AL, Umscheid CA, Arthos J, Cicala C, Lin Y, et al. (2003) HIV envelope induces virus expression from resting CD4+ T cells isolated from HIV-infected individuals in the absence of markers of cellular activation or apoptosis. Journal of Immunology 170: 2449–2455.
  59. 59. Chackerian B, Long EM, Luciw PA, Overbaugh J (1997) Human immunodeficiency virus type 1 coreceptors participate in postentry stages in the virus replication cycle and function in simian immunodeficiency virus infection. J Virol 71: 3932–3939.
  60. 60. Mori K, Ringler DJ, Desrosiers RC (1993) Restricted replication of simian immunodeficiency virus strain 239 in macrophages is determined by env but is not due to restricted entry. J Virol 67: 2807–2814.
  61. 61. Arthos J, Rubbert A, Rabin RL, Cicala C, Machado E, et al. (2000) CCR5 signal transduction in macrophages by human immunodeficiency virus and simian immunodeficiency virus envelopes. J Virol 74: 6418–6424.
  62. 62. Lin YL, Mettling C, Portales P, Reynes J, Clot J, et al. (2002) Cell surface CCR5 density determines the postentry efficiency of R5 HIV-1 infection. Proc Natl Acad Sci U S A 99: 15590–15595.
  63. 63. Grainger DJ, Lever AM (2005) Blockade of chemokine-induced signalling inhibits CCR5-dependent HIV infection in vitro without blocking gp120/CCR5 interaction. Retrovirology 2: 23.
  64. 64. Lin YL, Mettling C, Portales P, Reant B, Clot J, et al. (2005) G-protein signaling triggered by R5 human immunodeficiency virus type 1 increases virus replication efficiency in primary T lymphocytes. J Virol 79: 7938–7941.
  65. 65. Alfano M, Schmidtmayerova H, Amella CA, Pushkarsky T, Bukrinsky M (1999) The B-oligomer of pertussis toxin deactivates CC chemokine receptor 5 and blocks entry of M-tropic HIV-1 strains. J Exp Med 190: 597–605.
  66. 66. Alfano M, Pushkarsky T, Poli G, Bukrinsky M (2000) The B-oligomer of pertussis toxin inhibits human immunodeficiency virus type 1 replication at multiple stages. Journal of Virology 74: 8767–8770.
  67. 67. Bleul CC, Wu L, Hoxie JA, Springer TA, Mackay CR (1997) The HIV coreceptors CXCR4 and CCR5 are differentially expressed and regulated on human T lymphocytes. Proc Natl Acad Sci U S A 94: 1925–1930.
  68. 68. Brenchley JM, Schacker TW, Ruff LE, Price DA, Taylor JH, et al. (2004) CD4+ T cell depletion during all stages of HIV disease occurs predominantly in the gastrointestinal tract. J Exp Med 200: 749–759.
  69. 69. Iyengar S, Hildreth JE, Schwartz DH (1998) Actin-dependent receptor colocalization required for human immunodeficiency virus entry into host cells. J Virol 72: 5251–5255.
  70. 70. Bukrinskaya A, Brichacek B, Mann A, Stevenson M (1998) Establishment of a functional human immunodeficiency virus type 1 (HIV-1) reverse transcription complex involves the cytoskeleton. J Exp Med 188: 2113–2125.
  71. 71. Liu Y, Belkina NV, Shaw S (2009) HIV infection of T cells: actin-in and actin-out. Sci Signal 2: pe23.
  72. 72. Jimenez-Baranda S, Gomez-Mouton C, Rojas A, Martinez-Prats L, Mira E, et al. (2007) Filamin-A regulates actin-dependent clustering of HIV receptors. Nat Cell Biol 9: 838–846.
  73. 73. Cicala C, Arthos J, Selig SM, Dennis G Jr., Hosack DA, et al. (2002) HIV envelope induces a cascade of cell signals in non-proliferating target cells that favor virus replication. Proc Natl Acad Sci U S A 99: 9380–9385.
  74. 74. Cicala C, Arthos J, Martinelli E, Censoplano N, Cruz CC, et al. (2006) R5 and X4 HIV envelopes induce distinct gene expression profiles in primary peripheral blood mononuclear cells. Proc Natl Acad Sci U S A 103: 3746–3751.
  75. 75. Amarnath S, Dong L, Li J, Wu Y, Chen W (2007) Endogenous TGF-beta activation by reactive oxygen species is key to Foxp3 induction in TCR-stimulated and HIV-1-infected human CD4+CD25- T cells. Retrovirology 4: 57.
  76. 76. Moustakas A, Stournaras C (1999) Regulation of actin organisation by TGF-beta in H-ras-transformed fibroblasts. J Cell Sci 112(Pt8): 1169–1179.
  77. 77. Monteleone G, Mann J, Monteleone I, Vavassori P, Bremner R, et al. (2004) A failure of transforming growth factor-beta1 negative regulation maintains sustained NF-kappaB activation in gut inflammation. J Biol Chem 279: 3925–3932.
  78. 78. Banda NK, Bernier J, Kurahara DK, Kurrle R, Haigwood N, et al. (1992) Crosslinking CD4 by human immunodeficiency virus gp120 primes T cells for activation-induced apoptosis. J Exp Med 176: 1099–1106.
  79. 79. Herbein G, Mahlknecht U, Batliwalla F, Gregersen P, Pappas T, et al. (1998) Apoptosis of CD8+ T cells is mediated by macrophages through interaction of HIV gp120 with chemokine receptor CXCR4. Nature 395: 189–194.
  80. 80. Berndt C, Mopps B, Angermuller S, Gierschik P, Krammer PH (1998) CXCR4 and CD4 mediate a rapid CD95-independent cell death in CD4(+) T cells. Proc Natl Acad Sci U S A 95: 12556–12561.
  81. 81. Hesselgesser J, Taub D, Baskar P, Greenberg M, Hoxie J, et al. (1998) Neuronal apoptosis induced by HIV-1 gp120 and the chemokine SDF-1 alpha is mediated by the chemokine receptor CXCR4. Current Biology 8: 595–598.
  82. 82. Blanco J, Jacotot E, Cabrera C, Cardona A, Clotet B, et al. (1999) The implication of the chemokine receptor CXCR4 in HIV-1 envelope protein-induced apoptosis is independent of the G protein-mediated signalling. Aids 13: 909–917.
  83. 83. Espert L, Denizot M, Grimaldi M, Robert-Hebmann V, Gay B, et al. (2006) Autophagy is involved in T cell death after binding of HIV-1 envelope proteins to CXCR4. J Clin Invest 116: 2161–2172.
  84. 84. Ameisen JC, Capron A (1991) Cell dysfunction and depletion in AIDS: the programmed cell death hypothesis. Immunol Today 12: 102–105.
  85. 85. Westendorp MO, Frank R, Ochsenbauer C, Stricker K, Dhein J, et al. (1995) Sensitization of T cells to CD95-mediated apoptosis by HIV-1 Tat and gp120. Nature 375: 497–500.
  86. 86. Wong JK, Hezareh M, Gunthard HF, Havlir DV, Ignacio CC, et al. (1997) Recovery of replication-competent HIV despite prolonged suppression of plasma viremia [see comments]. Science 278: 1291–1295.
  87. 87. Finzi D, Blankson J, Siliciano JD, Margolick JB, Chadwick K, et al. (1999) Latent infection of CD4+ T cells provides a mechanism for lifelong persistence of HIV-1, even in patients on effective combination therapy. Nat Med 5: 512–517.
  88. 88. Venkatachari NJ, Buchanan WG, Ayyavoo V (2008) Human immunodeficiency virus (HIV-1) infection selectively downregulates PD-1 expression in infected cells and protects the cells from early apoptosis in vitro and in vivo. Virology 376: 140–153.
  89. 89. Brainard DM, Tharp WG, Granado E, Miller N, Trocha AK, et al. (2004) Migration of antigen-specific T cells away from CXCR4-binding human immunodeficiency virus type 1 gp120. Journal of Virology 78: 5184–5193.
  90. 90. Iyengar S, Schwartz DH, Hildreth JE (1999) T cell-tropic HIV gp120 mediates CD4 and CD8 cell chemotaxis through CXCR4 independent of CD4: implications for HIV pathogenesis. J Immunol 162: 6263–6267.
  91. 91. Poznansky MC, Olszak IT, Foxall R, Evans RH, Luster AD, et al. (2000) Active movement of T cells away from a chemokine. Nat Med 6: 543–548.
  92. 92. Masci AM, Galgani M, Cassano S, De Simone S, Gallo A, et al. (2003) HIV-1 gp120 induces anergy in naive T lymphocytes through CD4-independent protein kinase-A-mediated signaling. J Leukoc Biol 74: 1117–1124.
  93. 93. Lee C, Liu QH, Tomkowicz B, Yi Y, Freedman BD, et al. (2003) Macrophage activation through CCR5- and CXCR4-mediated gp120-elicited signaling pathways. Journal of Leukocyte Biology 74: 676–682.
  94. 94. Clerici M, Stocks NI, Zajac RA, Boswell RN, Lucey DR, et al. (1989) Detection of three distinct patterns of T helper cell dysfunction in asymptomatic, human immunodeficiency virus-seropositive patients. Independence of CD4+ cell numbers and clinical staging. J Clin Invest 84: 1892–1899.
  95. 95. Gurley RJ, Ikeuchi K, Byrn RA, Anderson K, Groopman JE (1989) CD4+ lymphocyte function with early human immunodeficiency virus infection. Proc Natl Acad Sci U S A 86: 1993–1997.
  96. 96. Hellerstein M, Hanley MB, Cesar D, Siler S, Papageorgopoulos C, et al. (1999) Directly measured kinetics of circulating T lymphocytes in normal and HIV-1-infected humans. Nat Med 5: 83–89.
  97. 97. Valdez H, Lederman MM (1997) Cytokines and cytokine therapies in HIV infection. AIDS Clin Rev 187–228.
  98. 98. Finzi D, Hermankova M, Pierson T, Carruth LM, Buck C, et al. (1997) Identification of a reservoir for HIV-1 in patients on highly active antiretroviral therapy [see comments]. Science 278: 1295–1300.
  99. 99. Lempicki RA, Kovacs JA, Baseler MW, Adelsberger JW, Dewar RL, et al. (2000) Impact of HIV-1 infection and highly active antiretroviral therapy on the kinetics of CD4+ and CD8+ T cell turnover in HIV-infected patients. Proc Natl Acad Sci U S A 97: 13778–13783.
  100. 100. Chen JJ, Cloyd MW (1999) The potential importance of HIV-induction of lymphocyte homing to lymph nodes. Int Immunol 11: 1591–1594.
  101. 101. Samstag Y, Eibert SM, Klemke M, Wabnitz GH (2003) Actin cytoskeletal dynamics in T lymphocyte activation and migration. J Leukoc Biol 73: 30–48.
  102. 102. Eibert SM, Lee KH, Pipkorn R, Sester U, Wabnitz GH, et al. (2004) Cofilin peptide homologs interfere with immunological synapse formation and T cell activation. Proc Natl Acad Sci U S A 101: 1957–1962.
  103. 103. Derry JM, Ochs HD, Francke U (1994) Isolation of a novel gene mutated in Wiskott-Aldrich syndrome. Cell 78: 635–644.
  104. 104. Samstag Y, Dreizler EM, Ambach A, Sczakiel G, Meuer SC (1996) Inhibition of constitutive serine phosphatase activity in T lymphoma cells results in phosphorylation of pp19/cofilin and induces apoptosis. J Immunol 156: 4167–4173.
  105. 105. Fenyo EM, Morfeldt-Manson L, Chiodi F, Lind B, von Gegerfelt A, et al. (1988) Distinct replicative and cytopathic characteristics of human immunodeficiency virus isolates. J Virol 62: 4414–4419.
  106. 106. Schuitemaker H, Koot M, Kootstra NA, Dercksen MW, de Goede RE, et al. (1992) Biological phenotype of human immunodeficiency virus type 1 clones at different stages of infection: progression of disease is associated with a shift from monocytotropic to T-cell-tropic virus population. J Virol 66: 1354–1360.
  107. 107. Tersmette M, Gruters RA, de Wolf F, de Goede RE, Lange JM, et al. (1989) Evidence for a role of virulent human immunodeficiency virus (HIV) variants in the pathogenesis of acquired immunodeficiency syndrome: studies on sequential HIV isolates. J Virol 63: 2118–2125.
  108. 108. Jensen MA, van 't Wout AB (2003) Predicting HIV-1 coreceptor usage with sequence analysis. AIDS Rev 5: 104–112.
  109. 109. Pastore C, Nedellec R, Ramos A, Pontow S, Ratner L, et al. (2006) Human immunodeficiency virus type 1 coreceptor switching: V1/V2 gain-of-fitness mutations compensate for V3 loss-of-fitness mutations. J Virol 80: 750–758.
  110. 110. Wu Y (2009) The co-receptor signaling model of HIV-1 pathogenesis in peripheral CD4 T cells. Retrovirology 6: 41.