CD8 T cells play an important role in protection and control of HIV-1 by direct cytolysis of infected cells and by suppression of viral replication by secreted factors. However, although HIV-1–infected individuals have a high frequency of HIV-1–specific CD8 T cells, viral reservoirs persist and progressive immunodeficiency generally ensues in the absence of continuous potent antiviral drugs. Freshly isolated HIV-specific CD8 T cells are often unable to lyse HIV-1–infected cells. Maturation into competent cytotoxic T lymphocytes may be blocked during the initial encounter with antigen because of defects in antigen presentation by interdigitating dendritic cells or HIV-infected macrophages. The molecular basis for impaired function is multifactorial, due to incomplete T-cell signaling and activation (in part related to CD3ζ and CD28 down-modulation), reduced perforin expression, and inefficient trafficking of HIV-specific CD8 T cells to lymphoid sites of infection. CD8 T-cell dysfunction can partially be corrected in vitro with short-term exposure to interleukin 2, suggesting that impaired HIV-specific CD4 T helper function may play a significant causal or exacerbating role. Functional defects are qualitatively different and more severe with advanced disease, when interferon γ production also becomes compromised.

During primary HIV-1 infection, plasma viremia increases to peak levels of up to 109 HIV-1 copies/mL. Viremia is thereafter partially controlled by the immune response and stabilizes 3 to 6 months later. However, in situ hybridization for HIV-1 RNA and quantification of cell-associated unspliced messenger and genomic RNA and DNA indicate that HIV-replicating and latently infected cells in lymphoid tissue increase over time.1,2 The majority of cells replicating HIV-1 are CD4 T cells, both early and late in disease.3 

Viral replication persists in lymphoid tissue despite a vigorous immune response. HIV-1–specific CD8 T cells, which should be able to target virally infected cells for elimination, exist at high frequencies in most HIV-1–infected patients. Previous limiting dilution estimates of the frequency of HIV-specific CD8 T cells reached around one per thousand cells.4,5 However, more sensitive techniques that use tetramer staining, which fluorescently labels T cells whose T-cell receptors (TCRs) recognize an antigenic peptide-major histocompatibility (MHC) pair, have shown that up to a few percent of circulating CD8 T cells recognize a particular HIV-1 peptide.6 The frequency of CD8 T cells producing interferon γ (IFNγ) in response to major HIV-1 proteins expressed by vaccinia virus ranges from 0.8% to 18%.7 In response to HIV-1–infected primary CD4 T cells, approximately 0.3% to 3% of HIV-seropositive donor CD8 T cells produce IFNγ.8 In those studies, the frequency of circulating HIV-specific CD8 T cells is generally above 1%. These results provide lower limits for the frequency of HIV-specific CD8 T cells because, as discussed below, not all specific cells are triggered to produce IFNγ.8 

After the first week of infection when viral dissemination is reined in by the innate immune response, viral-specific CD8 T cells form the cornerstone of immune control for most viruses, including HIV-1. Viral-specific cytotoxic T lymphocytes (CTLs) suppress HIV-1 replication in vitro by direct cytotoxicity and secretion of soluble factors.9-13 Acute viremia resolves with the appearance of HIV-specific CTLs.14,15 Moreover, the likelihood of progressing to AIDS increases in patients who lackgag-specific CTLs.4,16 This finding, together with the rapid disease course in neonates whose overall T-cell immunity is immature,17 suggests that CTLs are important in controlling HIV-1. Direct evidence was provided in rhesus macaques in which elimination of CD8 T cells results in a dramatic increase in simian immunodeficiency virus (SIV) load.18,19 

If there are so many HIV-1–specific CD8 T cells, why are these cells not doing a better job? In the absence of antiviral drugs, the immune response fails to halt progressive immunodeficiency in all but a small minority of long-term nonprogressor (LTNP) patients. A persistent viral reservoir might be explained by the fact that HIV-1 can establish latent cellular infection in which HIV-1 provirus is integrated into the host genome, but no viral proteins are produced to signal to roving T cells. However, the progressive nature of the disease suggests that antiviral CD8 T cells do not always efficiently destroy cells replicating HIV-1. As many as 10 billion virions are produced each day in asymptomatic, untreated individuals.20,21 The failure of CD8 T-cell function is clear when one compares the frequency of HIV-specific CD8 T cells with the numbers of HIV-1–infected cells replicating virus, generally substantially below one per thousand and frequently on the order of one per million mononuclear cells. Because the frequency of HIV-specific CD8 T cells is typically at least 10-fold higher and because in infected individuals there are usually more CD8 T cells than CD4 T cells and macrophages, the inability to control viral production is striking, especially because CTLs are “serial killers” able to lyse multiple cells.22,23 

Failure to control viral replication suggests that antiviral CD8 T cells are impaired in lysing HIV-1–infected cells and suppressing HIV-1 replication. Although early reports suggested that peripheral blood mononuclear cells (PBMCs) from 80% to 90% of HIV-1–infected donors demonstrate HIV-specific cytotoxicity above background in 6- to 16-hour assays,10,24,25 specific cytotoxicity by freshly isolated PBMCs is generally not much above background in direct 4-hour51Cr release assays.8,26-28 However, HIV-specific cytotoxicity by PBMCs is dramatically enhanced after overnight culture in an interleukin-2 (IL-2)–dependent manner.26 The rapid increase in cytotoxicity, typically to 10% to 30% specific cytotoxicity in a 4-hour assay at an effector/target (E:T) ratio of 25:1 to 50:1, cannot be explained by overnight clonal expansion of viral-specific T cells. A possible specific defect in HIV-specific cytotoxic capability was recently demonstrated by comparing HIV- and CMV-specific lysis by freshly isolated blood lymphocytes.28 HIV-specific lysis was considerably less than CMV-specific lysis, despite comparable numbers of tetramer-staining HIV-specific and CMV-specific CD8 T cells.

Failure of cytolysis could arise either from resistance of HIV-1–infected targets to CTL recognition or lysis29,30or from a lack of effector function by the CTLs themselves.26 Many viruses have strategies to evade CTL recognition, especially by interfering with antigen presentation (reviewed in Tortorella et al31). Two HIV-1 evasion strategies have been described. Viral mutation of the epitopic sequence recognized by HIV-specific CTL can sidestep CTL recognition.29,32,33 Moreover, mutated epitopes (altered peptide ligands) can interfere with subsequent triggering by cells expressing wild-type sequence.34 Although such mutations certainly exist, it is uncertain whether they play an important role. Convincing recent evidence for a role for escape mutation comes from the rapid emergence of viral mutations of recognized tat epitopes during acute SIV infection.35 

Another viral evasion mechanism is nef-induced down-modulation of MHC class I A and B molecules.36-40 Assays that measure T-cell functional responses to HIV-1–infected primary cells are invaluable for determining whether viral evasion significantly interferes with T-cell immune responses.8,40-43 In fact, CD8 clones that recognize 4 different HIV-1 proteins lyse HIV-1–infected primary CD4 T cells as well as they lyse cells infected with recombinant vaccinia virus expressing HIV-1 genes.43 Moreover, nef is a major antigen recognized by HIV-seropositive donor CD8 T cells.44 In addition, CD8 T cells can recognize target cells expressing only a few antigenic peptide-MHC molecules45 and the HLA class I C allele is unaffected by nef.46 Those results suggest that nef-mediated down-modulation of class I alleles may not block immune recognition of HIV-1–infected cells. Nef-mediated down-modulation of class I molecules, however, might be important for CD8 T cells with low-affinity TCRs or for antigens, synthesized at low levels or inefficiently processed and presented.30 

Because lysis of HIV-1–infected cells is dramatically up-regulated after brief exposure to IL-2 in vitro,26there is also a functional in vivo problem with HIV-specific CTLs. Previously activated antigen-specific T cells may be unable to proliferate or to perform the full complement of effector functions when they re-encounter antigen. Lack of proliferation or function is termed anergy.47 Although anergy can be defined as complete unresponsiveness on re-encounter with antigen, in reality cells more often become partially anergized. For example, a CD4 T cell may lose the ability to proliferate to antigen but may still express activation molecules and secrete some cytokines. Anergy is important to prevent immunopathology and autoimmunity. Anergy has been studied extensively for CD4 T cells, but less is known about regulation of CD8 T-cell function.

HIV-1–specific CD4 T-cell proliferation is undetectable within 3 months after primary infection in most HIV-1–infected individuals, with the exception of some LTNPs and early treated patients.48-54 However, HIV-specific CD4 T cells are not deleted but are present in low numbers in a partially anergized state.55 In response to HIV-1 gag p24, approximately 0.12% (range, 0%-0.66%) of CD4 T cells produce IFNγ by intracellular flow cytometry analysis. The median response is higher in nonprogressive HIV-1 disease (0.40%). However, those figures are about an order of magnitude lower than the corresponding CD8 T-cell activity, especially when the reduced numbers of CD4 T cells are considered.55 Moreover, CD4 proliferation to HIV-1 p24 can be induced in vitro by adding CD40L-trimer and IL-12, providing further support for persistence of anergized HIV-specific CD4 cells.56 

CD4 T cells are important to maintain effective antiviral CD8 CTLs. They secrete IL-2, an important growth factor for CD8 cell survival. However, high concentrations of IL-2 facilitate T-cell activation–induced apoptosis. The high frequency of HIV-specific CD8 T cells may be partly due to higher concentrations of stromal cell and monocyte-derived cytokines, such as IL-7, IL-12, and IL-15, which promote T-cell survival but not apoptosis, compared with CD4 T-cell–derived IL-2.57-62 CD4 T cells also express molecules whose ligation is directly or indirectly required for CTL differentiation. For example, ligation of CD40L on CD4 T cells activates dendritic cells (DCs) to provide signals for development of CD8 CTLs.63-65 Although antiviral CD8 T cells persist in CD4 T-cell–depleted mice, their function is compromised.66,67 In this review we argue that CD8 T-cell anergy contributes significantly to the progressive immunodeficiency of untreated HIV-1 infection, as was first suggested by Miedema68 and Lewis et al.69 

Lack of HIV-specific CD4 proliferation and CD8 cytotoxicity is usually not overcome by antiretroviral therapy. As was originally suggested by Macatonia et al70 and more recently by Shearer,71 CD4 and CD8 T-cell anergy in HIV-1 infection may be a result of changes in antigen presentation by DCs or macrophages (Figure 1). Interdigitating DCs bring antigen from peripheral tissues to the lymph nodes (LNs), where they activate naive T cells. Mature DCs strongly up-regulate MHC class II and costimulatory molecules72such as CD40, CD80, CD86, OX40L,73 and 4-1BBL,74 required for effective stimulation of a primary T-cell response. Although monocytes and macrophages are infected with HIV-1, whether DCs are infected remains controversial.75-79 HIV-1 binds to DCs through the association of gp120 with a recently described C-type lectin receptor, DC-SIGN. Tight binding of HIV to DCs facilitates CD4 T-cell infection.80,81 Because DC-SIGN is important for DC-mediated T-cell activation, HIV-1 particles or free gp120 might interfere with primary T-cell activation. However, this is just a conjecture. In fact, freshly isolated DCs from spleen or blood of HIV-1–infected donors have no significant differences in MHC class II or costimulatory molecule expression and are not defective antigen-presenting cells (APCs) after in vitro maturation.76,82-87 Although CD11c+ DCs are reduced in the blood of HIV-1–infected individuals, this reduction may be due to their migration to LNs, where they are increased.88 However, the APC function of cytokine-activated monocyte-derived DCs may have little relation to the function of DCs in vivo. Expression of critical costimulatory molecules on interdigitating DCs in situ in the paracortical regions of LNs needs to be studied.

Fig. 1.

Dendritic cells and macrophages in HIV-infected patients may hinder antigen presentation.

(A) Binding of HIV-1 gp120 to DC-SIGN on DCs enhances infection of CD4 T cells and may interfere with their activation. (B) Macrophages from HIV-infected patients have reduced expression of molecules required for effective stimulation of T cells and have increased expression of molecules that can trigger apoptosis. Macrophages with defective APC function are not necessarily infected with HIV. The size of the labels in this figure reflects the relative expression level on macrophages from healthy donors and HIV-infected donors. Part B modified from Shearer71 copyright 1998 and reprinted with permission from Elsevier Science.

Fig. 1.

Dendritic cells and macrophages in HIV-infected patients may hinder antigen presentation.

(A) Binding of HIV-1 gp120 to DC-SIGN on DCs enhances infection of CD4 T cells and may interfere with their activation. (B) Macrophages from HIV-infected patients have reduced expression of molecules required for effective stimulation of T cells and have increased expression of molecules that can trigger apoptosis. Macrophages with defective APC function are not necessarily infected with HIV. The size of the labels in this figure reflects the relative expression level on macrophages from healthy donors and HIV-infected donors. Part B modified from Shearer71 copyright 1998 and reprinted with permission from Elsevier Science.

Close modal

Although no clear DC functional deficit has been demonstrated, monocytes from HIV-infected donors aberrantly express molecules important for antigen presentation to T cells. HIV-infected macrophages are important for restimulating antigen-experienced HIV-specific T cells. Monocyte/macrophages from HIV-infected donors have reduced MHC class II and costimulatory CD80 and CD86 expression.89-91Moreover, they have up-regulated fas and fasL as well as CD16, molecules that may trigger apoptosis of HIV-specific T cells instead of activating a protective immune response.92,93 In vitro HIV-1 infection of macrophages also up-regulates fasL.92,94 Moreover, HIV-1 tat down-modulates transcription of the mannose receptor, important for binding pathogens and their efficient internalization for antigen presentation.95 

Monocyte-derived macrophages from HIV-infected donors are also likely to be impaired in inducing TH1 and CTL responses because they produce less IL-12 and more IL-10.91,96 This may be secondary to reduced CD4 helper function, because trimeric CD40L can help restore IL-12 production in vitro, by substituting for CD40L on activated CD4 T cells.97 CD40L is required to activate APCs to produce IL-12 and is required for generating functional CTLs.63-65 Interestingly, the combination of CD40L-trimer and IL-12 reverses CD4 and CD8 T-cell anergy in vitro.56,98 

CTLs kill cells either by the granule exocytosis pathway (by releasing cytotoxic granules that contain pore-forming perforin [PFP] and granzyme proteases) or by engagement of death domain–containing receptors, such as fas or the tumor necrosis factor receptor (reviewed in Henkart99). The granule-mediated pathway, critical for the immune defense against many viral infections, is absolutely dependent on both PFP and granzymes.100 Mice genetically deficient in PFP succumb to viral infections such as lymphocytic choriomeningitis virus.101 Children with familial hemophagocytic lymphohistiocytosis, linked to a defective perforin gene, often die of overwhelming viral infections.102 Mouse experiments have also shown that IFNγ production by viral-specific CD8 T cells is important to control some viral infections.103-105 Although HIV-1–infected CD4 T cells express twice as much fas as uninfected activated cells, HIV-1–infected primary T cells are predominantly lysed by the PFP-dependent granule-mediated pathway.43 This was shown by using HIV-specific CD8 T-cell lines and clones responding to a variety of HIV-1 proteins. Moreover, HIV-specific CD8 T cells do not express the principal death receptor ligand fasL, even after stimulation with HIV-1–infected CD4 cells.43 

Phenotypic analysis of tetramer-stained, HIV-specific CD8 T cells indicates that circulating viral-specific CD8+ T cells have been previously activated, because they no longer express cell surface markers, such as CD62L, CCR7, CD45RA, and CD28, present on naive cells, and also express CD45RO, granzyme A (GzmA), and bcl-2, not found in naive cells.6,8,28,106,107 (Table1). This is not surprising, given the high likelihood of prior encounter with HIV-1–infected cells. However, protein expression profiles suggest that HIV-specific CD8 T cells either have not fully differentiated into mature effector CTLs, have reverted to a memory type, or have become anergized. Another possibility is that specific cells die before they fully mature; however, this is unlikely, given the high frequency of HIV-specific CD8 T cells. Although no single surface marker unambiguously identifies CTLs, differentiation into effector CTLs is associated with down-modulation of CD27 and CD28 and re-expression of CD45RA.108-110 Although HIV-1 tetramer+ cells are CD28, they are uniformly CD27+ and CD45RA in most donors.8,28,111 These properties are distinct to HIV-specific cells, because they do not hold for cells specific for Epstein-Barr virus (EBV) and cytomegalovirus (CMV) in the same donors.28,107 

Table 1.

Phenotypic properties of tetramer+ CD8 T cells

Cell markerHIV-1 tetramer+ cellsEBV and CMV tetramer+ cells
CD3ζ − −  
CD28 − − 
NKRs ?  
CD62L − −  
CD45RA − ± to + 
CCR7 − −  
CD27 −  
GzmA 
Perforin − to ± − to ± to +  
Bcl-2 
Cell markerHIV-1 tetramer+ cellsEBV and CMV tetramer+ cells
CD3ζ − −  
CD28 − − 
NKRs ?  
CD62L − −  
CD45RA − ± to + 
CCR7 − −  
CD27 −  
GzmA 
Perforin − to ± − to ± to +  
Bcl-2 

The phenotypic properties of the majority of HIV-specific CD8 T cells are compared with those for other chronic viral infections. Data are a composite from published6,8,28,106,107,111,150 and unpublished (P.S., G. Chen, and J.L., March 2001) studies.

EBV indicates Epstein-Barr virus; CMV, cytomegalovirus; NKR, natural killer cell-inhibitory receptor; GzmA, granzyme A; ?, not known.

Naive CD8 T cells have limited function on their initial encounter with antigen; they are not cytolytic and do not produce effector cytokines rapidly. It takes approximately 5 days to differentiate into effector CTLs and to express cytolytic molecules.112 PFP and granzymes are up-regulated in parallel.112 Circulating CD8 T cells in HIV-1–infected subjects, compared with healthy donors, have an unusually high percentage of CD8 cells containing cytolytic granules, as measured by staining for the most abundant granule protease, GzmA.26 The overwhelming majority of HIV-1 tetramer+ CD8 T cells express GzmA.8,43However, there is a disjunction between PFP and GzmA protein expression in the LNs113 (Figure 2). Although granzyme+ CD8 T cells are increased in HIV-1–infected LNs compared with normal LNs, PFP+ cells are rare. This finding holds for both acute and chronic HIV-1 samples and is in contradistinction to acute infectious mononucleosis LNs in which PFP and GzmA are both increased.114 Lack of LN PFP-expressing CD8 T cells is not due to recent degranulation, because degranulation would deplete granzymes as well as PFP. PFP is also not expressed at high levels in circulating tetramer+HIV-specific CD8 T cells.28 Lack of PFP seems to affect HIV-specific cells preferentially, because about a third of all circulating CD8 T cells (34% ± 20%) are PFP+ in a diverse spectrum of HIV-1–infected patients.107 Moreover, in another study HIV-1 tetramer+ cells are PFP, whereas CMV tetramer+ cells in the same patient are PFP+.28 Because PFP is required to lyse HIV-1–infected targets,43 lack of PFP may be critical for impaired cytotoxicity, especially at LN sites of viral production.

Fig. 2.

Perforin staining is reduced compared with staining for GzmA in adjacent LN sections from a donor undergoing primary HIV infection.

In contrast, comparable levels of granzymes A and perforin were found in samples from acute infectious mononucleosis. Figure taken from Andersson et al113. Copyright 1999 and reprinted with permission from Lippincott Williams & Wilkins.

Fig. 2.

Perforin staining is reduced compared with staining for GzmA in adjacent LN sections from a donor undergoing primary HIV infection.

In contrast, comparable levels of granzymes A and perforin were found in samples from acute infectious mononucleosis. Figure taken from Andersson et al113. Copyright 1999 and reprinted with permission from Lippincott Williams & Wilkins.

Close modal

These results require further verification. The absence of PFP in HIV-specific CD8 T cells compared with CD8 T cells specific for another virus was reported in only one study. A word of caution also is necessary before generalizing from results for HIV-1 tetramer+ cells to conclusions about all HIV-specific CD8 T cells. HIV-1 epitopes for which tetramers are available are rarely immunodominant.115-118 CD8 T cells responding to subdominant or cryptogenic epitopes may have lower avidity interactions with target cells than T cells in the immunodominant response.119 They may revert to CD45RAPFP memory cells if they are not effectively activated by HIV-infected cells, especially if the viral epitope has mutated. Low avidity interactions may alter T-cell activation and differentiation, as has been shown for murine CD4 T cells.120 Therefore, lack of PFP in HIV-specific CD8 T cells needs to be confirmed by using a broader array of tetramers or other methods.8,26-28 

CD8 T cells also suppress viral production by secreting cytokines like IFNγ, inhibitory chemokines, such as macrophage inflammatory protein 1α (MIP-1α), MIP-1β, and regulated on activation, normal T-cell expressed and secreted (RANTES), and other factors, some of which, such as CD8+ cell antiviral factor (CAF), remain to be characterized.9-13 CAF is produced by CD28+ CD8 T cells, which are substantially reduced in HIV-1 infection.121 The inhibitory chemokines, and perhaps other suppressive molecules, are stored in cytotoxic granules with PFP and granzymes.122 Leukemia inhibitory factor, also produced by CD8 T cells, is a recently described potent HIV-1 suppressor factor, active at concentrations 100-fold lower than the β-chemokines.123 

IFNγ is important in antiviral immunity in genetically targeted murine models.124-126 In more advanced patients, fewer cells produce IFNγ in response to HIV-1, and the number of IFNγ-producing cells initially increases with highly active antiretrovial therapy.127-129 On a single cell basis, IFNγ production in response to HIV-1–infected cells is impaired, especially in late-stage patient samples.7,8,130 In one study, IFNγ production was measured by intracellular staining after stimulation with HIV-1–infected CD4 blasts in the presence or absence of IL-28 (Figure 3). In subjects with undetectable plasma viremia, adding IL-2 had no effect on IFNγ production. In some subjects with measurable plasma viremia, IL-2 increased the frequency of IFNγ-producing cells 4- to 5-fold. IFNγ production was also detected after IL-2 was added to some samples that had no IFNγ-producing cells without IL-2. Therefore, in more advanced patients, IFNγ production is compromised. In another approach, the numbers of HIV-1 tetramer+ CD8 T cells were compared with the numbers of cells producing IFNγ after stimulation with the tetrameric peptide. In mice, virtually every specific memory cell produces IFNγ in response to antigenic peptide.131,132 In most HIV-infected subjects, however, fewer than 25% of tetramer+ cells produce IFNγ after stimulation with tetrameric peptide. Subjects with a high proportion of IFNγ-producing tetramer+ cells are asymptomatic and have well-maintained CD4 counts (> 650/μL). Most subjects with less than 25% IFNγ-producing tetramer+ cells have had HIV-1–related stage B or C symptoms. Culturing tetramer+cells in IL-15 and IL-2 converts them from nonresponsive to IFNγ-producing cells.8 

Fig. 3.

The number of IFNγ-producing CD8 T cells in response to HIV-1–infected CD4 T cells increases substantially in some HIV-1–infected patient samples when IL-2 is added during the stimulation.

Background IFN production in the presence of IL-2 but not infected CD4 T cells was low (< 0.05%) and was subtracted to get the frequency of specific producers. Data are taken from Shankar et al.8 

Fig. 3.

The number of IFNγ-producing CD8 T cells in response to HIV-1–infected CD4 T cells increases substantially in some HIV-1–infected patient samples when IL-2 is added during the stimulation.

Background IFN production in the presence of IL-2 but not infected CD4 T cells was low (< 0.05%) and was subtracted to get the frequency of specific producers. Data are taken from Shankar et al.8 

Close modal

Therefore, IFNγ production in response to HIV-1, like cytotoxicity, is partially impaired in vivo, especially in later-stage subjects, because adding exogenous cytokines increases it. A few studies have not found impaired IFNγ production in HIV-1 infection, but they have either focused on donors with well-controlled or early infection28 or measured production following stimulation with antigen plus auxiliary costimulatory agents to amplify IFNγ production.133 A recent study in rhesus macaques infected with SIV mac239 confirms and illuminates the human data.134 Although virtually all tetramer+ CD8 T cells produce IFNγ after vaccination or early after infection, within 6 months the proportion of specific cells that produce IFNγ drops substantially. Loss of IFNγ production by antigen-specific CD8 T cells infiltrating melanoma tumors or circulating in melanoma patients has also been described.135-137 

The T-cell response can be dissected into a series of steps, any of which might be impaired in HIV-specific CD8 T cells. The TCR complex is composed of a clonotypic dimeric TCR, required for antigen recognition, in noncovalent association with CD3, a multicomponent signal transduction complex. CD3 consists of CD3δ, CD3ε, and CD3γ chains and a CD3ζ chain–containing dimer (reviewed in Weiss and Littman138). When the TCR encounters antigen on an infected cell, a signal is sent into the cell via sequential tyrosine phosphorylation of key TCR-associated signaling molecules. This signal initiates formation of an “immunologic synapse” between the T cell and its target. The TCR-CD3 complex, CD28 and associated signaling molecules, such as lck, fyn, and protein kinase Cθ, cluster in the center of the synapse, surrounded circumferentially by larger molecules like CD2 and LFA-1, which stabilize the synapse. The synapse forms within minutes of TCR engagement and lasts for more than an hour until the entire TCR complex is internalized.139 Synapse formation critically depends on the actin cytoskeleton to move molecules in and out of the forming synapse.140 CD3ζ is central to transmitting the TCR activation signal. It contains 3 immune-receptor tyrosine-activating motifs, which can be phosphorylated and serve as docking sites for SH2 domain–containing tyrosine kinases. CD3ζ also links TCR-CD3 to actin.141,142 

In antigen-experienced CD8 T cells, cytokine secretion and cytolysis are triggered rapidly following TCR engagement.143Cytolytic granules, which normally contain granzymes, PFP, and some other components, are transported toward the immunologic synapse, and the granule membrane fuses to the CTL plasma membrane, releasing its contents. Failure of cytolysis by HIV-specific CD8 T cells could be due to lack of recognition of HIV-1–infected primary cells by the TCR, failure or incomplete signaling by the TCR-CD3 complex, failure of degranulation, absence of key cytolytic components such as PFP in the cytolytic granules, or resistance of HIV-1–infected cells to cytolysis. Because β-chemokines are also stored in CTL granules,122 viral suppression would probably also be impaired if granule exocytosis is blocked.

HIV-specific tetramer+ CD8 T cells uniformly have down-modulated both CD3ζ and the principal costimulatory molecule CD28.26,111,144 Down-modulation of CD3ζ was first described in lymphocytes infiltrating murine tumors and has since been found in a variety of human cancers.145-148 Not surprisingly, circulating T cells of HIV-1–infected subjects also have signaling defects.111,149 Using immunoprecipitation and Western blotting, Stefanova et al149 found that kinase activities of lck, fyn, and ZAP70 were decreased in HIV-1–infected patients at varying disease stages but not in a group of LTNPs. CD3ζ down-modulation was also found by Western blotting in T cells from both acutely infected donors and AIDS patients.111,149 However, down-modulation of CD3ζ and CD28 also occurs in CD8 T cells during other viral infections and therefore, is part of normal immune regulation.150 CD3ζCD28CD8 T cells from HIV-1–infected and healthy donors have a selective defect in activating the IL-2 axis—they neither express the α-chain of the high affinity IL-2 receptor (CD25) nor produce IL-2.111,150 This finding may be especially critical in HIV-1 infection in which there is a special deficit in HIV-stimulated production of IL-2 by CD4 T cells.

Down-modulation of CD3ζ and CD28 on most HIV-specific CD8 T cells almost certainly raises the activation threshold for functions induced by TCR engagement.111,150,151 The threshold for some functions, such as cytotoxicity, may be higher than for other functions, such as IFNγ production. In fact, CD8 T cells with down-modulated CD3ζ and CD28 produce IFNγ but not IL-2 after TCR activation.111,150 Different activation thresholds also characterize CD4 T-cell cytokine production. Antigen-experienced CD4 T cells are heterogeneous in the amounts of cytokines produced and in the signals required to activate secretion.152 Cells that do not produce cytokines may be triggered in the presence of activated APCs or of antibodies to costimulatory pathways (CD28, CD49d, and CD5).153 Raising the threshold for cytotoxicity protects the host from immunopathogenic consequences of lysing uninfected cells expressing the low-affinity self-antigens on which the T cell was selected in the thymus. The requirement for participation of nearby antigen-specific helper cells may also help ensure that the target is indeed infected.

HIV-specific cytotoxicity and IFNγ production are greatly enhanced by adding more than 100 IU/mL IL-2.8,151Cytotoxicity requires IL-2 in most patient samples; IFNγ secretion is enhanced by IL-2, especially in advanced patients. Therefore, thresholds have been raised to different levels for different functions. A high IL-2 concentration triggers the low-affinity IL-2R, but much lower concentrations (∼1 IU/mL) trigger the high-affinity IL-2R. Because activated CD3ζ–down-modulated cells do not express high-affinity IL-2R, high IL-2 concentrations are needed to improve their function. Therefore, in vivo high-dose IL-2 therapy in HIV infection may do more than increase CD4 T cell numbers; it may also improve CD8 antiviral function.154 This possibility should be examined.

Correction of CD8 T-cell function with IL-2 suggests that CD4 T help may be important for protective CD8 T-cell function, as it is in mouse lymphocytic choriomeningitis virus infection.67 In HIV-infected humans, who lack a proliferative response to CMV or HIV, antigen-specific CD8 T cells persist, as they do in CD4 T- cell–depleted mice.155 Just as antigen-specific CD8 T cells are not protective in CD4 T-cell–depleted mice, lack of effective HIV-specific CD4 help may also be responsible for lack of immune protection by HIV-specific CD8 T cells.51,151,156-158 Although this hypothesis is attractive, supporting data are mostly indirect.

Another possible contributory factor to CD8 T-cell dysfunction in vivo could be inhibitory signaling by natural killer (NK) cell-inhibitory receptors (NKRs), expressed on some CD8 T cells after activation.159 The receptors belong to 2 distinct molecular types: (1) the immunoglobulin superfamily killer inhibitory receptors (KIRs), which recognize specific HLA allotypes with common structural features in the α-1 domain, and (2) C-type lectin receptors in the CD94 family, which display broad specificity for HLA class I molecules through lectin interactions.160-162CD94/NKG2-(A-C) and NKG2D lectinlike receptors, respectively, recognize HLA-E (not down-modulated by HIV-1 nef46) and MHC class I chain–related A (MICA). Additional immunoglobulinlike receptors (ILT receptors) are expressed by activated T cells and other leukocytes, and they interact broadly with class Ia molecules. Within the receptor families, homologues, which bind the same ligands, have opposite inhibitory or stimulatory activity. However, the inhibitory receptors usually have higher affinity. CD8 T cells expressing NKRs have oligoclonal TCRs, suggesting that they correspond to antigen-specific cell expansions. Recently, NKRs on melanoma-specific CD8 T cells have been shown to inhibit both cytotoxicity and IFNγ production.137 

Large numbers of CD8 T cells express NKRs in HIV-1 infection. Moreover, NKR expression is up-regulated by cytokines, such as IL-10, IL-15, and tumor growth factor β, which are increased in HIV-1 infection.159,163,164 In HIV-infected donor samples, blocking NKR engagement with specific antibodies augments in vitro HIV-specific cytotoxicity.165 Therefore, it makes sense to explore a role for NKR inhibitory signaling in blocking HIV-specific CD8 T-cell function. A first step would be to show increased expression of NKRs on HIV-1 tetramer+ CD8 T cells.

Another reason for incomplete HIV-1 control by CD8 T cells could be inefficient access to sites of HIV-1 infection in LNs. T-cell activation in vivo is associated with changes in surface phenotype, which reflect alterations in migration and functional capability. Initial activation of naive CD8 T cells occurs in T-cell zones of LNs where naive cells encounter DCs bearing the antigen, which fits its TCR. Naive T cells continuously recirculate from blood to LNs through specialized high endothelial venules (HEVs). Sequential engagement of L-selectin (CD62L) and LFA-1 on naive T cells to their respective ligands on HEVs sets into motion tethering, rolling, and firm adhesion to HEVs, prerequisite steps for transmigration into the surrounding LN (reviewed in Springer166). A chemokine receptor CCR7 has also emerged as an important determinant of T-cell homing to LNs.167-169 CCR7 on T cells interacts with the chemokine secondary lymphoid tissue chemokine on HEVs and regulates LN homing by delivering an activation signal for LFA-1 binding.169-171Studies in CCR7 knockout mice suggest that CCR7 is also important for the movement of T cells within LNs. If the naive cell encounters antigen, it proliferates and differentiates into effector cells in the specialized microenvironment of the LN. In keeping with the requirement of effector/memory cells to function at peripheral sites of infection, previously activated CD8 T cells up-regulate adhesins, including CD44 and β integrins, which contribute to their preferential homing to inflamed tissues, from which naive cells are normally excluded.172-175 The LN homing receptors L-selectin and CCR7 are down-modulated on effector cells but are heterogeneously expressed on memory cells, suggesting that memory cells recirculate through LNs as part of immune surveillance.110 

In HIV-1 infection there is a marked reduction in CCR7+ naive and long-term memory cells.107HIV-1, EBV, and CMV tetramer+ CD8 T cells from healthy donors and HIV-1–infected donors, even with well-controlled disease, do not express the lymphoid homing molecules CCR7 and CD62L.107 Exclusion of effector cells from LNs, which normally provides efficient division of labor among CD8 T cells and may protect LNs from damage by inflammatory cytokines and cytolytic enzymes, may work to the detriment of the host during infections, such as HIV-1 or EBV, that target lymphocytes by converting LNs into immunologically privileged sites. In fact, in paired PBMC and LN aspirate samples from HIV-infected donors, there were 3- to 4-fold fewer EBV tetramer+ cells in the LN than in the circulation, despite the concentration of EBV infection in lymphoid tissues.107 In acute SIV infection in macaques, SIV-specific CD8 T cells are also preferentially excluded from LNs; there are proportionately 4-fold fewer SIV tetramer+ cells in LNs than in the circulation in the first weeks after infection (P < .007).176 

Circumstantial evidence also supports selective exclusion of HIV-specific CD8 T cells from lymphoid sites of infection. The proportion of CD8 T cells in the LNs of HIV-infected patients is about half of what it is in the circulation.107,177,178Moreover, the number of copies of spliced and unspliced viral RNA per cell is much higher in LNs than in PBMCs, as is the frequency of cells carrying infectious virus, suggesting a greater failure of protective immunity in the LN.177 During primary and chronic infection, perturbations of T-cell receptor Vβ repertoire, indicative of expansions of antigen-specific CTLs, are more pronounced in blood CD8 T lymphocytes than in the LN.179,180 

Exclusion of CCR7 T cells from the LNs in HIV-1 infection is probably leaky. Gene-marked HIV-specific CD8 T-cell clones infused into HIV-1–infected subjects can be identified in LNs.181 How efficiently they traffic there is unclear, because billions of cells are administered in these clinical studies. The requirement of CD62L and CCR7 expression for lymphocyte homing via HEVs might also be superseded in an inflammatory setting. Moreover, some tissue-homing lymphocytes can enter LNs through afferent lymph. Thus, additional human LN studies are required to determine how much lack of LN homing molecules interferes with the ability of HIV-specific cells to get to infection sites.

HIV-specific CD8 T cells have multiple properties that might explain their inadequate immunoprotective role in HIV-1 infection: (1) they are not cytotoxic, (2) they are impaired in trafficking to the major LN sites of viral replication, (3) they do not produce IL-2 and cannot respond to low concentrations of IL-2, and (4) they have down-modulated key molecules for T-cell signaling (Figure4). Moreover, CD8 T-cell function may be compromised even more in late-stage patients, when some of the molecular defects no longer correct in vitro with exogenous IL-2111 or when cytokine production is impaired. We now need to understand the relative contributions of the recently described novel mechanisms for CD8 T-cell dysfunction and their etiology. Closer study of the properties of HIV-specific CD8 T cells in LTNPs will help sort out what aspects of dysfunction are especially important for effective immune protection. Some of CD8 T-cell dysfunction is part of normal immune regulation of CD8 T cells to prevent these serial killers from wreaking damage by the release of cytolytic enzymes or inflammatory cytokines. For example, down-modulation of CD3ζ and CD28 and of LN homing receptors on antigen-specific CD8 T cells occurs in other infections.107,150 Although much research has focused on how CD4 T cells are regulated to prevent autoimmunity, regulation of CD8 T cells is largely terra incognito. Understanding CD8 T-cell regulation is critical for understanding immune pathogenesis and formulating strategies for immune intervention for HIV-1 and other diseases, including cancer.

Fig. 4.

Model of mechanisms that may contribute to lack of protection by antiviral CD8 T cells in HIV infection.

(1) HIV-1–infected cells are not recognized because of viral escape mutations or nef-mediated down-modulation of class I molecules, (2) signaling is impaired because of CD3ζ and CD28 down-modulation, (3) cytotoxicity is inhibited by expression of NK inhibitory receptors (KIR), (4) cytotoxicity is ineffective because of lack of PFP in cytotoxic granules, and (5) antigen-specific CTLs do not efficiently reach sites of infection because of down-modulation of homing receptors. Figure modified from Chen et al.107 

Fig. 4.

Model of mechanisms that may contribute to lack of protection by antiviral CD8 T cells in HIV infection.

(1) HIV-1–infected cells are not recognized because of viral escape mutations or nef-mediated down-modulation of class I molecules, (2) signaling is impaired because of CD3ζ and CD28 down-modulation, (3) cytotoxicity is inhibited by expression of NK inhibitory receptors (KIR), (4) cytotoxicity is ineffective because of lack of PFP in cytotoxic granules, and (5) antigen-specific CTLs do not efficiently reach sites of infection because of down-modulation of homing receptors. Figure modified from Chen et al.107 

Close modal

Most tetramer+ HIV-specific CD8 T cells, unlike many EBV- or CMV-specific cells, do not have an effector CTL phenotype. Factors specific for HIV-1 infection that might influence CTL function are the continuous presence of high levels of infectious and noninfectious viral antigens, the paucity of functioning viral-specific CD4 helper cells, the effect of sustained inflammation on antigen-presenting function of macrophages and DCs, and the effect of particular HIV-1 gene products, such as tat or nef. A possible role of HIV proteins in blocking CD8 T-cell maturation into effector CTLs has yet to be studied. HIV-1 infection especially targets the helper CD4 T cells that are important in regulating effector CTLs. In mouse models in which CD4 T cells have been eliminated, protection against infection is lost, even though some antigen-specific CD8 T cells persist but with compromised function.66,67 These models may provide helpful parallels for HIV-1 infection, where CD8 T-cell protection may well correlate with preservation of HIV-specific CD4 T-cell function.156,182 This link needs to be studied in more depth by moving beyond counting the numbers of HIV-specific CD8 T cells to studying their function with assays that avoid more than a few hours of in vitro culture. Even after overnight culture, many CD8 functional defects correct, giving the false impression of functional competence. Studies that show no correlation between clinical status (such as LTNP) and numbers of antigen-specific CD8 T-cells need to be re-evaluated in light of new findings that many HIV-specific cells may not function in vivo.

Supported by grants AI-42519 and AI-45406 (J.L.) and AI-41536 (J.A.) from the National Institutes of Health and by grant 10850 (J.A.) from the Swedish Medical Research Council.

@ 2001 by The American Society of Hematology

1
Reinhart
TA
Rogan
MJ
Viglianti
GA
et al
A new approach to investigating the relationship between productive infection and cytopathicity in vivo.
Nat Med.
3
1997
218
221
2
Haase
AT
Henry
K
Zupancic
M
et al
Quantitative image analysis of HIV-1 infection in lymphoid tissue.
Science.
274
1996
985
989
3
Zhang
Z
Schuler
T
Zupancic
M
et al
Sexual transmission and propagation of SIV and HIV in resting and activated CD4+ T cells.
Science.
286
1999
1353
1357
4
Hoffenbach
A
Langlade-Demoyen
P
Dadaglio
G
et al
Unusually high frequencies of HIV-specific cytotoxic T lymphocytes in humans.
J Immunol.
142
1989
452
462
5
Moss
PA
Rowland-Jones
SL
Frodsham
PM
et al
Persistent high frequency of human immunodeficiency virus-specific cytotoxic T cells in peripheral blood of infected donors.
Proc Natl Acad Sci U S A.
92
1995
5773
5777
6
Altman
JD
Moss
PAH
Goulder
PJR
et al
Phenotypic analysis of antigen-specific T lymphocytes.
Science.
274
1996
94
96
7
Gea-Banacloche
JC
Migueles
SA
Martino
L
et al
Maintenance of large numbers of virus-specific CD8+ T cells in HIV-infected progressors and long-term nonprogressors.
J Immunol.
165
2000
1082
1092
8
Shankar
P
Russo
M
Harnisch
B
et al
Impaired function of circulating HIV-specific CD8(+) T cells in chronic human immunodeficiency virus infection.
Blood.
96
2000
3094
3101
9
Walker
CM
Moody
DJ
Stites
DP
Levy
JA
CD8+ lymphocytes can control HIV infection in vitro by suppressing virus replication.
Science.
234
1986
1563
1566
10
Walker
BD
Chakrabarti
S
Moss
B
et al
HIV-specific cytotoxic T lymphocytes in seropositive individuals.
Nature.
328
1987
345
348
11
Plata
F
Autran
B
Martins
LP
et al
AIDS virus-specific cytotoxic T lymphocytes in lung disorders.
Nature.
328
1987
348
351
12
Cocchi
F
DeVico
AL
Garzino-Demo
A
et al
Identification of RANTES, MIP-1 alpha, and MIP-1 beta as the major HIV-suppressive factors produced by CD8+ T cells.
Science.
270
1995
1811
1815
13
Baier
M
Werner
A
Bannert
N
Metzner
K
Kurth
R
HIV suppression by interleukin-16.
Nature.
378
1995
563
14
Koup
RA
Safrit
JT
Cao
Y
et al
Temporal association of cellular immune responses with the initial control of viremia in primary human immunodeficiency virus type 1 syndrome.
J Virol.
68
1994
4650
4655
15
Borrow
P
Lewicki
H
Wei
X
et al
Antiviral pressure exerted by HIV-1-specific cytotoxic T lymphocytes (CTLs) during primary infection demonstrated by rapid selection of CTL escape virus.
Nat Med.
3
1997
205
211
16
Riviere
Y
McChesney
MB
Porrot
F
et al
Gag-specific cytotoxic responses to HIV type 1 are associated with a decreased risk of progression to AIDS-related complex or AIDS.
AIDS Res Hum Retroviruses.
11
1995
903
907
17
Kourtis
AP
Ibegbu
C
Nahmias
AJ
et al
Early progression of disease in HIV-infected infants with thymus dysfunction.
N Engl J Med.
335
1996
1431
1436
18
Schmitz
JE
Kuroda
MJ
Santra
S
et al
Control of viremia in simian immunodeficiency virus infection by CD8+ lymphocytes.
Science.
283
1999
857
860
19
Jin
X
Bauer
DE
Tuttleton
SE
et al
Dramatic rise in plasma viremia after CD8(+) T cell depletion in simian immunodeficiency virus-infected macaques.
J Exp Med.
189
1999
991
998
20
Ho
DD
Neumann
AU
Perelson
AS
et al
Rapid turnover of plasma virions and CD4 lymphocytes in HIV-1 infection.
Nature.
373
1995
123
126
21
Wei
X
Ghosh
SK
Taylor
ME
et al
Viral dynamics in human immunodeficiency type 1 infection.
Nature.
373
1995
117
122
22
Isaaz
S
Baetz
K
Olsen
K
Podack
E
Griffiths
GM
Serial killing by cytotoxic T lymphocytes: T cell receptor triggers degranulation, re-filling of the lytic granules and secretion of lytic proteins via a non-granule pathway.
Eur J Immunol.
25
1995
1071
1079
23
Lieberman
J
Cytotoxic T lymphocyte adoptive immunotherapy for HIV infection.
Cytotoxic Cells: Basic Mechanisms and Medical Applications.
Sitkovsky
MV
Henkart
PA
1999
441
457
Lippincott Williams & Wilkins
Philadelphia, PA
24
Koup
RA
Sullivan
JL
Levine
PH
et al
Detection of major histocompatibility complex class Irestricted, HIV-specific cytotoxic T lymphocytes in the blood of infected hemophiliacs.
Blood.
73
1989
1909
1914
25
Riviere
Y
Tanneau-Salvadori
F
Regnault
A
et al
Human immunodeficiency virus-specific cytotoxic responses of seropositive individuals: distinct types of effector cells mediate killing of targets expressing gag and env proteins.
J Virol.
63
1989
2270
2277
26
Trimble
LA
Lieberman
J
Circulating CD8 T lymphocytes in human immunodeficiency virus-infected individuals have impaired function and downmodulate CD3 zeta, the signaling chain of the T-cell receptor complex.
Blood.
91
1998
585
594
27
Gray
CM
Lawrence
J
Schapiro
JM
et al
Frequency of class I HLA-restricted anti-HIV CD8+ T cells in individuals receiving highly active antiretroviral therapy (HAART).
J Immunol.
162
1999
1780
1788
28
Appay
V
Nixon
DF
Donahoe
SM
et al
HIV-specific CD8+ T cells produce antiviral cytokines but are impaired in cytolytic function.
J Exp Med.
192
2000
63
75
29
Phillips
RE
Rowland-Jones
S
Nixon
DF
et al
Human immunodeficiency virus genetic variation that can escape cytotoxic T cell recognition.
Nature.
354
1991
453
459
30
Tsomides
TJ
Aldovini
A
Johnson
RP
et al
Naturally processed viral peptides recognized by cytotoxic T lymphocytes on cells chronically infected by human immunodeficiency virus type 1.
J Exp Med.
180
1994
1283
1293
31
Tortorella
D
Gewurz
BE
Furman
MH
Schust
DJ
Ploegh
HL
Viral subversion of the immune system.
Annu Rev Immunol.
18
2000
861
926
32
Jameson
SC
Carbone
FR
Bevan
MJ
Clone-specific T cell receptor antagonists of major histocompatibility complex class I-restricted cytotoxic T cells.
J Exp Med.
177
1993
1541
1550
33
Klenerman
P
Rowland-Jones
S
McAdam
S
et al
Cytotoxic T-cell activity antagonized by naturally occurring HIV-1 Gag variants.
Nature.
369
1994
403
407
34
Evavold
BD
Allen
PM
Separation of IL-4 production from Th cell proliferation by an altered T cell receptor ligand.
Science.
252
1991
1308
1310
35
Allen
TM
O'Connor
DH
Jing
P
et al
Tat-specific cytotoxic T lymphocytes select for SIV escape variants during resolution of primary viraemia.
Nature.
407
2000
386
390
36
Schwartz
O
Riviere
Y
Heard
JM
Danos
O
Reduced cell surface expression of processed human immunodeficiency virus type 1 envelope glycoprotein in the presence of Nef.
J Virol.
67
1993
3274
3280
37
Schwartz
O
Marechal
V
Le Gall
S
Lemonnier
F
Heard
JM
Endocytosis of major histocompatibility class I molecules is induced by the HIV-Nef protein.
Nat Med.
2
1996
338
342
38
Le Gall
S
Heard
JM
Schwartz
O
Analysis of Nef-induced MHC-I endocytosis.
Res Virol.
148
1997
43
47
39
Greenberg
ME
Iafrate
AJ
Skowronski
J
The SH3 domain-binding surface and an acidic motif in HIV-1 Nef regulate trafficking of class I MHC complexes.
EMBO J.
17
1998
2777
2789
40
Collins
KL
Chen
BK
Kalams
SA
Walker
BD
Baltimore
D
HIV-1 Nef protein protects infected primary cells against killing by cytotoxic T lymphocytes.
Nature.
391
1998
397
401
41
Ferrari
G
Humphrey
W
McElrath
MJ
et al
Clade B-based HIV-1 vaccines elicit cross-clade cytotoxic T lymphocyte reactivities in uninfected volunteers.
Proc Natl Acad Sci U S A.
94
1997
1396
1401
42
Shankar
P
Sprang
H
Lieberman
J
Effective lysis of HIV-1-infected primary CD4+ T cells by a cytotoxic T-lymphocyte clone directed against a novel A2-restricted reverse-transcriptase epitope.
J Acquir Immune Defic Syndr Hum Retrovirol.
19
1998
111
120
43
Shankar
P
Xu
Z
Lieberman
J
Viral-specific cytotoxic T lymphocytes lyse HIV-infected primary T lymphocytes by the granule exocytosis pathway.
Blood.
94
1999
3084
3093
44
Culmann
B
Gomard
E
Kieny
MP
et al
Six epitopes reacting with human cytotoxic CD8+ T cells in the central region of the HIV-1 NEF protein.
J Immunol.
146
1991
1560
1565
45
Sykulev
Y
Joo
M
Vturina
I
Tsomides
TJ
Eisen
HN
Evidence that a single peptide-MHC complex on a target cell can elicit a cytolytic T cell response.
Immunity.
4
1996
565
571
46
Cohen
GB
Gandhi
RT
Davis
DM
et al
The selective downregulation of class I major histocompatibility complex proteins by HIV-1 protects HIV-infected cells from NK cells.
Immunity.
10
1999
661
671
47
Maier
CC
Greene
MI
Biochemical features of anergic T cells.
Immunol Res.
17
1998
133
140
48
Clerici
M
Stocks
NI
Zajac
RA
et al
Interleukin-2 production used to detect antigenic peptide recognition by T-helper lymphocytes from asymptomatic HIV-seropositive individuals.
Nature.
339
1989
383
385
49
Schwartz
D
Sharma
U
Busch
M
et al
Absence of recoverable infectious virus and unique immune responses in an asymptomatic HIV+ long-term survivor.
AIDS Res Hum Retroviruses.
10
1994
1703
1711
50
Miedema
F
Meyaard
L
Koot
M
et al
Changing virus-host interactions in the course of HIV-1 infection.
Immunol Rev.
140
1994
35
72
51
Rosenberg
ES
Billingsley
JM
Caliendo
AM
et al
Vigorous HIV-1-specific CD4+ T cell responses associated with control of viremia.
Science.
278
1997
1447
1450
52
Valentine
FT
Paolino
A
Saito
A
Holzman
RS
Lymphocyte-proliferative responses to HIV antigens as a potential measure of immunological reconstitution in HIV disease.
AIDS Res Hum Retroviruses.
14(suppl 2)
1998
S161
S166
53
Rosenberg
ES
LaRosa
L
Flynn
T
Robbins
G
Walker
BD
Characterization of HIV-1-specific T-helper cells in acute and chronic infection.
Immunol Lett.
66
1999
89
93
54
Lori
F
Jessen
H
Lieberman
J
et al
Treatment of human immunodeficiency virus infection with hydroxyurea, didanosine, and a protease inhibitor before seroconversion is associated with normalized immune parameters and limited viral reservoir.
J Infect Dis.
180
1999
1827
1832
55
Pitcher
CJ
Quittner
C
Peterson
DM
et al
HIV-1-specific CD4+ T cells are detectable in most individuals with active HIV-1 infection, but decline with prolonged viral suppression.
Nat Med.
5
1999
518
525
56
Dybul
M
Mercier
G
Belson
M
et al
CD40 ligand trimer and IL-12 enhance peripheral blood mononuclear cells and CD4+ T cell proliferation and production of IFN-gamma in response to p24 antigen in HIV-infected individuals: potential contribution of anergy to HIV-specific unresponsiveness.
J Immunol.
165
2000
1685
1691
57
Chouaib
S
Chehimi
J
Bani
L
et al
Interleukin 12 induces the differentiation of major histocompatibility complex class I-primed cytotoxic T-lymphocyte precursors into allospecific cytotoxic effectors.
Proc Natl Acad Sci U S A.
91
1994
12659
12663
58
Carini
C
Essex
M
Interleukin 2-independent interleukin 7 activity enhances cytotoxic immune response of HIV-1-infected individuals.
AIDS Res Hum Retroviruses.
10
1994
121
130
59
Refaeli
Y
Van Parijs
L
London
CA
Tschopp
J
Abbas
AK
Biochemical mechanisms of IL-2-regulated Fas-mediated T cell apoptosis.
Immunity.
8
1998
615
623
60
Zhang
X
Sun
S
Hwang
I
Tough
DF
Sprent
J
Potent and selective stimulation of memory-phenotype CD8+ T cells in vivo by IL-15.
Immunity.
8
1998
591
599
61
Ku
CC
Murakami
M
Sakamoto
A
Kappler
J
Marrack
P
Control of homeostasis of CD8+ memory T cells by opposing cytokines.
Science.
288
2000
675
678
62
Napolitano
LA
Grant
RM
Deeks
SG
et al
Increased production of IL-7 accompanies HIV-1-mediated T-cell depletion: implications for T-cell homeostasis.
Nat Med.
7
2001
73
79
63
Bennett
SR
Carbone
FR
Karamalis
F
et al
Help for cytotoxic-T-cell responses is mediated by CD40 signaling.
Nature.
393
1998
478
480
64
Borrow
P
Tough
DF
Eto
D
et al
CD40 ligand-mediated interactions are involved in the generation of memory CD8(+) cytotoxic T lymphocytes (CTL) but are not required for the maintenance of CTL memory following virus infection.
J Virol.
72
1998
7440
7449
65
Schoenberger
SP
Toes
RE
van der Voort
EI
Offringa
R
Melief
CJ
T-cell help for cytotoxic T lymphocytes is mediated by CD40-CD40L interactions.
Nature.
393
1998
480
483
66
Cardin
RD
Brooks
JW
Sarawar
SR
Doherty
PC
Progressive loss of CD8+ T cell-mediated control of a gamma-herpesvirus in the absence of CD4+ T cells.
J Exp Med.
184
1996
863
871
67
Zajac
AJ
Blattman
JN
Murali-Krishna
K
et al
Viral immune evasion due to persistence of activated T cells without effector function.
J Exp Med.
188
1998
2205
2213
68
Miedema
F
Immunological abnormalities in the natural history of HIV infection: mechanisms and clinical relevance.
Immunodefic Rev.
3
1992
173
193
69
Lewis
DE
Tang
DS
Adu-Oppong
A
Schober
W
Rodgers
JR
Anergy and apoptosis in CD8+ T cells from HIV-infected persons.
J Immunol.
153
1994
412
420
70
Macatonia
SE
Patterson
S
Knight
SC
Suppression of immune responses by dendritic cells infected with HIV.
Immunology.
67
1989
285
289
71
Shearer
GM
HIV-induced immunopathogenesis.
Immunity.
9
1998
587
593
72
McAdam
AJ
Schweitzer
AN
Sharpe
AH
The role of B7 co-stimulation in activation and differentiation of CD4+ and CD8+ T cells.
Immunol Rev.
165
1998
231
247
73
Chen
AI
McAdam
AJ
Buhlmann
JE
et al
Ox40-ligand has a critical costimulatory role in dendritic cell: T cell interactions.
Immunity.
11
1999
689
698
74
Gramaglia
I
Cooper
D
Miner
KT
Kwon
BS
Croft
M
Co-stimulation of antigen-specific CD4 T cells by 4-1BB ligand.
Eur J Immunol.
30
2000
392
402
75
Knight
SC
Patterson
S
Bone marrow-derived dendritic cells, infection with human immunodeficiency virus, and immunopathology.
Annu Rev Immunol.
15
1997
593
615
76
Cameron
PU
Forsum
U
Teppler
H
Granelli-Piperno
A
Steinman
RM
During HIV-1 infection most blood dendritic cells are not productively infected and can induce allogeneic CD4+ T cells clonal expansion.
Clin Exp Immunol.
88
1992
226
236
77
McIlroy
D
Autran
B
Cheynier
R
et al
Infection frequency of dendritic cells and CD4+ T lymphocytes in spleens of human immunodeficiency virus-positive patients.
J Virol.
69
1995
4737
4745
78
Weissman
D
Li
Y
Orenstein
JM
Fauci
AS
Both a precursor and a mature population of dendritic cells can bind HIV. However, only the mature population that expresses CD80 can pass infection to unstimulated CD4+ T cells.
J Immunol.
155
1995
4111
4117
79
Ayehunie
S
Garcia-Zepeda
EA
Hoxie
JA
et al
Human immunodeficiency virus-1 entry into purified blood dendritic cells through CC and CXC chemokine coreceptors.
Blood.
90
1997
1379
1386
80
Geijtenbeek
TB
Kwon
DS
Torensma
R
et al
DC-SIGN, a dendritic cell-specific HIV-1-binding protein that enhances trans-infection of T cells.
Cell.
100
2000
587
597
81
Geijtenbeek
TB
Torensma
R
van Vliet
SJ
et al
Identification of DC-SIGN, a novel dendritic cell-specific ICAM-3 receptor that supports primary immune responses.
Cell.
100
2000
575
585
82
Blauvelt
A
Chougnet
C
Shearer
GM
Katz
SI
Modulation of T cell responses to recall antigens presented by Langerhans cells in HIV-discordant identical twins by anti-interleukin (IL)-10 antibodies and IL-12.
J Clin Invest.
97
1996
1550
1555
83
Fan
Z
Huang
XL
Zheng
L
et al
Cultured blood dendritic cells retain HIV-1 antigen-presenting capacity for memory CTL during progressive HIV-1 infection.
J Immunol.
159
1997
4973
4982
84
Sapp
M
Engelmayer
J
Larsson
M
et al
Dendritic cells generated from blood monocytes of HIV-1 patients are not infected and act as competent antigen presenting cells eliciting potent T-cell responses.
Immunol Lett.
66
1999
121
128
85
Canque
B
Rosenzwajg
M
Camus
S
et al
The effect of in vitro human immunodeficiency virus infection on dendritic-cell differentiation and function.
Blood.
88
1996
4215
4228
86
McIlroy
D
Autran
B
Clauvel
JP
et al
Low CD83, but normal MHC class II and costimulatory molecule expression, on spleen dendritic cells from HIV+ patients.
AIDS Res Hum Retroviruses.
14
1998
505
513
87
Chougnet
C
Cohen
SS
Kawamura
T
et al
Normal immune function of monocyte-derived dendritic cells from HIV-infected individuals: implications for immunotherapy.
J Immunol.
163
1999
1666
1673
88
Grassi
F
Hosmalin
A
McIlroy
D
et al
Depletion in blood CD11c-positive dendritic cells from HIV-infected patients.
AIDS.
13
1999
759
766
89
Clerici
M
Landay
AL
Kessler
HA
et al
Multiple patterns of alloantigen presenting/stimulating cell dysfunction in patients with AIDS.
J Immunol.
146
1991
2207
2213
90
Dudhane
A
Conti
B
Orlikowsky
T
et al
Monocytes in HIV type 1-infected individuals lose expression of costimulatory B7 molecules and acquire cytotoxic activity.
AIDS Res Hum Retroviruses.
12
1996
885
892
91
Orlikowsky
T
Wang
ZQ
Dudhane
A
et al
The cell surface marker phenotype of macrophages from HIV-1-infected subjects reflects an IL-10-enriched and IFN-gamma-deprived donor environment.
J Interferon Cytokine Res.
16
1996
957
962
92
Badley
AD
McElhinny
JA
Leibson
PJ
et al
Upregulation of Fas ligand expression by human immunodeficiency virus in human macrophages mediates apoptosis of uninfected T lymphocytes.
J Virol.
70
1996
199
206
93
Orlikowsky
T
Wang
ZQ
Dudhane
A
et al
Cytotoxic monocytes in the blood of HIV type 1infected subjects destroy targeted T cells in a CD95-dependent fashion.
AIDS Res Hum Retroviruses.
13
1997
953
960
94
Dockrell
DH
Badley
AD
Villacian
JS
et al
The expression of Fas ligand by macrophages and its upregulation by human immunodeficiency virus infection.
J Clin Invest.
101
1998
2394
2405
95
Caldwell
RL
Egan
BS
Shepherd
VL
HIV-1 Tat represses transcription from the mannose receptor promoter.
J Immunol.
165
2000
7035
7041
96
Chougnet
C
Wynn
TA
Clerici
M
et al
Molecular analysis of decreased interleukin-12 production in persons infected with human immunodeficiency virus.
J Infect Dis.
174
1996
46
53
97
Chougnet
C
Thomas
E
Landay
AL
et al
CD40 ligand and IFN-gamma synergistically restore IL-12 production in HIV-infected patients.
Eur J Immunol.
28
1998
646
656
98
Ostrowski
MA
Justement
SJ
Ehler
L
et al
The role of CD4(+) T cell help and CD40 ligand in the in vitro expansion of HIV-1-specific memory cytotoxic CD8(+) T cell responses.
J Immunol.
165
2000
6133
6141
99
Henkart
PA
Lymphocyte-mediated cytotoxicity: two pathways and multiple effector molecules.
Immunity.
1
1994
343
346
100
Kagi
D
Ledermann
B
Burki
K
Zinkernagel
RM
Hengartner
H
Lymphocyte-mediated cytotoxicity in vitro and in vivo: mechanisms and significance.
Immunol Rev.
146
1995
95
115
101
Kagi
D
Ledermann
B
Burki
K
et al
Cytotoxicity mediated by T cells and natural killer cells is greatly impaired in perforin-deficient mice.
Nature.
369
1994
31
37
102
Stepp
SE
Dufourcq-Lagelouse
R
Le Deist
F
et al
Perforin gene defects in familial hemophagocytic lymphohistiocytosis.
Science.
286
1999
1957
1959
103
Schijns
VE
Wierda
CM
van Hoeij
M
Horzinek
MC
Exacerbated viral hepatitis in IFN-gamma receptor-deficient mice is not suppressed by IL-12.
J Immunol.
157
1996
815
821
104
Geiger
KD
Nash
TC
Sawyer
S
et al
Interferon-gamma protects against herpes simplex virus type 1-mediated neuronal death.
Virology.
238
1997
189
197
105
Bot
A
Bot
S
Bona
CA
Protective role of gamma interferon during the recall response to influenza virus.
J Virol.
72
1998
6637
6645
106
Ogg
GS
Jin
X
Bonhoeffer
S
et al
Quantitation of HIV-1-specific cytotoxic T lymphocytes and plasma load of viral RNA.
Science.
279
1998
2103
2106
107
Chen
G
Shankar
P
Lange
C
et al
CD8 T cells specific for human immunodeficiency virus, Epstein-Barr virus and cytomegalovirus lack molecules for homing to lymphoid sites of infection.
Blood.
98
2001
156
164
108
Hamann
D
Baars
PA
Rep
MH
et al
Phenotypic and functional separation of memory and effector human CD8+ T cells.
J Exp Med.
186
1997
1407
1418
109
Roederer
M
De Rosa
S
Gerstein
R
et al
8 color, 10-parameter flow cytometry to elucidate complex leukocyte heterogeneity.
Cytometry.
29
1997
328
339
110
Sallusto
F
Lenig
D
Forster
R
Lipp
M
Lanzavecchia
A
Two subsets of memory T lymphocytes with distinct homing potentials and effector functions.
Nature.
401
1999
708
712
111
Trimble
LA
Shankar
P
Patterson
M
Daily
JP
Lieberman
J
Human immunodeficiency virus-specific circulating CD8 T lymphocytes have down-modulated CD3zeta and CD28, key signaling molecules for T-cell activation.
J Virol.
74
2000
7320
7330
112
Garcia-Sanz
JA
Velotti
F
MacDonald
HR
et al
Appearance of granule-associated molecules during activation of cytolytic T-lymphocyte precursors by defined stimuli.
Immunology.
64
1988
129
134
113
Andersson
J
Behbahani
H
Lieberman
J
et al
Perforin is not co-expressed with granzyme A within cytotoxic granules in CD8 T lymphocytes present in lymphoid tissue during chronic HIV infection.
AIDS.
13
1999
1295
1303
114
Andersson J, Kinloch S, Sönnerborg A, et al. Dissociation between perforin and granzyme A expression in CD8+ T-lymphocyte granules in lymphoid tissue during primary HIV-1 infection. J Infect Dis. In press.
115
Lieberman
J
Fabry
JA
Kuo
MC
et al
Cytotoxic T lymphocytes from HIV-1 seropositive individuals recognize immunodominant epitopes in Gp160 and reverse transcriptase.
J Immunol.
148
1992
2738
2747
116
Lieberman
J
Fabry
JA
Fong
DM
Parkerson
GR
3rd
Recognition of a small number of diverse epitopes dominates the cytotoxic T lymphocytes response to HIV type 1 in an infected individual.
AIDS Res Hum Retroviruses.
13
1997
383
392
117
Betts
MR
Casazza
JP
Patterson
BA
et al
Putative immunodominant human immunodeficiency virus-specific CD8(+) T-cell responses cannot be predicted by major histocompatibility complex class I haplotype.
J Virol.
74
2000
9144
9151
118
Altfeld
MA
Livingston
B
Reshamwala
N
et al
Identification of novel HLA-A2-restricted human immunodeficiency virus type 1-specific cytotoxic T-lymphocyte epitopes predicted by the HLA-A2 supertype peptide-binding motif.
J Virol.
75
2001
1301
1311
119
Sercarz
EE
Lehmann
PV
Ametani
A
et al
Dominance and crypticity of T cell antigenic determinants.
Annu Rev Immunol.
11
1993
729
766
120
Leitenberg
D
Bottomly
K
Regulation of naive T cell differentiation by varying the potency of TCR signal transduction.
Semin Immunol.
11
1999
283
292
121
Barker
E
Bossart
KN
Fujimura
SH
Levy
JA
CD28 costimulation increases CD8+ cell suppression of HIV replication.
J Immunol.
159
1997
5123
5131
122
Wagner
L
Yang
OO
Garcia-Zepeda
EA
et al
Beta-chemokines are released from HIV-1-specific cytolytic T-cell granules complexed to proteoglycans.
Nature.
391
1998
908
911
123
Patterson
BK
Behbahani
H
Kabat
WJ
et al
Leukemia inhibitory factor inhibits HIV-1 replication and is upregulated in placentae from nontransmitting women.
J Clin Invest.
107
2001
287
294
124
Geiger
KD
Nash
TC
Sawyer
S
et al
Interferon-gamma protects against herpes simplex virus type 1-mediated neuronal death.
Virology.
238
1997
189
197
125
Sarawar
SR
Cardin
RD
Brooks
JW
et al
Gamma interferon is not essential for recovery from acute infection with murine gammaherpesvirus 68.
J Virol.
71
1997
3916
3921
126
Bot
A
Bot
S
Bona
CA
Protective role of gamma interferon during the recall response to influenza virus.
J Virol.
72
1998
6637
6645
127
Ullum
H
Cozzi Lepri
A
Bendtzen
K
et al
Low production of interferon gamma is related to disease progression in HIV infection: evidence from a cohort of 347 HIV-infected individuals.
AIDS Res Hum Retroviruses.
13
1997
1039
1046
128
Bailer
RT
Holloway
A
Sun
J
et al
IL-13 and IFN-gamma secretion by activated T cells in HIV-1 infection associated with viral suppression and a lack of disease progression.
J Immunol.
162
1999
7534
7542
129
Huang
XL
Fan
Z
Kalinyak
C
Mellors
JW
Rinaldo
CR
Jr
CD8(+) T-cell gamma interferon production specific for human immunodeficiency virus type 1 (HIV-1) in HIV-1-infected subjects.
Clin Diagn Lab Immunol.
7
2000
279
287
130
Goepfert
PA
Bansal
A
Edwards
BH
et al
A significant number of human immunodeficiency virus epitope-specific cytotoxic T lymphocytes detected by tetramer binding do not produce gamma interferon.
J Virol.
74
2000
10249
10255
131
Murali-Krishna
K
Altman
JD
Suresh
M
et al
Counting antigen-specific CD8 T cells: a reevaluation of bystander activation during viral infection.
Immunity.
8
1998
177
187
132
Butz
EA
Bevan
MJ
Massive expansion of antigen-specific CD8+ T cells during an acute virus infection.
Immunity.
8
1998
167
175
133
Goulder
PJ
Tang
Y
Brander
C
et al
Functionally inert HIV-specific cytotoxic T lymphocytes do not play a major role in chronically infected adults and children.
J Exp Med.
192
2000
1819
1832
134
Vogel
TU
Allen
TM
Altman
JD
Watkins
DI
Functional impairment of simian immunodeficiency virus-specific CD8+ T cells during the chronic phase of infection.
J Virol.
75
2001
2458
2461
135
Luscher
U
Filgueira
L
Juretic
A
et al
The pattern of cytokine gene expression in freshly excised human metastatic melanoma suggests a state of reversible anergy of tumor-infiltrating lymphocytes.
Int J Cancer.
57
1994
612
619
136
Lee
PP
Yee
C
Savage
PA
et al
Characterization of circulating T cells specific for tumor-associated antigens in melanoma patients.
Nat Med.
5
1999
677
685
137
Huard
B
Karlsson
L
A subpopulation of CD8+ T cells specific for melanocyte differentiation antigens expresses killer inhibitory receptors (KIR) in healthy donors: evidence for a role of KIR in the control of peripheral tolerance.
Eur J Immunol.
30
2000
1665
1675
138
Weiss
A
Littman
DR
Signal transduction by lymphocyte antigen receptors.
Cell.
76
1994
263
274
139
Dustin
ML
Chan
AC
Signaling takes shape in the immune system.
Cell.
103
2000
283
294
140
van der Merwe
AP
Davis
SJ
Shaw
AS
Dustin
ML
Cytoskeletal polarization and redistribution of cell-surface molecules during T cell antigen recognition.
Semin Immunol.
12
2000
5
21
141
Rozdzial
MM
Malissen
B
Finkel
TH
Tyrosine-phosphorylated T cell receptor zeta chain associates with the actin cytoskeleton upon activation of mature T lymphocytes.
Immunity.
3
1995
623
633
142
Caplan
S
Zeliger
S
Wang
L
Baniyash
M
Cell-surface-expressed T-cell antigen-receptor zeta chain is associated with the cytoskeleton.
Proc Natl Acad Sci U S A.
92
1995
4768
4772
143
Dutton
RW
Bradley
LM
Swain
SL
T cell memory.
Annu Rev Immunol.
16
1998
201
223
144
Geertsma
MF
van Wengen-Stevenhagen
A
van Dam
EM
et al
Decreased expression of zeta molecules by T lymphocytes is correlated with disease progression in human immunodeficiency virus-infected persons.
J Infect Dis.
180
1999
649
658
145
Mizoguchi
H
O'Shea
JJ
Longo
DL
et al
Alterations in signal transduction molecules in T lymphocytes from tumor-bearing mice.
Science.
258
1992
1795
1798
146
Finke
JH
Zea
AH
Stanley
J
et al
Loss of T-cell receptor zeta chain and p56lck in T-cells infiltrating human renal cell carcinoma.
Cancer Res.
53
1993
5613
5616
147
Massaia
M
Attisano
C
Beggiato
E
Bianchi
A
Pileri
A
Correlation between disease activity and T-cell CD3 zeta chain expression in a B-cell lymphoma.
Br J Haematol.
88
1994
886
888
148
Renner
C
Ohnesorge
S
Held
G
et al
T cells from patients with Hodgkin's disease have a defective T-cell receptor zeta chain expression that is reversible by T-cell stimulation with CD3 and CD28.
Blood.
88
1996
236
241
149
Stefanova
I
Saville
MW
Peters
C
et al
HIV infection-induced posttranslational modification of T cell signaling molecules associated with disease progression.
J Clin Invest.
98
1996
1290
1297
150
Trimble
LA
Kam
LW
Friedman
RS
Xu
Z
Lieberman
J
CD3zeta and CD28 down-modulation on CD8 T cells during viral infection.
Blood.
96
2000
1021
1029
151
Trimble
LA
Down-modulation of CD3zeta, the signaling chain of the T cell receptor complex, is associated with CD8 T cell dysfunction in HIV infection [dissertation].
1998
Sackler School of Graduate Biomedical Sciences, Tufts University
Boston, MA
152
Helms
T
Boehm
BO
Asaad
RJ
et al
Direct visualization of cytokine-producing recall antigen-specific CD4 memory T cells in healthy individuals and HIV patients.
J Immunol.
164
2000
3723
3732
153
Waldrop
SL
Davis
KA
Maino
VC
Picker
LJ
Normal human CD4+ memory T cells display broad heterogeneity in their activation threshold for cytokine synthesis.
J Immunol.
161
1998
5284
5295
154
Kovacs
JA
Vogel
S
Albert
JM
et al
Controlled trial of interleukin-2 infusions in patients infected with the human immunodeficiency virus.
N Engl J Med.
335
1996
1350
1356
155
Spiegel
HM
Ogg
GS
DeFalcon
E
et al
Human immunodeficiency virus type 1- and cytomegalovirus-specific cytotoxic T lymphocytes can persist at high frequency for prolonged periods in the absence of circulating peripheral CD4(+) T cells.
J Virol.
74
2000
1018
1022
156
Kalams
SA
Walker
BD
The critical need for CD4 help in maintaining effective cytotoxic T lymphocyte responses.
J Exp Med.
188
1998
2199
2204
157
Rosenberg
ES
Altfeld
M
Poon
SH
et al
Immune control of HIV-1 after early treatment of acute infection.
Nature.
407
2000
523
526
158
Altfeld
M
Rosenberg
ES
The role of CD4(+) T helper cells in the cytotoxic T lymphocyte response to HIV-1.
Curr Opin Immunol.
12
2000
375
380
159
De Maria
A
Moretta
L
HLA-class I-specific inhibitory receptors in HIV-1 infection.
Hum Immunol.
61
2000
74
81
160
Mingari
MC
Ponte
M
Vitale
C
et al
Inhibitory receptors for HLA class I molecules on cytolytic T lymphocytes: functional relevance and implications for anti-tumor immune responses.
Int J Clin Lab Res.
27
1997
87
94
161
Vales-Gomez
M
Reyburn
H
Strominger
J
Molecular analyses of the interactions between human NK receptors and their HLA ligands.
Hum Immunol.
61
2000
28
38
162
Mingari
MC
Ponte
M
Vitale
C
Bellomo
R
Moretta
L
Expression of HLA class I-specific inhibitory receptors in human cytolytic T lymphocytes: a regulated mechanism that controls T-cell activation and function.
Hum Immunol.
61
2000
44
50
163
Galiani
MD
Aguado
E
Tarazona
R
et al
Expression of killer inhibitory receptors on cytotoxic cells from HIV-1-infected individuals.
Clin Exp Immunol.
115
1999
472
476
164
De Maria
A
Mavilio
D
Costa
P
et al
Multiple HLA-class I-specific inhibitory NK receptor expression and IL-4/IL-5 production by CD8+ T-cell clones in HIV-1 infection.
Immunol Lett.
72
2000
179
182
165
De Maria
A
Ferraris
A
Guastella
M
et al
Expression of HLA class I-specific inhibitory natural killer cell receptors in HIV-specific cytolytic T lymphocytes: impairment of specific cytolytic functions.
Proc Natl Acad Sci U S A.
94
1997
10285
10288
166
Springer
TA
Traffic signals for lymphocyte recirculation and leukocyte emigration: the multistep paradigm.
Cell.
76
1994
301
314
167
Nakano
H
Mori
S
Yonekawa
H
et al
A novel mutant gene involved in T-lymphocyte-specific homing into peripheral lymphoid organs on mouse chromosome 4.
Blood.
91
1998
2886
2895
168
Forster
R
Schubel
A
Breitfeld
D
et al
CCR7 coordinates the primary immune response by establishing functional microenvironments in secondary lymphoid organs.
Cell.
99
1999
23
33
169
Stein
JV
Rot
A
Luo
Y
et al
The CC chemokine thymus-derived chemotactic agent 4 (TCA-4, secondary lymphoid tissue chemokine, 6Ckine, exodus-2) triggers lymphocyte function-associated antigen 1-mediated arrest of rolling T lymphocytes in peripheral lymph node high endothelial venules.
J Exp Med.
191
2000
61
76
170
Campbell
JJ
Bowman
EP
Murphy
K
et al
6-C-kine (SLC), a lymphocyte adhesion-triggering chemokine expressed by high endothelium, is an agonist for the MIP-3beta receptor CCR7.
J Cell Biol.
141
1998
1053
1059
171
Tangemann
K
Gunn
MD
Giblin
P
Rosen
SD
A high endothelial cell-derived chemokine induces rapid, efficient, and subset-selective arrest of rolling T lymphocytes on a reconstituted endothelial substrate.
J Immunol.
161
1998
6330
6337
172
Mackay
CR
Marston
WL
Dudler
L
Naive and memory T cells show distinct pathways of lymphocyte recirculation.
J Exp Med.
171
1990
801
817
173
Vitetta
ES
Berton
MT
Burger
C
et al
Memory B and T cells.
Annu Rev Immunol.
9
1991
193
217
174
Gray
D
Immunological memory.
Annu Rev Immunol.
11
1993
49
77
175
Swain
SL
Croft
M
Dubey
C
et al
From naive to memory T cells.
Immunol Rev.
150
1996
143
167
176
Kuroda
MJ
Schmitz
JE
Charini
WA
et al
Emergence of CTL coincides with clearance of virus during primary simian immunodeficiency virus infection in rhesus monkeys.
J Immunol.
162
1999
5127
5133
177
Meylan
PR
Burgisser
P
Weyrich-Suter
C
Spertini
F
Viral load and immunophenotype of cells obtained from lymph nodes by fine needle aspiration as compared with peripheral blood cells in HIV-infected patients.
J Acquir Immune Defic Syndr Hum Retrovirol.
13
1996
39
47
178
Landay
AL
Bethel
J
Schnittman
S
Phenotypic variability of lymphocyte populations in peripheral blood and lymph nodes from HIV-infected individuals and the impact of antiretroviral therapy. DATRI 003 Study Group. Division of AIDS Treatment Research Initiative.
AIDS Res Hum Retroviruses.
14
1998
445
451
179
Carbonari
M
Cibati
M
Pesce
AM
et al
Comparison of the Vbeta repertoire in peripheral blood and in lymph nodes of HIV-infected subjects reveals skewed usage predominantly in CD8+ T cells.
Clin Immunol Immunopathol.
81
1996
200
209
180
Pantaleo
G
Soudeyns
H
Demarest
JF
et al
Accumulation of human immunodeficiency virusspecific cytotoxic T lymphocytes away from the predominant site of virus replication during primary infection.
Eur J Immunol.
27
1997
3166
3173
181
Brodie
SJ
Lewinsohn
DA
Patterson
BK
et al
In vivo migration and function of transferred HIV-1-specific cytotoxic T cells.
Nat Med.
5
1999
34
41
182
Kalams
SA
Buchbinder
SP
Rosenberg
ES
et al
Association between virus-specific cytotoxic T-lymphocyte and helper responses in human immunodeficiency virus type 1 infection.
J Virol.
73
1999
6715
6720

Author notes

Judy Lieberman, The Center for Blood Research, 800 Huntington Ave, Boston, MA 02115; e-mail:lieberman@cbr.med.harvard.edu.

Sign in via your Institution