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.
2.
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   

3.
4.
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   

5.
Previous studies have identified a central role for HLA-B alleles in influencing control of HIV infection. An alternative possibility is that a small number of HLA-B alleles may have a very strong impact on HIV disease outcome, dominating the contribution of other HLA alleles. Here, we find that even following the exclusion of subjects expressing any of the HLA-B class I alleles (B*57, B*58, and B*18) identified to have the strongest influence on control, the dominant impact of HLA-B alleles on virus set point and absolute CD4 count variation remains significant. However, we also find that the influence of HLA on HIV control in this C-clade-infected cohort from South Africa extends beyond HLA-B as HLA-Cw type remains a significant predictor of virus and CD4 count following exclusion of the strongest HLA-B associations. Furthermore, there is evidence of interdependent protective effects of the HLA-Cw*0401-B*8101, HLA-Cw*1203-B*3910, and HLA-A*7401-B*5703 haplotypes that cannot be explained solely by linkage to a protective HLA-B allele. Analysis of individuals expressing both protective and detrimental alleles shows that even the strongest HLA alleles appear to have an additive rather than dominant effect on HIV control at the individual level. Finally, weak but significant frequency-dependent effects in this cohort can be detected only by looking at an individual''s combined HLA allele frequencies. Taken together, these data suggest that although individual HLA alleles, particularly HLA-B, can have a strong impact, HIV control overall is likely to be influenced by the additive effect of some or all of the other HLA alleles present.HIV-specific CD8+ T cells play a central role in resolution of primary viremia and the long-term suppression of viral replication (13). Supporting this notion is the observed correlation between possession of particular human leukocyte antigen (HLA) class I alleles and control of HIV, measured both directly by time-to-AIDS (5, 6) and indirectly via clinical markers of disease progression (viral load [VL] and CD4 count) (15, 26, 28). Specific HLA class I alleles have been associated with relatively successful control of viral replication and slow disease progression, most notably, alleles HLA-B*57 and HLA-B*27 (1, 7, 12, 15, 21, 23), and also with relatively ineffective control of viral replication and rapid disease progression [B*35(Px), B*5802, and B*18] (5, 15, 17, 23). In addition, general trends suggesting an HLA class I heterozygote advantage (5) and rare allele advantage (28) and, most recently, a correlation between levels of surface expression linked to certain HLA-Cw alleles (11, 27) and HIV control has also been described.Among the different HLA class I loci, the HIV-specific CD8+ T-cell responses restricted by HLA-B alleles are thought to play the central role in determining disease outcome: the majority of detectable HIV-specific CD8+ T-cell responses are restricted by HLA-B alleles (3, 15, 16), HLA-B-restricted responses typically express a more effective “polyfunctional” phenotype (14), the strongest HLA-associations with either slow or rapid progression are with HLA-B alleles (5, 10, 11, 15), and HLA-B-restricted CD8+ T cells exert the strongest selection pressure on the virus (15, 19, 24). However, whether this apparent association between HIV immune control and HLA-B is a general and causal trend or, rather, is biased by the coincidence that the strongest HLA associations with either extreme of disease control happen, by chance, to involve HLA-B alleles remains uncertain.In order to further investigate the correlation between HLA type and HIV infection control, we here examine a cohort now comprising >1,200 chronically HIV C-clade-infected, treatment-naïve subjects from Durban, South Africa, in an extended analysis following from our previous studies of a smaller cohort (15). We first address the question of whether the dominant role of HLA-B in this population compared to the roles of HLA-A or HLA-C results from the influence of HLA-B alleles in general or is dependent on a few known strong associations, such as that between HLA-B*57 alleles and low viremia. Second, in light of recent data (11, 27), we assess the impact of HLA-C alleles on HIV disease outcome and examine the effect of HLA haplotypes on observed HLA associations with disease control. Third, we investigate the question of whether the impact of certain HLA-B alleles on HIV outcome dominates that of other HLA-B alleles to negate the contribution of the latter or whether the impact of individual HLA alleles can be additive. Finally, we compare the impact of individual HLA alleles on HIV on immune control to the impact of heterozygote and rare allele advantage in this cohort.  相似文献   

6.
7.
8.
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   

9.
The structural precursor polyprotein, Gag, encoded by all retroviruses, including the human immunodeficiency virus type 1 (HIV-1), is necessary and sufficient for the assembly and release of particles that morphologically resemble immature virus particles. Previous studies have shown that the addition of Ca2+ to cells expressing Gag enhances virus particle production. However, no specific cellular factor has been implicated as mediator of Ca2+ provision. The inositol (1,4,5)-triphosphate receptor (IP3R) gates intracellular Ca2+ stores. Following activation by binding of its ligand, IP3, it releases Ca2+ from the stores. We demonstrate here that IP3R function is required for efficient release of HIV-1 virus particles. Depletion of IP3R by small interfering RNA, sequestration of its activating ligand by expression of a mutated fragment of IP3R that binds IP3 with very high affinity, or blocking formation of the ligand by inhibiting phospholipase C-mediated hydrolysis of the precursor, phosphatidylinositol-4,5-biphosphate, inhibited Gag particle release. These disruptions, as well as interference with ligand-receptor interaction using antibody targeted to the ligand-binding site on IP3R, blocked plasma membrane accumulation of Gag. These findings identify IP3R as a new determinant in HIV-1 trafficking during Gag assembly and introduce IP3R-regulated Ca2+ signaling as a potential novel cofactor in viral particle release.Assembly of the human immunodeficiency virus (HIV) is determined by a single gene that encodes a structural polyprotein precursor, Gag (71), and may occur at the plasma membrane or within late endosomes/multivesicular bodies (LE/MVB) (7, 48, 58; reviewed in reference 9). Irrespective of where assembly occurs, the assembled particle is released from the plasma membrane of the host cell. Release of Gag as virus-like particles (VLPs) requires the C-terminal p6 region of the protein (18, 19), which contains binding sites for Alix (60, 68) and Tsg101 (17, 37, 38, 41, 67, 68). Efficient release of virus particles requires Gag interaction with Alix and Tsg101. Alix and Tsg101 normally function to sort cargo proteins to LE/MVB for lysosomal degradation (5, 15, 29, 52). Previous studies have shown that addition of ionomycin, a calcium ionophore, and CaCl2 to the culture medium of cells expressing Gag or virus enhances particle production (20, 48). This is an intriguing observation, given the well-documented positive role for Ca2+ in exocytotic events (33, 56). It is unclear which cellular factors might regulate calcium availability for the virus release process.Local and global elevations in the cytosolic Ca2+ level are achieved by ion release from intracellular stores and by influx from the extracellular milieu (reviewed in reference 3). The major intracellular Ca2+ store is the endoplasmic reticulum (ER); stores also exist in MVB and the nucleus. Ca2+ release is regulated by transmembrane channels on the Ca2+ store membrane that are formed by tetramers of inositol (1,4,5)-triphosphate receptor (IP3R) proteins (reviewed in references 39, 47, and 66). The bulk of IP3R channels mediate release of Ca2+ from the ER, the emptying of which signals Ca2+ influx (39, 51, 57, 66). The few IP3R channels on the plasma membrane have been shown to be functional as well (13). Through proteomic analysis, we identified IP3R as a cellular protein that was enriched in a previously described membrane fraction (18) which, in subsequent membrane floatation analyses, reproducibly cofractionated with Gag and was enriched in the membrane fraction only when Gag was expressed. That IP3R is a major regulator of cytosolic calcium concentration (Ca2+) is well documented (39, 47, 66). An IP3R-mediated rise in cytosolic Ca2+ requires activation of the receptor by a ligand, inositol (1,4,5)-triphosphate (IP3), which is produced when phospholipase C (PLC) hydrolyzes phosphatidylinositol-4,5-bisphosphate [PI(4,5)P2] at the plasma membrane (16, 25, 54). Paradoxically, PI(4,5)P2 binds to the matrix (MA) domain in Gag (8, 55, 59), and the interaction targets Gag to PI(4,5)P2-enriched regions on the plasma membrane; these events are required for virus release (45). We hypothesized that PI(4,5)P2 binding might serve to target Gag to plasma membrane sites of localized Ca2+ elevation resulting from PLC-mediated PI(4,5)P2 hydrolysis and IP3R activation. This idea prompted us to investigate the role of IP3R in Gag function.Here, we show that HIV-1 Gag requires steady-state levels of IP3R for its efficient release. Three isoforms of IP3R, types 1, 2, and 3, are encoded in three independent genes (39, 47). Types 1 and 3 are expressed in a variety of cells and have been studied most extensively (22, 39, 47, 73). Depletion of the major isoforms in HeLa or COS-1 cells by small interfering RNA (siRNA) inhibited viral particle release. Moreover, we show that sequestration of the IP3R activating ligand or blocking ligand formation also inhibited Gag particle release. The above perturbations, as well as interfering with receptor expression or activation, led to reduced Gag accumulation at the cell periphery. The results support the conclusion that IP3R activation is required for efficient HIV-1 viral particle release.  相似文献   

10.
11.
12.
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   

13.
Little is known about the transmission or tropism of the newly discovered human retrovirus, human T-cell lymphotropic virus type 3 (HTLV-3). Here, we examine the entry requirements of HTLV-3 using independently expressed Env proteins. We observed that HTLV-3 surface glycoprotein (SU) binds efficiently to both activated CD4+ and CD8+ T cells. This contrasts with both HTLV-1 SU, which primarily binds to activated CD4+ T cells, and HTLV-2 SU, which primarily binds to activated CD8+ T cells. Binding studies with heparan sulfate proteoglycans (HSPGs) and neuropilin-1 (NRP-1), two molecules important for HTLV-1 entry, revealed that these molecules also enhance HTLV-3 SU binding. However, unlike HTLV-1 SU, HTLV-3 SU can bind efficiently in the absence of both HSPGs and NRP-1. Studies of entry performed with HTLV-3 Env-pseudotyped viruses together with SU binding studies revealed that, for HTLV-1, glucose transporter 1 (GLUT-1) functions at a postbinding step during HTLV-3 Env-mediated entry. Further studies revealed that HTLV-3 SU binds efficiently to naïve CD4+ T cells, which do not bind either HTLV-1 or HTLV-2 SU and do not express detectable levels of HSPGs, NRP-1, and GLUT-1. These results indicate that the complex of receptor molecules used by HTLV-3 to bind to primary T lymphocytes differs from that of both HTLV-1 and HTLV-2.The primate T-cell lymphotropic virus (PTLV) group of deltaretroviruses consists of three types of human T-cell lymphotropic viruses (HTLVs) (HTLV-1, HTLV-2, HTLV-3), their closely related simian T-cell lymphotropic viruses (STLVs) (STLV-1, STLV-2, STLV-3), an HTLV (HTLV-4) for which a simian counterpart has not been yet identified, and an STLV (STLV-5) originally described as a divergent STLV-1 (5-7, 30, 35, 37, 38, 45, 51, 53). HTLV-1 and HTLV-2, which have a 70% nucleotide homology, differ in both their pathobiology and tropism (reviewed in reference 13). While HTLV-1 causes a neurological disorder (tropical spastic paraparesis/HTLV-1-associated myelopathy) and a hematological disease (adult T-cell leukemia/lymphoma) (15, 42, 55), HTLV-2 is only rarely associated with tropical spastic paraparesis/HTLV-1-associated myelopathy-like disease and is not definitively linked to any lymphoproliferative disease (12, 20). In vivo, both HTLV-1 and HTLV-2 infect T cells. Although HTLV-1 is primarily found in CD4+ T cells, other cell types in the peripheral blood of infected individuals have been found to contain HTLV-1, including CD8+ T cells, dendritic cells, and B cells (19, 29, 33, 36, 46).Binding and entry of retroviruses requires specific interactions between the Env glycoproteins on the virus and cell surface receptor complexes on target cells. For HTLV-1, three molecules have been identified as important for entry, as follows: heparan sulfate proteoglycans (HSPGs), neuropilin-1 (NRP-1), and glucose transporter 1 (GLUT-1) (16, 22, 26, 28, 29, 34, 39, 44). Recent studies support a model in which HSPG and NRP-1 function during the initial binding of HTLV-1 to target cells, and GLUT-1 functions at a postattachment stage, most likely to facilitate fusion (29, 34, 49). Efficient HTLV-2 binding and entry requires NRP-1 and GLUT-1 but not HSPGs (16, 26, 39, 49).This difference in the molecules required for binding to target cells reflects differences in the T-cell tropisms of these two viruses. Activated CD4+ T cells express much higher levels of HSPGs than CD8+ T cells (26). In infected individuals, HTLV-1 is primarily found in CD4+ T cells, while HTLV-2 is primarily found in CD8+ T cells (21, 43, 46). In vitro, HTLV-1 preferentially transforms CD4+ T cells while HTLV-2 preferentially transforms CD8+ T cells, and this difference has been mapped to the Env proteins (54).We and others have reported the discovery of HTLV-3 in two Cameroonese inhabitants (6, 7, 53). We recently uncovered the presence of a third HTLV-3 strain in a different population living several hundred kilometers away from the previously identified groups (5), suggesting that this virus may be common in central Africa. Since the HTLV-3 sequences were obtained by PCR amplification of DNA isolated from peripheral blood mononuclear cells (PBMCs) of infected individuals, little is known about its tropism and pathobiology in vivo. Based on the correlation between HSPG expression levels and viral tropisms of HTLV-1 and HTLV-2, we reasoned that knowledge about the HTLV-3 receptors might provide insight into the tropism of this virus. We therefore generated vectors expressing HTLV-3 Env proteins and used them to begin to characterize the receptor complex used by HTLV-3 to bind and enter cells.  相似文献   

14.
We describe a structural rearrangement that can occur in parvovirus minute virus of mice (MVMp) virions following prolonged exposure to buffers containing 0.5 mM EDTA. Such particles remain stable at 4°C but undergo a conformational shift upon heating to 37°C at pH 7.2 that leads to the ejection of much of the viral genome in a 3′-to-5′ direction, leaving the DNA tightly associated with the otherwise intact capsid. This rearrangement can be prevented by the addition of 1 mM CaCl2 or MgCl2 prior to incubation at 37°C, suggesting that readily accessible divalent cation binding sites in the particle are critical for genome retention. Uncoating was not seen following the incubation of virions at pH 5.5 and 37°C or at pH 7.2 and 37°C in particles with subgenomic DNA, suggesting that pressure exerted by the full-length genome may influence this process. Uncoated genomes support complementary-strand synthesis by T7 DNA polymerase, but synthesis aborts upstream of the right-hand end, which remains capsid associated. We conclude that viral genomes are positioned so that their 3′ termini and coding sequences can be released from intact particles at physiological temperatures by a limited conformational rearrangement. In the presence of divalent cations, incremental heating between 45°C and 65°C induces structural transitions that first lead to the extrusion of VP1 N termini, followed by genome exposure. However, in cation-depleted virions, the sequence of these shifts is blurred. Moreover, cation-depleted particles that have been induced to eject their genomes at 37°C continue to sequester their VP1 N termini within the intact capsid, suggesting that these two extrusion events represent separable processes.The capsids of nonenveloped viruses have evolved to protect their genomes from hostile extra- and intracellular environments and to undergo a specific program of conformational shifts in response to cellular cues, which sequentially expose trafficking effector structures required to transport the nucleic acid to the appropriate cellular location, at the correct time, for its successful replication (22, 25, 31, 42, 43, 46). In many cases, such shifts are possible because the virion is poised in a metastable configuration, prevented from collapsing to an alternate structure by the energy barrier between the two forms. Cellular or environmental triggers effectively compromise this energy barrier during cell entry, inducing the particle to rearrange. Accordingly, the pattern of particle morphogenesis is in part encrypted within the virion so that probing its structure in vitro, using elevated temperatures or biased ionic conditions, provides an avenue for assessing the types of rearrangements that the virion is capable of sustaining (5, 12, 22, 24, 29, 34, 35, 46, 47, 50). In this study we exploit in vitro manipulation to gain insight into the structure and potential rearrangements programmed into the minute virus of mice (MVM) virion.As for all members of the family Parvoviridae, virions of MVM are exceptionally small and structurally simple, comprising a single copy of a ∼5-kb negative-sense, self-priming, single-stranded, linear DNA enclosed in a nonenveloped protein capsid with an external diameter of 280 Å and T=1 icosahedral symmetry (1, 30). Capsids are assembled from 60 polypeptides encoded by a single structural gene, which are expressed in an approximately 1:5 ratio as two size variants, VP1 (83 kDa) and VP2 (63 kDa) (44). They share 587 amino acids of carboxy-terminal sequence, while VP1 has an N-terminal extension of 142 amino acids that contains a phospholipase A2 (PLA2) domain essential for endosomal escape during entry (19, 53). As can be seen in the X-ray structure, capsid shells are constructed from 549 amino acids common to the C termini of VP1 and VP2, leaving signal-rich N-terminal extensions of 38 residues for VP2 and 180 residues for VP1, whose disposition within the assembled virion remains uncertain (1). The core structure comprises the same eight-stranded, antiparallel, β-barrel motif found in numerous viral capsid structures (2, 36, 39) with individual β-strands separated by large insertion loops that form the surface of the virion. Each of the 60 asymmetric units is constructed by the interdigitation of several neighboring VP polypeptides, intimately wrapped around each other, particularly at the icosahedral 3-fold axes, creating an apparently rugged assembly. Projecting “cylinders” at each icosahedral 5-fold axis appear to be critically involved in particle dynamics. These cylinders are created by the parallel juxtaposition of β-hairpin ribbons from each of the 5-fold related capsid proteins and enclose a small central pore (∼8 Å in diameter at its tightest constriction) that penetrates to the particle interior. The cylinders are themselves encircled by highly conserved canyon-like surface depressions of uncertain function. Despite the apparently minimal diameter of the 5-fold pores, genetic and structural approaches implicate the cylinders as the major portals of entry into and egress from the virion (1, 3, 4, 17, 18).The best-documented example of such transit concerns the N-terminal peptides of VP2, which are sequestered within empty particles but start to become surface exposed early during DNA packaging (12, 14). Twenty amino acids from the N termini of most VP2 molecules can be removed from intact virions by proteolysis, leaving a short, N-terminal, glycine-rich sequence that has been modeled into a partially ordered electron density observed stretching through the 5-fold pores of full, but not empty, particles in X-ray diffraction studies (1, 48, 51). For some crystals this observed density indicates that as many as two-thirds of the 5-fold pores may be thus occupied (9). This represents, at most, 8 of the ∼50 VP2 N termini present in the particle, but ultimately, most of these peptides can be proteolytically cleaved from the virion, suggesting that the 5-fold cylinders may open repeatedly during maturation and cell entry to allow the sequential surface exposure of these N termini. Genetic experiments suggest that a 5-fold pore also serves as the entry portal for viral DNA (3, 4, 17, 18). Newly displaced progeny DNA strands are translocated, in a 3′-to-5′ direction (14), into a preformed “empty” capsid in a reaction that is likely driven by the helicase function of the viral replication initiator protein, NS1 (27). Although precisely where the 3′ end of the genome is positioned in the particle remains uncertain, packaging leaves ∼22 nucleotides from the 5′ end of the DNA exposed at the particle exterior, still covalently attached to an externally located NS1 molecule (13).Where or how the viral genome is uncoated in vivo remains uncertain, but the protein chains of the virion are intricately interwoven and largely resistant to proteolytic assault, except for the short VP2 N-terminal peptides. As part of an unrelated study, we recently removed these VP2 peptides from large batches of MVM virions by trypsin digestion at 37°C and then rebanded the digested product on density gradients. Remarkably, virions in a subset of virus stocks appeared to have undergone a reaction that caused them to band at a lower density, indicating that a major structural rearrangement had likely occurred. In this paper we characterize the nature of this rearrangement, explore the conditions that allow it to occur, and discuss the implications of these findings for virion structure and the genome-uncoating process.  相似文献   

15.
Ebola virus (EBOV) protein VP35 is a double-stranded RNA (dsRNA) binding inhibitor of host interferon (IFN)-α/β responses that also functions as a viral polymerase cofactor. Recent structural studies identified key features, including a central basic patch, required for VP35 dsRNA binding activity. To address the functional significance of these VP35 structural features for EBOV replication and pathogenesis, two point mutations, K319A/R322A, that abrogate VP35 dsRNA binding activity and severely impair its suppression of IFN-α/β production were identified. Solution nuclear magnetic resonance (NMR) spectroscopy and X-ray crystallography reveal minimal structural perturbations in the K319A/R322A VP35 double mutant and suggest that loss of basic charge leads to altered function. Recombinant EBOVs encoding the mutant VP35 exhibit, relative to wild-type VP35 viruses, minimal growth attenuation in IFN-defective Vero cells but severe impairment in IFN-competent cells. In guinea pigs, the VP35 mutant virus revealed a complete loss of virulence. Strikingly, the VP35 mutant virus effectively immunized animals against subsequent wild-type EBOV challenge. These in vivo studies, using recombinant EBOV viruses, combined with the accompanying biochemical and structural analyses directly correlate VP35 dsRNA binding and IFN inhibition functions with viral pathogenesis. Moreover, these studies provide a framework for the development of antivirals targeting this critical EBOV virulence factor.Ebola viruses (EBOVs) are zoonotic, enveloped negative-strand RNA viruses belonging to the family Filoviridae which cause lethal viral hemorrhagic fever in humans and nonhuman primates (47). Currently, information regarding EBOV-encoded virulence determinants remains limited. This, coupled with our lack of understanding of biochemical and structural properties of virulence factors, limits efforts to develop novel prophylactic or therapeutic approaches toward these infections.It has been proposed that EBOV-encoded mechanisms to counter innate immune responses, particularly interferon (IFN) responses, are critical to EBOV pathogenesis (7). However, a role for viral immune evasion functions in the pathogenesis of lethal EBOV infection has yet to be demonstrated. Of the eight major EBOV gene products, two viral proteins have been demonstrated to counter host IFN responses. The VP35 protein is a viral polymerase cofactor and structural protein that also inhibits IFN-α/β production by preventing the activation of interferon regulatory factor (IRF)-3 and -7 (3, 4, 8, 24, 27, 34, 41). VP35 also inhibits the activation of PKR, an IFN-induced, double-stranded RNA (dsRNA)-activated kinase with antiviral activity, and inhibits RNA silencing (17, 20, 48). The VP24 protein is a minor structural protein implicated in virus assembly and regulation of viral RNA synthesis, and changes in VP24 coding sequences are also associated with adaptation of EBOVs to mice and guinea pigs (2, 13, 14, 27, 32, 37, 50, 52). Further, VP24 inhibits cellular responses to both IFN-α/β and IFN-γ by preventing the nuclear accumulation of tyrosine-phosphorylated STAT1 (44, 45). The functions of VP35 and VP24 proteins are manifested in EBOV-infected cells by the absence of IRF-3 activation, impaired production of IFN-α/β, and severely reduced expression of IFN-induced genes, even after treatment of infected cells with IFN-α (3, 19, 21, 22, 24, 25, 28).Previous studies proposed that VP35 basic residues 305, 309, and 312 are required for VP35 dsRNA binding activity (26). VP35 residues K309 and R312 were subsequently identified as critical for binding to dsRNA, and mutation of these residues impaired VP35 suppression of IFN-α/β production (8). In vivo, an EBOV engineered to carry a VP35 R312A point mutation exhibited reduced replication in mice (23). However, because the parental recombinant EBOV into which the mutation was built did not cause disease in these animals, the impact of the mutation on viral pathogenesis could not be fully evaluated. Further, the lack of available structural and biochemical data to explain how the R312A mutation affects VP35 function limited avenues for the therapeutic targeting of critical VP35 functions. Recent structural analyses of the VP35 carboxy-terminal interferon inhibitory domain (IID) suggested that additional residues from the central basic patch may contribute to VP35 dsRNA binding activity and IFN-antagonist function (30). However, a direct correlation between dsRNA and IFN inhibitory functions of VP35 with viral pathogenesis is currently lacking.In order to further define the molecular basis for VP35 dsRNA binding and IFN-antagonist function and to define the contribution of these functions to EBOV pathogenesis, an integrated molecular, structural, and virological approach was taken. The data presented below identify two VP35 carboxy-terminal basic amino acids, K319 and R322, as required for its dsRNA binding and IFN-antagonist functions. Interestingly, these residues are outside the region originally identified as being important for dsRNA binding and IFN inhibition (26). However, they lie within the central basic patch identified by prior structural studies (26, 30). Introduction of these mutations (VP35 with these mutations is designated KRA) into recombinant EBOV renders this otherwise fully lethal virus avirulent in guinea pigs. KRA-infected animals also develop EBOV-specific antibodies and become fully resistant to subsequent challenge with wild-type (WT) virus. Our data further reveal that the KRA EBOV is immunogenic and likely replicates to low levels early after infection in vivo. However, the mutant virus is subsequently cleared by host immune responses. These data demonstrate that the VP35 central basic patch is important not only for IFN-antagonist function but also for EBOV immune evasion and pathogenesis in vivo. High-resolution structural analysis, coupled with our in vitro and in vivo analyses of the recombinant Ebola viruses, provides the molecular basis for loss of function by the VP35 mutant and highlights the therapeutic potential of targeting the central basic patch with small-molecule inhibitors and for future vaccine development efforts.  相似文献   

16.
17.
18.
19.
Identifying the functions of human immunodeficiency virus (HIV)-specific CD8+ T cells that are not merely modulated by the level of virus but clearly distinguish patients with immune control from those without such control is of paramount importance. Features of the HIV-specific CD8+ T-cell response in antiretroviral-treated patients (designated Rx <50) and untreated patients (long-term nonprogressors [LTNP]) matched for very low HIV RNA levels were comprehensively examined. The proliferative capacity of HIV-specific CD8+ T cells was not restored in Rx <50 to the level observed in LTNP, even though HIV-specific CD4+ T-cell proliferation in the two patient groups was comparable. This diminished HIV-specific CD8+ T-cell proliferation in Rx <50 was primarily due to a smaller fraction of antigen-specific cells recruited to divide and not to the numbers of divisions that proliferating cells had undergone. Exogenous interleukin-2 (IL-2) induced proliferating cells to divide further but did not rescue the majority of antigen-specific cells with defective proliferation. In addition, differences in HIV-specific CD8+ T-cell proliferation could not be attributed to differences in cellular subsets bearing a memory phenotype, IL-2 production, or PD-1 expression. Although polyfunctionality of HIV-specific CD8+ T cells in Rx <50 was not restored to the levels observed in LTNP despite prolonged suppression of HIV RNA levels, per-cell cytotoxic capacity was the functional feature that most clearly distinguished the cells of LTNP from those of Rx <50. Taken together, these data suggest that there are selective qualitative abnormalities within the HIV-specific CD8+ T-cell compartment that persist under conditions of low levels of antigen.Understanding the features of an effective immune response to human immunodeficiency virus (HIV) is among the most important goals for the design of HIV vaccines and immunotherapies. Most HIV-infected patients develop persistent viremia and CD4+ T-cell decline in the absence of antiviral therapy. However, evidence that immunologic control of HIV is possible can be drawn from a small group of rare patients who maintain normal CD4+ T-cell counts and restrict HIV replication to below 50 copies/ml plasma for up to 25 years without antiretroviral therapy (ART) (4, 22, 31, 40). Historically, these unique individuals were included within heterogeneous cohorts referred to as long-term survivors or long-term nonprogressors (LTNP), categorized solely based on their disease-free survival exceeding 7 to 10 years and their stable CD4+ T-cell counts (21). Over time, it became apparent that only a small subset of individuals within these cohorts had truly nonprogressive infection, maintaining good health with nondeclining CD4+ T-cell counts, and these true nonprogressors tended to have HIV type 1 (HIV-1) RNA levels below the lower detection limits of the newly available assays (23, 31). Some investigators have adopted other designations more recently, including elite controllers, elite suppressors, or HIV controllers. These designations vary by institution and, in some cases, rely only upon viral load measurements without a requirement for stable CD4+ T-cell counts (4, 22, 40). However, for our designation of true LTNP, we employ the inclusion criteria of stable health, nondeclining CD4+ T-cell counts, and maintenance of plasma viral RNA levels below 50 copies/ml without ART (29-31).Several lines of evidence strongly suggest that CD8+ T cells mediate this control of HIV in LTNP. HLA B*5701 is highly overrepresented in these patients, and in B*5701+ patients, the HIV-specific CD8+ T-cell response is largely focused on peptides restricted by the B57 protein (15, 31). In addition, similar control of simian immunodeficiency virus replication has been described in rhesus macaques carrying the Mamu B*08 or B*17 allele (25, 49). In these macaques, CD8+ T-cell depletion studies have strongly suggested that control of viral replication is mediated by CD8+ T cells (14). Although these results support the idea that CD8+ T cells are responsible for immunologic control, the mechanism remains incompletely understood.Several lines of evidence suggest that immunologic control in LTNP is not simply due to differences in autologous virus recognition by CD8+ T cells. The frequencies of CD8+ T cells specific for HIV or individual HIV-encoded gene products in the peripheral blood are not different in LTNP and untreated progressors (reviewed in reference 32). Putative “escape” mutations are found in viruses of both HLAB*57+ LTNP and HLA-matched progressors (4, 6, 28, 33, 34). In addition, comparable frequencies of CD8+ T cells of LTNP and progressors recognize autologous CD4+ T cells infected with the autologous virus (12, 28). Similar observations have recently been made in the rhesus macaque model (26). Collectively, these observations strongly suggest that features of the CD8+ T-cell response associated with immunologic control are not due to quantitative differences in the numbers of HIV-specific cells or to differential abilities of the autologous virus gene products to be recognized between patient groups.Several qualitative features in the HIV-specific CD8+ T-cell response have been associated with immunologic control in LTNP. LTNP have been found to have higher frequencies of “polyfunctional” CD8+ T cells, named for their ability to degranulate and produce multiple cytokines, including interleukin-2 (IL-2) (2, 5, 51). However, these cells comprise an extremely small proportion of the HIV-specific CD8+ T-cell response. In addition, there is considerable overlap between patient groups, and many LTNP have few or no such cells. Compared to those of progressors, HIV-specific CD8+ T cells of LTNP have a dramatically higher proliferative capacity, a greater ability to upregulate granzyme B (GrB) and perforin production, and a greater cytolytic capacity against autologous HIV-infected CD4+ T cells (3, 17, 24, 29, 30). Increased HIV-specific CD8+ T-cell proliferative capacity in LTNP compared to progressors has also been associated with lower PD-1 expression or IL-2 production by HIV-specific CD4+ or CD8+ T cells (11, 24, 48, 51).Considerable controversy exists over the cause-and-effect relationships between these qualitative differences in the CD8+ T-cell response and HIV viremia between patient groups. High levels of antigen can have potent effects on diverse cell types in humans and in animal models. For HIV, lowering the level of viremia through ART has been observed to increase the function of CD4+ and CD8+ T cells, NK cells, monocytes, and plasmacytoid dendritic cells (16, 18, 20, 37, 41, 45-47, 50). However, the vast majority of treated progressors will not control HIV replication when ART is interrupted (7, 9, 35), suggesting that many of the qualitative differences in the CD4+ or CD8+ T-cell response between LTNP and untreated progressors are not the cause of control over HIV but rather are likely an effect of viremia. In some but not all studies, ART was sufficient to restore the proliferative capacity, phenotype, and cytokine production by CD4+ T cells to levels similar to responses to other viruses or to the HIV-specific response of LTNP (13, 16, 18, 20, 37, 46, 50). Because better IL-2 production or function of HIV-specific CD4+ T cells has been associated with increased CD8+ T-cell proliferative capacity (24), it has also been suggested that diminished proliferative capacity of progressor CD8+ T cells may be an effect of viremia during the chronic phase of infection. In some studies, ART is sufficient to increase the frequency of polyfunctional HIV-specific CD8+ T cells or to decrease PD-1 expression (30, 41). However, the interpretations of the observations within these studies have relied on extrapolations between studies based upon cohorts with differing levels and durations of viral suppression or on examination of a limited number of functions or subsets in either CD4+ or CD8+ T cells.In the present study, we extended our earlier work and comprehensively examined a broad array of functions of HIV-specific T cells derived from two large patient groups, LTNP and progressors on ART, who possess comparable levels of HIV viremia as determined by a sensitive single-copy assay. In response to autologous HIV-infected CD4+ T cells, HIV-specific CD8+ T-cell proliferative capacity, IL-2 responsiveness, surface phenotype, PD-1 expression, polyfunctionality, and cytotoxic capacity were measured in considerable detail. We observe that although ART results in restoration of many of these functions, HIV-specific CD8+ T-cell polyfunctionality and proliferative and killing capacities are not restored to levels observed in LTNP.  相似文献   

20.
Rapid depletion of memory CD4+ T cells and delayed induction of neutralizing antibody (NAb) responses are characteristics of human immunodeficiency virus (HIV) and simian immunodeficiency virus (SIV) infections. Although it was speculated that postinfection NAb induction could have only a limited suppressive effect on primary HIV replication, a recent study has shown that a single passive NAb immunization of rhesus macaques 1 week after SIV challenge can result in reduction of viral loads at the set point, indicating a possible contribution of postinfection NAb responses to virus control. However, the mechanism accounting for this NAb-triggered SIV control has remained unclear. Here, we report rapid induction of virus-specific polyfunctional T-cell responses after the passive NAb immunization postinfection. Analysis of SIV Gag-specific responses of gamma interferon, tumor necrosis factor alpha, interleukin-2, macrophage inflammatory protein 1β, and CD107a revealed that the polyfunctionality of Gag-specific CD4+ T cells, as defined by the multiplicity of these responses, was markedly elevated in the acute phase in NAb-immunized animals. In the chronic phase, despite the absence of detectable NAbs, virus control was maintained, accompanied by polyfunctional Gag-specific T-cell responses. These results implicate virus-specific polyfunctional CD4+ T-cell responses in this NAb-triggered virus control, suggesting possible synergism between NAbs and T cells for control of HIV/SIV replication.Virus-specific CD4+ and CD8+ T-cell responses are crucial for the control of pathogenic human immunodeficiency virus type 1 (HIV-1) and simian immunodeficiency virus (SIV) infections (5, 6, 20, 23, 30, 39, 40). However, CD4+ T cells, especially CCR5+ memory CD4+ T cells, are themselves targets for these viruses, which may be an obstacle to potent virus-specific CD4+ T-cell induction (10, 47, 52). Indeed, HIV-1/SIV infection causes rapid, massive depletion of memory CD4+ T cells (26, 31), and host immune responses fail to contain viral replication and allow persistent chronic infection, although virus-specific CD8+ T-cell responses exert suppressive pressure on viral replication (15).Recently, the importance of T-cell quality in virus containment has been high-lighted, and T-cell polyfunctionality, which is defined by their multiplicity of antigen-specific cytokine production, has been analyzed as an indicator of T-cell quality (4, 8, 11, 41). However, there has been no evidence indicating an association of polyfunctional T-cell responses in the acute phase with HIV-1/SIV control. Even in the chronic phase, whether polyfunctional CD4+ T-cell responses may be associated with virus control has been unclear, although an inverse correlation between polyfunctional CD8+ T-cell responses and viral loads has been shown in HIV-1-infected individuals (4).Another characteristic of HIV-1/SIV infections is the absence of potent neutralizing antibody (NAb) induction during the acute phase (7). This is mainly due to the unusually neutralization-resistant nature of the virus, such as masking of target epitopes in viral envelope proteins (24). Whether this lack of effective NAb response contributes to the failure to control the virus, and whether NAb induction in the acute phase can contribute to virus control, remains unclear. Previous studies documenting virus escape from NAb recognition suggested that NAbs can also exert selective pressure on viral replication to a certain extent (38, 45, 49), but it was speculated that postinfection NAb induction could have only a limited suppressive effect on primary HIV-1/SIV replication (34, 37).By passive NAb immunization of rhesus macaques after SIV challenge, we recently provided evidence indicating that the presence of NAbs during the acute phase can result in SIV control (50). In that study, passive NAb immunization 1 week after SIVmac239 challenge resulted in transient detectable NAb responses followed by reduction in set point viral loads compared to unimmunized macaques. However, the mechanism of this virus control has remained unclear. In the present study, we found rapid appearance of polyfunctional Gag-specific CD4+ T-cell responses after such passive NAb immunization postinfection. These animals maintained virus control for more than 1 year in the absence of detectable plasma NAbs, which was accompanied by potent Gag-specific T-cell responses. These results implicate virus-specific polyfunctional CD4+ T-cell responses in this NAb-triggered primary and long-term SIV control.  相似文献   

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