首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
During untreated human immunodeficiency virus type 1 (HIV-1) infection, virus-specific CD8+ T cells partially control HIV replication in peripheral lymphoid tissues, but host mechanisms of HIV control in the central nervous system (CNS) are incompletely understood. We characterized HIV-specific CD8+ T cells in cerebrospinal fluid (CSF) and peripheral blood among seven HIV-positive antiretroviral therapy-naïve subjects. All had grossly normal brain magnetic resonance imaging and spectroscopy and normal neuropsychometric testing. Frequencies of epitope-specific CD8+ T cells by direct tetramer staining were on average 2.4-fold higher in CSF than in blood (P = 0.0004), while HIV RNA concentrations were lower. Cells from CSF were readily expanded ex vivo and responded to a broader range of HIV-specific human leukocyte antigen class I restricted optimal peptides than did expanded cells from blood. HIV-specific CD8+ T cells, in contrast to total CD8+ T cells, in CSF and blood were at comparable maturation states, as assessed by CD45RO and CCR7 staining. The strong relationship between higher T-cell frequencies and lower levels of viral antigen in CSF could be the result of increased migration to and/or preferential expansion of HIV-specific T cells within the CNS. This suggests an important role for HIV-specific CD8+ T cells in control of intrathecal viral replication.Human immunodeficiency virus type 1 (HIV-1) invades the central nervous system (CNS) early during primary infection (21, 30, 35), and proviral DNA persists in the brain throughout the course of HIV-1 disease (7, 25, 29, 47, 77, 83). Limited data from human and nonhuman primate studies suggest that little or no viral replication occurs in the brain during chronic, asymptomatic infection, based on the absence of demonstrable viral RNA or proteins (8, 85). In contrast, cognitive impairment affects approximately 40% of patients who progress to advanced AIDS without highly active antiretroviral therapy (21, 30, 35, 65). During HIV-associated dementia, there is active HIV-1 replication in the brain (23, 52, 61, 81), and viral sequence differences between cerebrospinal fluid (CSF) and peripheral tissues suggest distinct anatomic compartments of replication (18, 19, 22, 53, 75, 76, 78). Host mechanisms that control viral replication in the CNS during chronic, asymptomatic HIV-1 infection are incompletely understood.Anti-HIV CD8+ T cells are present in blood and peripheral tissues throughout the course of chronic HIV-1 infection (2, 14). Multiple lines of evidence support a critical role for these cells in controlling HIV-1 replication. During acute HIV-1 infection, the appearance of CD8+ T-cell responses correlates temporally with a decline in viremia (11, 43), and a greater proliferative capacity of peripheral blood HIV-specific CD8+ T cells correlates with better control of viremia (36, 54). In addition, the presence of certain major histocompatibility complex class I human leukocyte antigen (HLA) alleles, notably HLA-B*57, predicts slower progression to AIDS and death during chronic, untreated HIV-1 infection (55, 62). Finally, in the simian immunodeficiency virus (SIV) model, macaques depleted of CD8+ T cells experience increased viremia and rapid disease progression (39, 51, 67).Little is known regarding the role of intrathecal anti-HIV CD8+ T cells in HIV neuropathogenesis. Nonhuman primate studies have identified SIV-specific CD8+ T cells in the CNS early after infection (16, 80). Increased infiltration of SIV antigen-specific CD8+ T cells and cytotoxic T lymphocytes has been detected only in CSF of slow progressors without neurological symptoms (72). In chronically infected macaques with little or no SIV replication in the brain, the frequency of HIV-specific T cells was higher in CSF than in peripheral blood but did not correlate with the level of plasma viremia or CD4+ T-cell counts (56). Although intrathecal anti-HIV CD8+ T cells may help control viral replication, a detrimental role in the neuropathogenesis of HIV-1 has also been postulated (38). Immune responses contribute to neuropathogenesis in models of other infectious diseases, and during other viral infections cytotoxic T lymphocytes can worsen disease through direct cytotoxicity or release of inflammatory cytokines such as gamma interferon (IFN-γ) (3, 17, 31, 37, 42, 44, 71).We tested the hypothesis that quantitative and/or qualitative differences in HIV-specific CD8+ T-cell responses are present in CSF compared to blood during chronic, untreated HIV-1 infection. We characterized HIV-specific CD8+ T-cell responses in CSF among seven antiretroviral therapy-naïve adults with chronic HIV-1 infection, relatively high peripheral blood CD4+ T-cell counts, and low plasma HIV-1 RNA concentrations. We show that among these HIV-positive individuals with no neurological symptoms and with little or no HIV-1 RNA in CSF, frequencies of HIV-specific T cells are significantly higher in CSF than in blood. These CSF cells are at a state of differentiation similar to that of T cells in blood and are functionally competent for expansion and IFN-γ production. The higher frequency of functional HIV-specific CD8+ T cells in CSF, in the context of low or undetectable virus in CSF, suggests that these cells play a role in the control of intrathecal viral replication.  相似文献   

2.
Cytotoxicity and proliferation capacity are key functions of antiviral CD8 T cells. In the present study, we investigated a series of markers to define these functions in virus-specific CD8 T cells. We provide evidence that there is a lack of coexpression of perforin and CD127 in human CD8 T cells. CD127 expression on virus-specific CD8 T cells correlated positively with proliferation capacity and negatively with perforin expression and cytotoxicity. Influenza virus-, cytomegalovirus-, and Epstein-Barr virus/human immunodeficiency virus type 1-specific CD8 T cells were predominantly composed of CD127+ perforin/CD127 perforin+, and CD127/perforin CD8 T cells, respectively. CD127/perforin and CD127/perforin+ cells expressed significantly more PD-1 and CD57, respectively. Consistently, intracellular cytokine (gamma interferon, tumor necrosis factor alpha, and interleukin-2 [IL-2]) responses combined to perforin detection confirmed that virus-specific CD8 T cells were mostly composed of either perforin+/IL-2 or perforin/IL-2+ cells. In addition, perforin expression and IL-2 secretion were negatively correlated in virus-specific CD8 T cells (P < 0.01). As previously shown for perforin, changes in antigen exposure modulated also CD127 expression. Based on the above results, proliferating (CD127+/IL-2-secreting) and cytotoxic (perforin+) CD8 T cells were contained within phenotypically distinct T-cell populations at different stages of activation or differentiation and showed different levels of exhaustion and senescence. Furthermore, the composition of proliferating and cytotoxic CD8 T cells for a given antiviral CD8 T-cell population appeared to be influenced by antigen exposure. These results advance our understanding of the relationship between cytotoxicity, proliferation capacity, the levels of senescence and exhaustion, and antigen exposure of antiviral memory CD8 T cells.Cytotoxic CD8 T cells are a fundamental component of the immune response against viral infections and mediate an important role in immunosurveillance (7, 10, 55), and the induction of vigorous CD8 T-cell responses after vaccination is thought to be a key component of protective immunity (37, 41, 49, 50, 58, 60, 69). Cytotoxic CD8 T cells exert their antiviral and antitumor activity primarily through the secretion of cytotoxic granules containing perforin (pore-forming protein) and several granule-associated proteases, including granzymes (Grms) (5, 15, 20, 44). Several studies have recently advanced the characterization of the mechanism of granule-dependent cytotoxic activity and performed a comprehensive investigation of the content of cytotoxic granules in human virus-specific CD8 T cells (2, 19, 29, 44, 53).Heterogeneous profiles of cytotoxic granules have been identified in different virus-specific memory CD8 T cells and associated with distinct differentiation stages of memory CD8 T cells (2, 19, 29, 44). Furthermore, we have observed a hierarchy among the cytotoxic granules in setting the efficiency of cytotoxic activity and demonstrated that perforin (and to a lesser extent GrmB) but not GrmA or GrmK were associated with cytotoxic activity (29). Recently, a novel mechanism of perforin-dependent granule-independent CTL cytotoxicity has also been demonstrated (45).Major advances in the characterization of antigen (Ag)-specific CD4 and CD8 T cells have been made recently and have aimed at identifying functional profiles that may correlate with protective CD8 T-cell responses (1, 3, 4, 12, 13, 24, 28, 36-38, 40, 41, 49, 50, 56-58, 60, 64, 68). In particular, the functional characterization of antigen-specific T cells was mainly performed on the basis of (i) the pattern of cytokines secreted (i.e., gamma interferon [IFN-γ], tumor necrosis factor alpha [TNF-α], interleukin-2 [IL-2], or macrophage inflammatory protein 1β [MIP-1β]), (ii) the proliferation capacity, and (iii) the cytotoxic capacity (13, 28, 59). Of note, degranulation activity (i.e., CD107a mobilization following specific stimulation) has been used as a surrogate marker of cytotoxic activity (11, 13).The term “polyfunctional” has been used to define T-cell immune responses that, in addition to typical effector functions such as secretion of IFN-γ, TNF-α, or MIP-1β and cytotoxic activity (measured by the degranulation capacity), comprise distinct T-cell populations able to secrete IL-2 and retain proliferation capacity (13, 28, 49, 50). Some evidence indicates that a hallmark of protective immune responses is the presence of polyfunctional T-cell responses (59). Furthermore, the ability to secrete IL-2 was shown to be linked to proliferation capacity, and both factors have been associated with protective antiviral immunity (13, 28, 49, 50). Although a lack of correlation between degranulation activity and GrmB expression was reported in mice (65), the relationship between degranulation activity and perforin expression has never been comprehensively investigated in mice and in humans.The private α chain of the IL-7 receptor (IL-7Rα, also called CD127) has been suggested to selectively identify CD8 T cells that will become long-lived memory cells (6, 34, 36). Moreover, it was shown in mice (34, 36) and humans (14, 48, 63) that the CD127high memory-precursor CD8 T cells produced IL-2 in contrast to CD127low effector CD8 T cells. Of interest, CD127 expression has also been shown to correlate with Ag-specific proliferation capacity in mice (34, 36). A similar correlation was observed in humans, although only for polyclonal stimulations (48). With the exception of studies performed in HIV-1 infection, where an association between CD127 expression and HIV-1 viremia has been shown (21, 22, 42, 48, 54), very limited information is available on the CD127 expression in human virus-specific CD8 T cells other that HIV-1.Although cytotoxic activity and proliferation capacity are key components of the antiviral cellular immune response, the relationship between these functions has been only investigated in nonprogressive HIV-1 infection (46), where these two functions were shown to be related. However, it still remains to be determined whether these functions are mediated by the same or by different T-cell populations.In the present study, we performed a comprehensive characterization of virus-specific CD8 T-cell responses against HIV-1, cytomegalovirus (CMV), Epstein Barr virus (EBV), and influenza virus (Flu) in order to (i) analyze the degree of concordance between degranulation activity and perforin/Grm expression; (ii) identify the relevance of CD127 in identifying virus-specific CD8 T cells endowed with proliferation capacity; (iii) delineate the relationship between proliferation capacity, cytotoxic activity, activation/differentiation stage, and level of exhaustion of CD8 T cells; and (iv) determine the influence of antigen exposure in shaping the functional composition of virus-specific CD8 T cells.Our data indicate that cytotoxic (as defined by perforin expression) and proliferating (as defined by CD127 expression or IL-2 secretion) virus-specific CD8 T cells are contained within distinct CD8 T-cell populations. Furthermore, the proportion of proliferating and cytotoxic T cells within a given virus-specific CD8 T-cell population appears to be influenced by antigen exposure. These results advance our understanding of the relationship between cytotoxicity, proliferative capacity, differentiation stage, and Ag exposure of memory CD8 T cells.  相似文献   

3.
Direct cell-to-cell spread of human immunodeficiency virus type 1 (HIV-1) between T cells at the virological synapse (VS) is an efficient mechanism of viral dissemination. Tetherin (BST-2/CD317) is an interferon-induced, antiretroviral restriction factor that inhibits nascent cell-free particle release. The HIV-1 Vpu protein antagonizes tetherin activity; however, whether tetherin also restricts cell-cell spread is unclear. We performed quantitative cell-to-cell transfer analysis of wild-type (WT) or Vpu-defective HIV-1 in Jurkat and primary CD4+ T cells, both of which express endogenous levels of tetherin. We found that Vpu-defective HIV-1 appeared to disseminate more efficiently by cell-to-cell contact between Jurkat cells under conditions where tetherin restricted cell-free virion release. In T cells infected with Vpu-defective HIV-1, tetherin was enriched at the VS, and VS formation was increased compared to the WT, correlating with an accumulation of virus envelope proteins on the cell surface. Increasing tetherin expression with type I interferon had only minor effects on cell-to-cell transmission. Furthermore, small interfering RNA (siRNA)-mediated depletion of tetherin decreased VS formation and cell-to-cell transmission of both Vpu-defective and WT HIV-1. Taken together, these data demonstrate that tetherin does not restrict VS-mediated T cell-to-T cell transfer of Vpu-defective HIV-1 and suggest that under some circumstances tetherin might promote cell-to-cell transfer, either by mediating the accumulation of virions on the cell surface or by regulating integrity of the VS. If so, inhibition of tetherin activity by Vpu may balance requirements for efficient cell-free virion production and cell-to-cell transfer of HIV-1 in the face of antiviral immune responses.Human immunodeficiency virus type 1 can disseminate between and within hosts by cell-free infection or by direct cell-cell spread. Cell-cell spread of HIV-1 between CD4+ T cells is an efficient means of viral dissemination (65) and has been estimated to be several orders of magnitude more rapid than cell-free virus infection (6, 8, 41, 64, 74). Cell-cell transmission of HIV-1 takes place at the virological synapse (VS), a multimolecular structure that forms at the interface between an HIV-1-infected T cell and an uninfected target T cell during intercellular contact (27). Related structures that facilitate cell-cell spread of HIV-1 between dendritic cells and T cells (42) and between macrophages and T cells (16, 17) and for cell-cell spread of the related retrovirus human T-cell leukemia virus type 1 (HTLV-1) (24) have also been described. Moreover, more long-range cell-cell transfer can occur via cellular projections, including filopodia (71) and membrane nanotubes (75). The VS is initiated by binding of the HIV-1 envelope glycoprotein (Env), which is expressed on the surfaces of infected T cells, to HIV-1 entry receptors (CD4 and either CXCR4 or CCR5) present on the target cell membrane (6, 22, 27, 41, 61, 73). Interactions between LFA-1 and ICAM-1 and ICAM-3 further stabilize the conjugate interface and, together with Env receptor binding, help trigger the recruitment of viral proteins, CD4/coreceptor, and integrins to the contact site (27, 28, 61). The enrichment of viral and cellular proteins at the VS is an active process, dependent on cytoskeletal remodeling, and in the infected T cell both the actin and tubulin network regulate polarization of HIV-1 proteins at the cell-cell interface, thus directing HIV-1 assembly and egress toward the engaged target cell (27, 29). Virus is transferred by budding into the synaptic cleft, and virions subsequently attach to the target cell membrane to mediate entry, either by fusion at the plasma membrane or possibly following endocytic uptake (2, 22). In this way, the VS promotes more rapid infection kinetics and may enhance HIV-1 pathogenesis in vivo.Cells have evolved a number of barriers to resist invading microorganisms. One mechanism that appears to be particularly important in counteracting HIV-1 infection is a group of interferon-inducible, innate restriction factors that includes TRIM5α, APOBEC3G, and tetherin (38, 49, 69, 79). Tetherin (BST-2/CD317) is a host protein expressed by many cell types, including CD4+ T cells, that acts at a late stage of the HIV-1 life cycle to trap (or “tether”) mature virions at the plasma membranes of virus-producing cells, thereby inhibiting cell-free virus release (49, 56, 81). This antiviral activity of tetherin is not restricted to HIV-1, and tetherin can also inhibit the release of other enveloped viruses from infected cells (31, 40, 54, 62). What the cellular function of tetherin is besides its antiviral activity is unclear, but because expression is upregulated following alpha/beta interferon (IFN-α/β) treatment (1) and tetherin can restrict a range of enveloped viruses, tetherin has been postulated to be a broad-acting mediator of the innate immune defense against enveloped viruses.To circumvent restriction of particle release, HIV-1 encodes the 16-kDa accessory protein Vpu, which antagonizes tetherin and restores normal virus budding (47, 78). The molecular mechanisms by which Vpu does this are not entirely clear, but evidence suggests that Vpu may exert its antagonistic function by downregulating tetherin from the cell surface, trapping it in the trans-Golgi network (10) and targeting it for degradation by the proteasome (12, 39, 81) or lysosome (9, 25, 44); however, degradation of tetherin may be dispensable for Vpu activity (13), and in HIV-1-infected T cells, surface downregulation of tetherin has been reported to be minor (45), suggesting that global removal of tetherin from the plasma membrane may not be necessary to antagonize its function.Tetherin-mediated restriction of HIV-1 and antagonism by Vpu have been the focus of much research, and inhibition of cell-free virus infection has been well documented (33, 47-49, 77, 81, 82). In contrast, less studied is the impact of tetherin on direct cell-cell dissemination. For example, it is not clear if tetherin-mediated restriction inhibits T cell-T cell spread as efficiently as cell-free release or whether tetherin affects VS formation. To address these questions, we analyzed Vpu+ and Vpu viruses for their ability to spread directly between Jurkat T cells and primary CD4+ T cells in the presence or absence of endogenous tetherin. Our data suggest that tetherin does not restrict HIV-1 in the context of cell-to-cell transmission of virus between T cells expressing endogenous tetherin. Interestingly, we also that observed that Vpu-defective virus may disseminate more efficiently by cell-cell spread at the VS. We postulate that cell-cell spread may favor viral pathogenesis by allowing HIV-1 to disseminate in the presence of tetherin during an interferon-producing innate response.  相似文献   

4.
5.
6.
Human immunodeficiency virus (HIV) envelope (Env)-mediated bystander apoptosis is known to cause the progressive, severe, and irreversible loss of CD4+ T cells in HIV-1-infected patients. Env-induced bystander apoptosis has been shown to be gp41 dependent and related to the membrane hemifusion between envelope-expressing cells and target cells. Caveolin-1 (Cav-1), the scaffold protein of specific membrane lipid rafts called caveolae, has been reported to interact with gp41. However, the underlying pathological or physiological meaning of this robust interaction remains unclear. In this report, we examine the interaction of cellular Cav-1 and HIV gp41 within the lipid rafts and show that Cav-1 modulates Env-induced bystander apoptosis through interactions with gp41 in SupT1 cells and CD4+ T lymphocytes isolated from human peripheral blood. Cav-1 significantly suppressed Env-induced membrane hemifusion and caspase-3 activation and augmented Hsp70 upregulation. Moreover, a peptide containing the Cav-1 scaffold domain sequence markedly inhibited bystander apoptosis and apoptotic signal pathways. Our studies shed new light on the potential role of Cav-1 in limiting HIV pathogenesis and the development of a novel therapeutic strategy in treating HIV-1-infected patients.HIV infection causes a progressive, severe, and irreversible depletion of CD4+ T cells, which is responsible for the development of AIDS (9). The mechanism through which HIV infection induces cell death involves a variety of processes (58). Among these processes, apoptosis is most likely responsible for T-cell destruction in HIV-infected patients (33), because active antiretroviral therapy has been associated with low levels of CD4+ T-cell apoptosis (7), and AIDS progression was shown previously to correlate with the extent of immune cell apoptosis (34). Importantly, bystander apoptosis of uninfected cells was demonstrated to be one of the major processes involved in the destruction of immune cells (58), with the majority of apoptotic CD4+ T cells in the peripheral blood and lymph nodes being uninfected in HIV patients (22).Binding to uninfected cells or the entry of viral proteins released by infected cells is responsible for the virus-mediated killing of innocent-bystander CD4+ T cells (2-4, 9, 65). The HIV envelope glycoprotein complex, consisting of gp120 and gp41 subunits expressed on an HIV-infected cell membrane (73), is believed to induce bystander CD4+ T-cell apoptosis (58). Although there is a soluble form of gp120 in the blood, there is no conclusive agreement as to whether the concentration is sufficient to trigger apoptosis (57, 58). The initial step in HIV infection is mediated by the Env glycoprotein gp120 binding with high affinity to CD4, the primary receptor on the target cell surface, which is followed by interactions with the chemokine receptor CCR5 or CXCR4 (61). This interaction triggers a conformational change in gp41 and the insertion of its N-terminal fusion peptide into the target membrane (30). Next, a prehairpin structure containing leucine zipper-like motifs is formed by the two conserved coiled-coil domains, called the N-terminal and C-terminal heptad repeats (28, 66, 70). This structure quickly collapses into a highly stable six-helix bundle structure with an N-terminal heptad repeat inside and a hydrophobic C-terminal heptad repeat outside (28, 66, 70). The formation of the six-helix bundle leads to a juxtaposition and fusion with the target cell membrane (28, 66, 70). The fusogenic potential of HIV Env is proven to correlate with the pathogenesis of both CXCR4- and CCR5-tropic viruses by not only delivering the viral genome to uninfected cells but also mediating Env-induced bystander apoptosis (71). Initial infection is dominated by the CCR5-tropic strains, with the CXCR4-tropic viruses emerging in the later stages of disease (20). Studies have shown that CXCR4-tropic HIV-1 triggers more depletion of CD4+ T cells than CCR5-tropic strains (36).Glycolipid- and cholesterol-enriched membrane microdomains, termed lipid rafts, are spatially organized plasma membranes and are known to have many diverse functions (26, 53). These functions include membrane trafficking, endocytosis, the regulation of cholesterol and calcium homeostasis, and signal transduction in cellular growth and apoptosis. Lipid rafts have also been implicated in HIV cell entry and budding processes (19, 46, 48, 51). One such organelle is the caveola, which is a small, flask-shaped (50 to 100 nm in diameter) invagination in the plasma membrane (5, 62). The caveola structure, which is composed of proteins known as caveolins, plays a role in various functions by serving as a mobile platform for many receptors and signal proteins (5, 62). Caveolin-1 (Cav-1) is a 22- to 24-kDa major coat protein responsible for caveola assembly (25, 47). This scaffolding protein forms a hairpin-like structure and exists as an oligomeric complex of 14 to 16 monomers (21). Cav-1 has been shown to be expressed by a variety of cell types, mostly endothelial cells, type I pneumocytes, fibroblasts, and adipocytes (5, 62). In addition, Cav-1 expression is evident in immune cells such as macrophages and dendritic cells (38, 39). However, Cav-1 is not expressed in isolated thymocytes (49). Furthermore, Cav-1 and caveolar structures are absent in human or murine T-cell lines (27, 41, 68). Contrary to this, there has been one report showing evidence of Cav-1 expression in bovine primary cell subpopulations of CD4+, CD8+, CD21+, and IgM+ cells with Cav-1 localized predominantly in the perinuclear region (38). That report also demonstrated a membrane region staining with Cav-1-specific antibody of human CD21+ and CD26+ peripheral blood lymphocytes (PBLs). Recently, the expression of Cav-1 in activated murine B cells, with a potential role in the development of a thymus-independent immune response, was also reported (56). It remains to be determined whether Cav-1 expression is dependent on the activation state of lymphocytes. For macrophages, however, which are one of the main cell targets for HIV infection, Cav-1 expression has been clearly documented (38).The scaffolding domain of Cav-1, located in the juxtamembranous region of the N terminus, is responsible for its oligomerization and binding to various proteins (5, 62, 64). It recognizes a consensus binding motif, ΦXΦXXXXΦ, ΦXXXXΦXXΦ, or ΦXΦXXXXΦXXΦ, where Φ indicates an aromatic residue (F, W, or Y) and X indicates any residue (5, 62, 64). A Cav-1 binding motif (WNNMTWMQW) has been identified in the HIV-1 envelope protein gp41 (42, 43). Cav-1 has been shown to associate with gp41 by many different groups under various circumstances, including the immunoprecipitation of gp41 and Cav-1 in HIV-infected cells (42, 43, 52). However, the underlying pathological or physiological functions of this robust interaction between Cav-1 and gp41 remain unclear.Here, we report that the interaction between Cav-1 and gp41 leads to a modification of gp41 function, which subsequently regulates Env-induced T-cell bystander apoptosis. Moreover, we show that a peptide containing the Cav-1 scaffold domain sequence is capable of modulating Env-induced bystander apoptosis, which suggests a novel therapeutic application for HIV-1-infected patients.  相似文献   

7.
8.
CD4 T cells have been shown to play an important role in the immunity and immunopathogenesis of respiratory syncytial virus (RSV) infection. We identified two novel CD4 T-cell epitopes in the RSV M and M2 proteins with core sequences M213-223 (FKYIKPQSQFI) and M227-37 (YFEWPPHALLV). Peptides containing the epitopes stimulated RSV-specific CD4 T cells to produce gamma interferon (IFN-γ), interleukin 2 (IL-2), and other Th1- and Th2-type cytokines in an I-Ab-restricted pattern. Construction of fluorochrome-conjugated peptide-I-Ab class II tetramers revealed RSV M- and M2-specific CD4 T-cell responses in RSV-infected mice in a hierarchical pattern. Peptide-activated CD4 T cells from lungs were more activated and differentiated, and had greater IFN-γ expression, than CD4 T cells from the spleen, which, in contrast, produced greater levels of IL-2. In addition, M209-223 peptide-activated CD4 T cells reduced IFN-γ and IL-2 production in M- and M2-specific CD8 T-cell responses to Db-M187-195 and Kd-M282-90 peptides more than M225-39 peptide-stimulated CD4 T cells. This correlated with the fact that I-Ab-M209-223 tetramer-positive cells responding to primary RSV infection had a much higher frequency of FoxP3 expression than I-Ab-M226-39 tetramer-positive CD4 T cells, suggesting that the M-specific CD4 T-cell response has greater regulatory function. Characterization of epitope-specific CD4 T cells by novel fluorochrome-conjugated peptide-I-Ab tetramers allows detailed analysis of their roles in RSV pathogenesis and immunity.CD4 T lymphocytes play an important role in the resolution of primary viral infections and the prevention of reinfection by regulating a variety of humoral and cellular immune responses. CD4 T cells provide cytokines and other molecules to support the differentiation and expansion of antigen-specific CD8 T cells, which are major effectors for both virus clearance and immunopathology during primary infection with respiratory syncytial virus (RSV) (3, 17, 42, 43). CD4 T-cell help is mandatory for an effective B-cell response (14), which is necessary for producing neutralizing antibodies that prevent secondary RSV infection (12, 18, 21). A concurrent CD4 T-cell response also promotes the maintenance of CD8 T-cell surveillance and effector capacity (9). Previous studies have shown that interleukin 2 (IL-2) from CD4 T cells can restore CD8 T-cell function in lungs (10) and that IL-2 supplementation can increase the production of gamma interferon (IFN-γ) by CD8 T cells upon peptide stimulation in vitro (45).While CD4 T cells are important for providing support to host immunity, they have also been associated with immunopathogenesis by playing a key role in the Th2-biased T-cell response (34, 46), which may be the common mechanism of enhanced lung pathology and other disease syndromes shown in murine studies (2, 16, 17, 19, 35). Earlier studies showed the positive association of formalin-inactivated RSV (FI-RSV) immunization-mediated enhanced illness upon subsequent natural RSV infection with a Th2-biased CD4 T-cell response (19, 44). Th2-orientated CD4 T cells elicit severe pneumonia with extensive eosinophilic infiltrates in the lungs of FI-RSV-immunized mice (13, 24, 48). Patients with severe RSV disease showed an elevated Th2/Th1 cytokine ratio in nasal secretions and peripheral blood mononuclear cells (27, 29, 31, 38). Increased disease severity has also been associated with polymorphisms in Th2-related cytokine genes, such as the IL-4, IL-4 receptor, and IL-13 genes (11, 23, 36). Th2 cytokines from CD4 T cells can also diminish the CD8 T-cell response and delay viral clearance (4, 8).The evaluation of CD4 T-cell responses in viral infection is particularly relevant in the RSV model because of the association of RSV and allergic inflammation, which is largely mediated by CD4 T cells. Understanding the influence of CD4 T cells on CD8 T-cell responses and other immunological effector mechanisms is central to understanding RSV pathogenesis and developing preventive vaccine strategies for RSV. Our lab and others have demonstrated that CD8 T cells target RSV M and M2 proteins with cytolytic effector activities (28, 30, 39). In this study, we found that both RSV M and M2 proteins also contain CD4 T-cell epitopes. These epitopes have 11-mer amino acid core sequences and are associated with the major histocompatibility complex (MHC) class II molecule I-Ab. Fluorochrome-conjugated peptide-I-Ab molecule tetrameric complexes can identify RSV M- and M2-specific CD4 T cells from CB6F1 mice following RSV infection in a hierarchical pattern. Peptides containing the epitopes can stimulate CD4 T cells from RSV M or M2 DNA-immunized and virus-challenged mice and can lead to the production of IFN-γ, IL-2, and other Th1- and Th2-type cytokines that can modulate the CD8 T-cell response to RSV M and M2. We also found that CD4 T cells from the lungs and spleens of immunized mice have different phenotype and cytokine profiles upon in vitro stimulation. These observations suggest a regulatory role for CD4 T cells in the host response to RSV infection. The development of novel MHC class II tetramer reagents allows the characterization of epitope-specific CD4 T-cell responses to RSV and will enable the investigation of basic mechanisms by which CD4 T cells affect pathogenesis and immunity to viral infections.  相似文献   

9.
10.
11.
The immune correlates of human/simian immunodeficiency virus control remain elusive. While CD8+ T lymphocytes likely play a major role in reducing peak viremia and maintaining viral control in the chronic phase, the relative antiviral efficacy of individual virus-specific effector populations is unknown. Conventional assays measure cytokine secretion of virus-specific CD8+ T cells after cognate peptide recognition. Cytokine secretion, however, does not always directly translate into antiviral efficacy. Recently developed suppression assays assess the efficiency of virus-specific CD8+ T cells to control viral replication, but these assays often use cell lines or clones. We therefore designed a novel virus production assay to test the ability of freshly ex vivo-sorted simian immunodeficiency virus (SIV)-specific CD8+ T cells to suppress viral replication from SIVmac239-infected CD4+ T cells. Using this assay, we established an antiviral hierarchy when we compared CD8+ T cells specific for 12 different epitopes. Antiviral efficacy was unrelated to the disease status of each animal, the protein from which the tested epitopes were derived, or the major histocompatibility complex (MHC) class I restriction of the tested epitopes. Additionally, there was no correlation with the ability to suppress viral replication and epitope avidity, epitope affinity, CD8+ T-cell cytokine multifunctionality, the percentage of central and effector memory cell populations, or the expression of PD-1. The ability of virus-specific CD8+ T cells to suppress viral replication therefore cannot be determined using conventional assays. Our results suggest that a single definitive correlate of immune control may not exist; rather, a successful CD8+ T-cell response may be comprised of several factors.CD8+ T cells may play a critical role in blunting peak viremia and controlling human immunodeficiency virus (HIV) and simian immunodeficiency virus (SIV) replication. The transient depletion of CD8+ cells in SIV-infected macaques results in increased viral replication (26, 31, 51, 70). The emergence of virus-specific CD8+ T cells coincides with the reduction of peak viremia (12, 39, 42, 63), and CD8+ T-cell pressure selects for escape mutants (6, 9, 13, 28, 29, 38, 60, 61, 85). Furthermore, particular major histocompatibility complex (MHC) class I alleles are overrepresented in SIV- and HIV-infected elite controllers (15, 29, 33, 34, 46, 56, 88).Because it has been difficult to induce broadly neutralizing antibodies (Abs), the AIDS vaccine field is currently focused on developing a vaccine designed to elicit HIV-specific CD8+ T cells (8, 52, 53, 82). Investigators have tried to define the immune correlates of HIV control. Neither the magnitude nor the breadth of epitopes recognized by virus-specific CD8+ T-cell responses correlates with the control of viral replication (1). The quality of the immune response may, however, contribute to the antiviral efficacy of the effector cells. It has been suggested that the number of cytokines that virus-specific CD8+ T cells secrete may correlate with viral control, since HIV-infected nonprogressors appear to maintain CD8+ T cells that secrete several cytokines, compared to HIV-infected progressors (11, 27). An increased amount of perforin secretion may also be related to the proliferation of HIV-specific CD8+ T cells in HIV-infected nonprogressors (55). While those studies offer insight into the different immune systems of progressors and nonprogressors, they did not address the mechanism of viral control. Previously, we found no association between the ability of SIV-specific CD8+ T-cell clones to suppress viral replication in vitro and their ability to secrete gamma interferon (IFN-γ), tumor necrosis factor alpha (TNF-α), or interleukin-2 (IL-2) (18).Evidence suggests that some HIV/SIV proteins may be better vaccine targets than others. CD8+ T cells recognize epitopes derived from Gag as early as 2 h postinfection, whereas CD8+ T cells specific for epitopes in Env recognize infected cells only at 18 h postinfection (68). Additionally, a previously reported study of HIV-infected individuals showed that an increased breadth of Gag-specific responses was associated with lower viral loads (35, 59, 65, 66). CD8+ T-cell responses specific for Env, Rev, Tat, Vif, Vpr, Vpu, and Nef were associated with higher viral loads, with increased breadth of Env in particular being significantly associated with a higher chronic-phase viral set point.None of the many sophisticated methods employed for analyzing the characteristics of HIV- or SIV-specific immune responses clearly demarcate the critical qualities of an effective antiviral response. In an attempt to address these questions, we developed a new assay to measure the antiviral efficacy of individual SIV-specific CD8+ T-cell responses sorted directly from fresh peripheral blood mononuclear cells (PBMC). Using MHC class I tetramers specific for the epitope of interest, we sorted freshly isolated virus-specific CD8+ T cells and determined their ability to suppress virus production from SIV-infected CD4+ T cells. We then looked for a common characteristic of efficacious epitope-specific CD8+ T cells using traditional methods.  相似文献   

12.
13.
Highly active antiretroviral therapy (HAART) can reduce human immunodeficiency virus type 1 (HIV-1) viremia to clinically undetectable levels. Despite this dramatic reduction, some virus is present in the blood. In addition, a long-lived latent reservoir for HIV-1 exists in resting memory CD4+ T cells. This reservoir is believed to be a source of the residual viremia and is the focus of eradication efforts. Here, we use two measures of population structure—analysis of molecular variance and the Slatkin-Maddison test—to demonstrate that the residual viremia is genetically distinct from proviruses in resting CD4+ T cells but that proviruses in resting and activated CD4+ T cells belong to a single population. Residual viremia is genetically distinct from proviruses in activated CD4+ T cells, monocytes, and unfractionated peripheral blood mononuclear cells. The finding that some of the residual viremia in patients on HAART stems from an unidentified cellular source other than CD4+ T cells has implications for eradication efforts.Successful treatment of human immunodeficiency virus type 1 (HIV-1) infection with highly active antiretroviral therapy (HAART) reduces free virus in the blood to levels undetectable by the most sensitive clinical assays (18, 36). However, HIV-1 persists as a latent provirus in resting, memory CD4+ T lymphocytes (6, 9, 12, 16, 48) and perhaps in other cell types (45, 52). The latent reservoir in resting CD4+ T cells represents a barrier to eradication because of its long half-life (15, 37, 40-42) and because specifically targeting and purging this reservoir is inherently difficult (8, 25, 27).In addition to the latent reservoir in resting CD4+ T cells, patients on HAART also have a low amount of free virus in the plasma, typically at levels below the limit of detection of current clinical assays (13, 19, 35, 37). Because free virus has a short half-life (20, 47), residual viremia is indicative of active virus production. The continued presence of free virus in the plasma of patients on HAART indicates either ongoing replication (10, 13, 17, 19), release of virus after reactivation of latently infected CD4+ T cells (22, 24, 31, 50), release from other cellular reservoirs (7, 45, 52), or some combination of these mechanisms. Finding the cellular source of residual viremia is important because it will identify the cells that are still capable of producing virus in patients on HAART, cells that must be targeted in any eradication effort.Detailed analysis of this residual viremia has been hindered by technical challenges involved in working with very low concentrations of virus (13, 19, 35). Recently, new insights into the nature of residual viremia have been obtained through intensive patient sampling and enhanced ultrasensitive sequencing methods (1). In a subset of patients, most of the residual viremia consisted of a small number of viral clones (1, 46) produced by a cell type severely underrepresented in the peripheral circulation (1). These unique viral clones, termed predominant plasma clones (PPCs), persist unchanged for extended periods of time (1). The persistence of PPCs indicates that in some patients there may be another major cellular source of residual viremia (1). However, PPCs were observed in a small group of patients who started HAART with very low CD4 counts, and it has been unclear whether the PPC phenomenon extends beyond this group of patients. More importantly, it has been unclear whether the residual viremia generally consists of distinct virus populations produced by different cell types.Since the HIV-1 infection in most patients is initially established by a single viral clone (23, 51), with subsequent diversification (29), the presence of genetically distinct populations of virus in a single individual can reflect entry of viruses into compartments where replication occurs with limited subsequent intercompartmental mixing (32). Sophisticated genetic tests can detect such population structure in a sample of viral sequences (4, 39, 49). Using two complementary tests of population structure (14, 43), we analyzed viral sequences from multiple sources within individual patients in order to determine whether a source other than circulating resting CD4+ T cells contributes to residual viremia and viral persistence. Our results have important clinical implications for understanding HIV-1 persistence and treatment failure and for improving eradication strategies, which are currently focusing only on the latent CD4+ T-cell reservoir.  相似文献   

14.
CD4C/HIVnef transgenic (Tg) mice express Nef in CD4+ T cells and in the cells of the macrophage/monocyte/dendritic lineage, and they develop an AIDS-like disease similar to human AIDS. In these mice, Nef is constitutively expressed throughout life. To rule out the contribution of any developmental defects caused by early expression of Nef, we generated inducible human immunodeficiency virus type 1 (HIV-1) Nef Tg mice by using the tetracycline-inducible system. Faithful expression of the Nef transgene was induced in (CD4C/rtTA × TRE/HIVNef) or (CD4C/rtTA2S-M2 × TRE/HIVNef) double-Tg mice upon doxycycline (DOX) treatment in drinking water. Long-term treatment of these mice with DOX also led to loss, apoptosis, and activation of CD4+ T cells, this latter phenotype being observed even with low levels of Nef. These phenotypes could be transferred by bone marrow (BM) transplantation, indicating a hematopoietic cell autonomous effect. In addition, in mixed Tg:non-Tg BM chimeras, only Tg and not non-Tg CD4+ T cells exhibited an effector/memory phenotype in the absence of lymphopenia. Finally, the DOX-induced double-Tg mice developed nonlymphoid organ diseases similar to those of CD4C/HIVNef Tg mice and of humans infected with HIV-1. These results show for the first time that adult mice are susceptible to the detrimental action of Nef and that Nef-mediated T-cell activation can be independent of lymphopenia. These Tg mice represent a unique model which is likely to be instrumental for understanding the cellular and molecular pathways of Nef action as well as the main characteristics of immune reconstitution following DOX withdrawal.Small animal models able to express the entire human immunodeficiency virus (HIV) genome or selected HIV genes have provided useful information on the pathogenesis of AIDS and still represent important research tools toward this goal. Among these models, transgenic (Tg) mice containing intact copies of HIV DNA, defective provirus with the gag and pol genes deleted, or individual HIV-1 genes have been reported to develop various pathologies, some of which resemble those found in human AIDS (2, 3, 8, 9, 16, 17, 18, 24, 27, 29, 30, 38, 44, 45, 46, 49, 51, 52). The cell type context in which the HIV-1 transgene is expressed in these Tg mice appears to play an important role in determining the type of pathological lesions. Tg mice generated in our laboratory and expressing the entire coding sequence of HIV-1 (CD4C/HIVWT) or HIV-1 Nef alone (CD4C/HIVNef) in the relevant target cells of HIV-1, namely, CD4+ T cells, macrophages, and dendritic cells, develop pathologies very similar to those in human AIDS (17, 18). The AIDS-like disease of CD4C/HIVNef Tg mice is characterized by immunodeficiency, loss of CD4+ T cells, thymic atrophy, activation of T cells and pathologies in heart, lungs, and kidneys (18, 53). Similarly, expression of simian immunodeficiency virus (SIV) Nef in Tg mice under the control of the same promoter sequences (CD4C) results in an AIDS-like disease (42). These studies demonstrated that Nef plays an important role in the development of the AIDS-like disease induced by HIV-1 or SIV in Tg mice.Among the AIDS-like phenotypes of these models, the T-cell activation observed by a number of groups in Tg mice expressing Nef (3, 33, 44, 53) may be of special interest for its resemblance to that of humans or macaques infected with HIV-1 or SIV, respectively. HIV infection results in a state of chronic immune activation which correlates very closely with disease progression in humans (11, 14, 23). Similarly, SIV-infected macaques which develop AIDS show aberrant immune activation (35), while SIV-infected sooty mangabey monkeys, natural hosts of SIV, do not develop immunopathologies and do not show immune activation either (41). Various factors may contribute to this immune activation, including increased plasma lipopolysaccharide levels due to microbial translocation from the gut (4), impaired regulatory T cell function (32), or the action of the HIV-1 gene products themselves, such as Env gp120 and Nef (10, 12, 43). Consistent with this latter scenario, we reported that in CD4C/HIVNef Tg mice the extent of T-cell activation correlates with levels of Nef expression in CD4+ T cells, thus suggesting a direct involvement of Nef in this activation (53). In contrast, Koenen and coworkers reported that T-cell activation in CD2/Nef Tg mice is induced indirectly by lymphophenia (26). In that study, chimeric mice, which were generated from a mixture of non-Tg and Nef Tg bone marrow (BM) cells, were not lymphopenic, and the donor-derived Nef-expressing Tg T cells did not show an activated phenotype. However, the donor Nef Tg T cells constituted only 1 to 2% of peripheral T cells of these chimeric mice (26). Clearly, alternative experimental approaches are needed to study this phenotype in a more physiological context.In the previously described CD4C/HIVNef Tg mice (18), Nef expression begins early in life and is constitutively expressed throughout the life of the animal. The AIDS-like disease caused by this early expression of Nef best represents a model for pediatric AIDS. However, in these Tg mice, Nef may interfere with normal developmental processes and these latter defects may contribute to some of the phenotypes observed. To assess the effects of Nef in fully mature adult animals, and thus develop a model of adult AIDS, temporal regulation of Nef expression in adult mice using an inducible system is required.In the present study, we chose the tet-On (rtTA and rtTA2S-M2) system (13, 15, 25, 48) to induce expression of HIV-1 Nef in CD4+ T cells and cells of the macrophage/dendritic lineage of mice using the CD4C tissue-specific regulatory elements. These CD4C sequences were previously used to generate the constitutively Nef-expressing CD4C/HIVNef Tg mice (18). These inducible adult (TRE/HIVNef × CD4C/rtTA) and (TRE/HIVNef × CD4C/rtTA2S-M2) double-Tg (DTg) mice express Nef when treated with doxycycline (DOX) and develop an AIDS-like disease very similar to that seen in constitutively Nef-expressing CD4C/HIVNef Tg mice. We took advantage of this novel biological system to reassess the role of Nef in T-cell activation. Using a mixed chimera made with BM cells from these inducible Nef Tg mice and from non-Tg mice, we could document CD4+ T-cell activation only in donor-derived Nef-expressing Tg cells, but not in non-Tg cells, in the absence of lymphopenia. This result strongly suggests that this CD4+ T-cell activation phenotype is most likely driven by expression of Nef in these cells.  相似文献   

15.
16.
We previously showed that agonistic antibodies to CD40 could substitute for CD4 T-cell help and prevent reactivation of murine gammaherpesvirus 68 (MHV-68) in the lungs of major histocompatibility complex (MHC) class II−/− (CII−/−) mice, which are CD4 T cell deficient. Although CD8 T cells were required for this effect, no change in their activity was detected in vitro. A key question was whether anti-CD40 treatment (or CD4 T-cell help) changed the function of CD8 T cells or another cell type in vivo. To address this question, in the present study, we showed that adoptive transfer of CD8 T cells from virus-infected wild-type mice or anti-CD40-treated CII−/− mice caused a significant reduction in lung viral titers, in contrast to those from control CII−/− mice. Anti-CD40 treatment also greatly prolonged survival of infected CII−/− mice. This confirms that costimulatory signals cause a change in CD8 T cells enabling them to maintain effective long-term control of MHV-68. We investigated the nature of this change and found that expression of the inhibitory receptor PD-1 was significantly increased on CD8 T cells in the lungs of MHV-68-infected CII−/−, CD40−/−, or CD80/86−/− mice, compared with that in wild-type or CD28/CTLA4−/− mice, correlating with the level of viral reactivation. Furthermore, blocking PD-1-PD-L1 interactions significantly reduced viral reactivation in CD4 T-cell-deficient mice. In contrast, the absence of another inhibitory receptor, NKG2A, had no effect. These data suggest that CD4 T-cell help programs a change in CD8 T-cell function mediated by altered PD-1 expression, which enables effective long-term control of MHV-68.Murine gammaherpesvirus 68 (MHV-68) is a naturally occurring rodent pathogen which is closely related to Epstein-Barr virus (EBV) and Kaposi''s sarcoma-associated herpesvirus (KSHV) (17, 64). Intranasal administration of MHV-68 to mice results in acute productive infection of lung epithelial cells and a latent infection in various cell types, including B lymphocytes, dendritic cells, epithelial cells, and macrophages (18, 19, 52, 53, 61, 65). The virus induces an inflammatory infiltrate in the lungs, lymph node enlargement, splenomegaly, and mononucleosis comprising increased numbers of activated CD8 T cells in the blood (53, 58). It has also been reported to induce lymphoproliferative disease/lymphoma in immunocompromised mice (30, 55, 60). Thus, the pathogenesis resembles that of EBV in humans, although structurally, the virus is more closely related to KSHV.Infectious MHV-68 is cleared from the lungs by a T-cell-dependent mechanism 10 to 15 days after infection (18, 53, 56). In wild-type mice, the lungs remain clear of replicating virus thereafter. Although CD4 T cells are not essential for primary clearance of replicating virus, they are required for effective long-term control (11). Thus, major histocompatibility complex (MHC) class II−/− mice that lack CD4 T cells or mice rendered CD4 deficient by antibody treatment initially clear infectious virus from the lungs. However, infectious virus reactivates in the lungs 10 to 15 days later and gradually increases in titer (11, 43). The infected CD4-deficient mice eventually die, apparently from long-term lung damage due to continuing lytic viral replication (11). MHC class II−/− mice do not produce antibody to T-dependent antigens (10). Cytotoxic T-lymphocyte (CTL) epitopes have been identified in open reading frame (ORF) 6 (p56, H-2Db-restricted), and ORF 61 (p79, H-2Kb-restricted) gene products, which appear to encode early lytic-phase proteins (32, 49). The epitopes are presented during two distinct phases during MHV-68 infection, which changes the pattern of CTL dominance (32, 51). However, there is no significant difference in the numbers of CD8 T cells specific for each epitope in wild-type mice and CD4 T-cell-deficient mice (4, 50). In addition, CTL activity measured in vitro does not differ substantially in the lungs of wild-type mice or CD4 T-cell-deficient mice (4, 11, 50). Furthermore, postexposure vaccination with the p56 epitope failed to prevent viral reactivation in class II−/− mice, despite dramatically expanding the number of CD8 T cells specific for the peptide (5). In contrast, vaccination of wild-type mice against these epitopes reduced lytic viral titers in the lung dramatically on subsequent challenge with MHV-68. B-cell-deficient mice clear MHV-68 with the kinetics of wild-type mice and do not show viral reactivation in the lungs (13, 61), suggesting that antibody is not essential for control of the virus. Depletion of CD4 T cells during the latent phase of infection in B-cell-deficient mice does not induce viral reactivation, whereas depletion of both CD4 and CD8 T-cell subsets provokes viral reactivation in the lungs (52). Short-term depletion of both CD4 and CD8 T-cell subsets during the latent phase of infection in wild-type mice does not lead to viral reactivation probably due to the presence of neutralizing antibody (11). Taken together, these results suggest that CD4 and CD8 T cells and B cells play overlapping roles in preventing or controlling reactivation of MHV-68 during the latent phase of infection. However, the B-cell- and CD8 T-cell-mediated control mechanisms do not develop in the absence of CD4 T cells.We, and others, have previously shown that the costimulatory molecule CD28 is not required for long-term control of MHV-68 (28, 29). However, interestingly, mice lacking both of the ligands for CD28, CD80 and CD86, show viral reactivation in the lung (21, 35). Our previously published data showed that agonistic antibodies to CD40 could substitute for CD4 T-cell function in the long-term control of MHV-68 (46). CD8 T-cell receptor-positive (TCR+) cells were required for this effect, while antibody production was not restored (45, 46). MHV-68-infected CD40L−/− mice (7) and CD40−/− mice (29) also showed viral reactivation in the lungs. However, no change in CD8 CTL activity was detected in in vitro assays following anti-CD40 treatment (46). A key question was whether anti-CD40 treatment (or CD4 T-cell help) caused a direct change in CD8 T-cell function or whether both CD8 T cells and an independent anti-CD40-sensitive step were required for viral control. To address this question, we used adoptive transfer of CD8 T cells from MHV-68-infected wild-type mice, anti-CD40-treated mice, or control MHC class II−/− mice to MHV-68-infected class II−/− recipients. We also investigated whether anti-CD40 treatment prolonged survival in addition to reducing lung viral titers. The heterodimeric molecule CD94/NKG2A has been implicated in negatively regulating the CD8 T-cell response to polyomavirus (38) and herpes simplex virus (HSV) (54), while the inhibitory receptor PD-1 (programmed death 1) has been implicated in T-cell exhaustion following infection with several other persistent viruses (2, 15, 20, 22, 26, 36, 39-41, 57, 67). In the present study, we investigated the effect of signaling via various costimulatory molecules on the expression of NKG2A and PD-1 and how these molecules influenced viral control.  相似文献   

17.
We previously reported that CD4C/human immunodeficiency virus (HIV)Nef transgenic (Tg) mice, expressing Nef in CD4+ T cells and cells of the macrophage/dendritic cell (DC) lineage, develop a severe AIDS-like disease, characterized by depletion of CD4+ T cells, as well as lung, heart, and kidney diseases. In order to determine the contribution of distinct populations of hematopoietic cells to the development of this AIDS-like disease, five additional Tg strains expressing Nef through restricted cell-specific regulatory elements were generated. These Tg strains express Nef in CD4+ T cells, DCs, and macrophages (CD4E/HIVNef); in CD4+ T cells and DCs (mCD4/HIVNef and CD4F/HIVNef); in macrophages and DCs (CD68/HIVNef); or mainly in DCs (CD11c/HIVNef). None of these Tg strains developed significant lung and kidney diseases, suggesting the existence of as-yet-unidentified Nef-expressing cell subset(s) that are responsible for inducing organ disease in CD4C/HIVNef Tg mice. Mice from all five strains developed persistent oral carriage of Candida albicans, suggesting an impaired immune function. Only strains expressing Nef in CD4+ T cells showed CD4+ T-cell depletion, activation, and apoptosis. These results demonstrate that expression of Nef in CD4+ T cells is the primary determinant of their depletion. Therefore, the pattern of Nef expression in specific cell population(s) largely determines the nature of the resulting pathological changes.The major cell targets and reservoirs for human immunodeficiency virus type 1 (HIV-1)/simian immunodeficiency virus (SIV) infection in vivo are CD4+ T lymphocytes and antigen-presenting cells (macrophages and dendritic cells [DC]) (21, 24, 51). The cell specificity of these viruses is largely dependent on the expression of CD4 and of its coreceptors, CCR5 and CXCR-4, at the cell surface (29, 66). Infection of these immune cells leads to the severe disease, AIDS, showing widespread manifestations, including progressive immunodeficiency, immune activation, CD4+ T-cell depletion, wasting, dementia, nephropathy, heart and lung diseases, and susceptibility to opportunistic pathogens, such as Candida albicans (1, 27, 31, 37, 41, 82, 93, 109). It is reasonable to assume that the various pathological changes in AIDS result from the expression of one or many HIV-1/SIV proteins in these immune target cells. However, assigning the contribution of each infected cell subset to each phenotype has been remarkably difficult, despite evidence that AIDS T-cell phenotypes can present very differently depending on the strains of infecting HIV-1 or SIV or on the cells targeted by the virus (4, 39, 49, 52, 72). For example, the T-cell-tropic X4 HIV strains have long been associated with late events and severe CD4+ T-cell depletion (22, 85, 96). However, there are a number of target cell subsets expressing CD4 and CXCR-4, and identifying which one is responsible for this enhanced virulence has not been achieved in vivo. Similarly, the replication of SIV in specific regions of the thymus (cortical versus medullary areas), has been associated with very different outcomes but, unfortunately, the critical target cells of the viruses were not identified either in these studies (60, 80). The task is even more complex, because HIV-1 or SIV can infect several cell subsets within a single cell population. In the thymus, double (CD4 CD8)-negative (DN) or triple (CD3 CD4 CD8)-negative (TN) T cells, as well as double-positive (CD4+ CD8+) (DP) T cells, are infectible by HIV-1 in vitro (9, 28, 74, 84, 98, 99, 110) and in SCID-hu mice (2, 5, 91, 94). In peripheral organs, gut memory CCR5+ CD4+ T cells are primarily infected with R5 SIV, SHIV, or HIV, while circulating CD4+ T cells can be infected by X4 viruses (13, 42, 49, 69, 70, 100, 101, 104). Moreover, some detrimental effects on CD4+ T cells have been postulated to originate from HIV-1/SIV gene expression in bystander cells, such as macrophages or DC, suggesting that other infected target cells may contribute to the loss of CD4+ T cells (6, 7, 32, 36, 64, 90).Similarly, the infected cell population(s) required and sufficient to induce the organ diseases associated with HIV-1/SIV expression (brain, heart, and kidney) have not yet all been identified. For lung or kidney disease, HIV-specific cytotoxic CD8+ T cells (1, 75) or infected podocytes (50, 95), respectively, have been implicated. Activated macrophages have been postulated to play an important role in heart disease (108) and in AIDS dementia (35), although other target cells could be infected by macrophage-tropic viruses and may contribute significantly to the decrease of central nervous system functions (11, 86, 97), as previously pointed out (25).Therefore, because of the widespread nature of HIV-1 infection and the difficulty in extrapolating tropism of HIV-1/SIV in vitro to their cell targeting in vivo (8, 10, 71), alternative approaches are needed to establish the contribution of individual infected cell populations to the multiorgan phenotypes observed in AIDS. To this end, we developed a transgenic (Tg) mouse model of AIDS using a nonreplicating HIV-1 genome expressed through the regulatory sequences of the human CD4 gene (CD4C), in the same murine cells as those targeted by HIV-1 in humans, namely, in immature and mature CD4+ T cells, as well as in cells of the macrophage/DC lineages (47, 48, 77; unpublished data). These CD4C/HIV Tg mice develop a multitude of pathologies closely mimicking those of AIDS patients. These include a gradual destruction of the immune system, characterized among other things by thymic and lymphoid organ atrophy, depletion of mature and immature CD4+ T lymphocytes, activation of CD4+ and CD8+ T cells, susceptibility to mucosal candidiasis, HIV-associated nephropathy, and pulmonary and cardiac complications (26, 43, 44, 57, 76, 77, 79, 106). We demonstrated that Nef is the major determinant of the HIV-1 pathogenicity in CD4C/HIV Tg mice (44). The similarities of the AIDS-like phenotypes of these Tg mice to those in human AIDS strongly suggest that such a Tg mouse approach can be used to investigate the contribution of distinct HIV-1-expressing cell populations to their development.In the present study, we constructed and characterized five additional mouse Tg strains expressing Nef, through distinct regulatory elements, in cell populations more restricted than in CD4C/HIV Tg mice. The aim of this effort was to assess whether, and to what extent, the targeting of Nef in distinct immune cell populations affects disease development and progression.  相似文献   

18.
19.
For recognition of infected cells by CD8 T cells, antigenic peptides are presented at the cell surface, bound to major histocompatibility complex class I (MHC-I) molecules. Downmodulation of cell surface MHC-I molecules is regarded as a hallmark function of cytomegalovirus-encoded immunoevasins. The molecular mechanisms by which immunoevasins interfere with the MHC-I pathway suggest, however, that this downmodulation may be secondary to an interruption of turnover replenishment and that hindrance of the vesicular transport of recently generated peptide-MHC (pMHC) complexes to the cell surface is the actual function of immunoevasins. Here we have used the model of murine cytomegalovirus (mCMV) infection to provide experimental evidence for this hypothesis. To quantitate pMHC complexes at the cell surface after infection in the presence and absence of immunoevasins, we generated the recombinant viruses mCMV-SIINFEKL and mCMV-Δm06m152-SIINFEKL, respectively, expressing the Kb-presented peptide SIINFEKL with early-phase kinetics in place of an immunodominant peptide of the viral carrier protein gp36.5/m164. The data revealed ∼10,000 Kb molecules presenting SIINFEKL in the absence of immunoevasins, which is an occupancy of ∼10% of all cell surface Kb molecules, whereas immunoevasins reduced this number to almost the detection limit. To selectively evaluate their effect on preexisting pMHC complexes, cells were exogenously loaded with SIINFEKL peptide shortly after infection with mCMV-SIINFEKA, in which endogenous presentation is prevented by an L174A mutation of the C-terminal MHC-I anchor residue. The data suggest that pMHC complexes present at the cell surface in advance of immunoevasin gene expression are downmodulated due to constitutive turnover in the absence of resupply.CD8 T cells recognize infected cells by interaction of their T-cell receptor (TCR) with a cell surface presentation complex composed of a cognate antigenic peptide bound to a presenting allelic form of a major histocompatibility complex class I (MHC-I) glycoprotein (77, 85, 97, 98). The number of such “peptide receptors” per cell has been estimated to be on the order of 105 to 106 for each MHC-I allomorph (for a review, see reference 82). Viral antigenic peptides are generated within infected cells by proteolytic processing of viral proteins, usually in the proteasome, and associate with nascent MHC-I proteins in the endoplasmic reticulum (ER) before the peptide-MHC (pMHC) complexes travel to the cell surface with the cellular vesicular flow (for reviews, see references 13, 87, 92, and 93). CD8 T cells have long been known to protect against cytomegalovirus (CMV) infection and disease in animal models (60, 72; reviewed in references 33 and 36) and in humans (9, 61, 67, 75, 76). As shown only recently in the murine CMV (mCMV) model of infection of immunocompromised mice by adoptive transfer of epitope-specific CD8 T cells, antiviral protection against CMV is indeed TCR mediated and epitope dependent. Specifically, memory cells purified by TCR-based epitope-specific cell sorting, as well as cells of a peptide-selected cytolytic T-lymphocyte line, protected against mCMV expressing the cognate antigenic peptide, the IE1 peptide 168-YPHFMPTNL-176 in this example, but failed to control infection with a recombinant mCMV expressing a peptide analogue in which the C-terminal MHC-I anchor residue leucine was replaced with alanine (3).Interference with the MHC-I pathway of antigen presentation has evolved as a viral immune evasion mechanism of CMVs and other viruses, mediated by virally encoded proteins that inhibit MHC-I trafficking to the cell surface (for reviews, see references 1, 24, 27, 29, 63, 70, 71, 84, and 95). These molecules are known as immunoevasins (50, 70, 89), as “viral proteins interfering with antigen presentation” (VIPRs) (95), or as negative “viral regulators of antigen presentation” (vRAPs) (34). Although the detailed molecular mechanisms differ between different CMV species in their respective hosts, the common biological outcome is the inhibition of antigen presentation. Accordingly, downmodulation of MHC-I cell surface expression is a hallmark of molecular immune evasion and actually led to the discovery of this class of molecules. Since CD8 T cells apparently protect against infection with wild-type CMV strains despite the expression of immunoevasins, the in vivo relevance of these molecules is an issue of current interest and investigation (for a review, see reference 14). As shown recently with the murine model, antigen presentation in infected host cells is not completely blocked for all epitopes, because pMHC complexes that are constitutively formed in sufficiently large amounts can exhaust the inhibitory capacity of the immunoevasins (40). Likewise, enhancing antigen processing conditionally with gamma interferon (IFN-γ) aids in peptide presentation in the presence of immunoevasins (18, 28). Thus, by raising the threshold of the amount of peptide required for presentation, immunoevasins determine whether a particular viral peptide can function as a protective epitope—an issue of relevance for rational vaccine design as well (94). Whereas deletion of immunoevasin genes gives only incremental improvement to the control of infection in immunocompetent mice (22, 51), expression of immunoevasins reduces the protective effect of adoptively transferred CD8 T cells in immunocompromised recipients (37, 40, 47, 48). In a bone marrow transplantation model, immunoevasins were recently found to contribute to enhanced and prolonged virus replication during hematopoietic reconstitution and, consequently, also to higher latent viral genome loads in the lungs and a higher incidence of virus recurrence (4). Notably, however, immunoevasins do not inhibit but, rather, enhance CD8 T-cell priming (5, 21, 22, 56), due to higher viral replication levels in draining lymph nodes associated with sustained antigen supply for the cross-priming of CD8 T cells by uninfected antigen-presenting cells (5).For mCMV, three molecules are proposed to function as vRAPs, only two of which are confirmed negative regulators that downmodulate cell surface MHC-I (34, 62, 89) and inhibit the presentation of antigenic peptides to CD8 T cells (34, 62). Immunoevasin gp40/m152 transiently interacts with MHC-I molecules and mediates their retention in a cis-Golgi compartment (96), whereas gp48/m06 stably binds to MHC-I molecules in the ER and mediates sorting of the complexes for lysosomal degradation by a mechanism that involves the cellular cargo sorting adaptor proteins AP1-A and AP3-A (73, 74). The third proposed immunoevasin of mCMV, gp34/m04 (46), also binds stably to MHC-I molecules. A function as a CD8 T-cell immunoevasin was predicted from some alleviation of immune evasion for certain epitopes and MHC-I molecules in cells infected with the deletion mutant mCMV-Δm04 (34, 42, 89), but gp34/m04 does not reduce the steady-state level of cell surface class I molecules and does not inhibit peptide presentation when expressed selectively after infection with mCMV-Δm06m152 (34, 62). The m04-MHC-I complexes are expressed on the cell surface (46) and appear to be involved in the modulation of natural killer cell activity (45).Here we give the first report on quantitating the efficacy of immunoevasins in terms of absolute numbers of pMHC complexes displayed at the cell surface. By comparing the fate of pMHC complexes already present at the cell surface in advance of immunoevasin gene expression with that of newly formed pMHC complexes, our data provide direct evidence to conclude that downmodulation of cell surface MHC-I molecules is secondary to an interruption of the flow of newly formed pMHC complexes to the cell surface.(Part of this work was presented at the 12th International CMV/Betaherpesvirus Workshop, 10 to 14 May 2009, Boston, MA.)  相似文献   

20.
设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号