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 46 毫秒
1.
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   

2.
The Nef protein of human immunodeficiency virus type 1 downregulates the CD4 coreceptor from the surface of host cells by accelerating the rate of CD4 endocytosis through a clathrin/AP-2 pathway. Herein, we report that Nef has the additional function of targeting CD4 to the multivesicular body (MVB) pathway for eventual delivery to lysosomes. This targeting involves the endosomal sorting complex required for transport (ESCRT) machinery. Perturbation of this machinery does not prevent removal of CD4 from the cell surface but precludes its lysosomal degradation, indicating that accelerated endocytosis and targeting to the MVB pathway are separate functions of Nef. We also show that both CD4 and Nef are ubiquitinated on lysine residues, but this modification is dispensable for Nef-induced targeting of CD4 to the MVB pathway.Primate immunodeficiency viruses infect helper T lymphocytes and cells of the macrophage/monocyte lineage by binding of their viral envelope glycoprotein, Env, to a combination of two host cell-specific surface proteins, CD4 and either the CCR5 or CXCR4 chemokine receptors (reviewed in reference 62). Ensuing fusion of the viral envelope with the host cell plasma membrane delivers the viral genetic material into the cytoplasm. Remarkably, the most highly transcribed viral gene in the early phase of infection does not encode an enzyme or structural protein but an accessory protein named Nef. Early expression of Nef is thought to reprogram the host cell for optimal replication of the virus. Indeed, Nef has been shown to enhance virus production (19, 24, 59, 74) and to promote progression to AIDS (23, 47, 48), making it an attractive candidate for pharmacologic intervention.Nef is an N-terminally myristoylated protein with a molecular mass of 27 kDa for human immunodeficiency virus type 1 (HIV-1) and 35 kDa for HIV-2 and simian immunodeficiency virus (27, 29, 50, 65). Nef has been ascribed many functions, the best characterized of which is the downregulation of the CD4 coreceptor from the surface of infected cells (28, 35, 57). CD4 downregulation is believed to prevent superinfection (8, 52) and to preclude the cellular retention of newly synthesized Env (8, 49), thus allowing the establishment of a robust infection (30, 71).The molecular mechanism by which Nef downregulates CD4 has been extensively studied. A consensus has emerged that Nef accelerates the endocytosis of cell surface CD4 (2, 64) by linking the cytosolic tail of CD4 to the heterotetrameric (α-β2-μ2-σ2) adaptor protein-2 (AP-2) complex (17, 25, 34, 45, 67). Determinants in the CD4 tail bind to a hydrophobic pocket comprising tryptophan-57 and leucine-58 on the folded core domain of Nef (34). On the other hand, a dileucine motif (i.e., ENTSLL, residues 160 to 165) (14, 22, 32) and a diacidic motif (i.e., DD, residues 174 and 175) (3) (residues correspond to the NL4-3 clone of HIV-1) within a C-terminal, flexible loop of Nef bind to the α and σ2 subunits of AP-2 (17, 18, 25, 51). AP-2, in turn, binds to clathrin, leading to the concentration of CD4 within clathrin-coated pits (15, 33). These pits eventually bud from the plasma membrane as clathrin-coated vesicles that deliver internalized CD4 to endosomes. In essence, then, Nef acts as a connector that confers on CD4 the ability to be rapidly internalized in a manner similar to endocytic receptors (75).Unlike typical endocytic recycling receptors like the transferrin receptor or the low-density lipoprotein receptor, however, CD4 that is forcibly internalized by Nef does not return to the cell surface but is delivered to lysosomes for degradation (4, 64, 68). Thus, expression of Nef decreases both the surface and total levels of CD4. What keeps internalized CD4 from returning to the plasma membrane? We hypothesized that Nef might additionally act on endosomes to direct CD4 to lysosomes. This is precisely the fate followed by signaling receptors, transporters, and other transmembrane proteins that undergo ubiquitination-mediated internalization and targeting to the multivesicular body (MVB) pathway (40, 46). This targeting involves the endosomal sorting complex required for transport (ESCRT), including the ESCRT-0, -I, -II, and -III complexes, which function to sort ubiquitinated cargoes into intraluminal vesicles of MVBs for eventual degradation in lysosomes (40, 46). Herein, we show that Nef indeed plays a novel role in targeting internalized CD4 from endosomes to the MVB pathway in an ESCRT-dependent manner. We also show that both Nef and CD4 undergo ubiquitination on lysine residues, but, strikingly, this modification is not required for CD4 targeting to the MVB pathway.  相似文献   

3.
4.
A critical function of the human immunodeficiency virus type 1 Nef protein is the downregulation of CD4 from the surfaces of infected cells. Nef is believed to act by linking the cytosolic tail of CD4 to the endocytic machinery, thereby increasing the rate of CD4 internalization. In support of this model, weak binary interactions between CD4, Nef, and the endocytic adaptor complex, AP-2, have been reported. In particular, dileucine and diacidic motifs in the C-terminal flexible loop of Nef have been shown to mediate binding to a combination of the α and σ2 subunits of AP-2. Here, we report the identification of a potential binding site for the Nef diacidic motif on α-adaptin. This site comprises two basic residues, lysine-297 and arginine-340, on the α-adaptin trunk domain. The mutation of these residues specifically inhibits the ability of Nef to bind AP-2 and downregulate CD4. We also present evidence that the diacidic motif on Nef and the basic patch on α-adaptin are both required for the cooperative assembly of a CD4-Nef-AP-2 complex. This cooperativity explains how Nef is able to efficiently downregulate CD4 despite weak binary interactions between components of the tripartite complex.CD4, a type I transmembrane glycoprotein that serves as a coreceptor for major histocompatibility complex class II (MHC-II) molecules, is expressed on the surfaces of helper T lymphocytes and cells of the monocyte/macrophage lineage (8). Primate immunodeficiency viruses gain access to these cells by virtue of the interaction of the viral envelope glycoprotein (Env) with a combination of CD4 and a chemokine receptor (63). This interaction causes a conformational change within the Env protein that promotes the fusion of the viral envelope with the plasma membrane. Upon the delivery of the viral genetic material into the cytoplasm of the host cells, one of the first virally encoded proteins to be expressed is Nef, an accessory factor that modulates specific signal transduction and protein-trafficking pathways in a manner that optimizes the intracellular environment for viral replication (reviewed in references 21, 39, and 65). Perhaps the best characterized function of Nef is the downregulation of CD4 from the surfaces of the host cells (6, 22, 29, 45). CD4 downregulation prevents superinfection (6, 41) and enhances virion release (19, 38, 48, 66, 76), thereby contributing to the establishment of a robust infective state (24, 72).The mechanism used by the Nef protein of human immunodeficiency virus type 1 (HIV-1) to downregulate CD4 has been the subject of extensive study, but only recently have the molecular details of this process begun to be unraveled. It is generally acknowledged that HIV-1 Nef accelerates the internalization of CD4 from the plasma membrane by linking the cytosolic tail of the receptor to the clathrin-associated endocytic machinery (1, 12, 20, 34, 40, 64). In support of this model, a hydrophobic pocket comprising W57 and L58 on the folded core domain of Nef binds with millimolar affinity to the cytosolic tail of CD4 (28) (all residues and numbers correspond to the NL4-3 variant of HIV-1 Nef used in this study). In addition, a dileucine motif (ENTSLL, residues 160 to 165) (10, 16, 26) and a diacidic motif (D174 and D175) (2) on the C-terminal flexible loop of Nef mediate an interaction of micromolar affinity with the clathrin-associated, heterotetrameric (α-β2-μ2-σ2) adaptor protein 2 (AP-2) complex (12, 20, 40, 49). These interactions draw CD4 into clathrin-coated pits that eventually bud inwards as clathrin-coated vesicles (11, 27). Internalized CD4 is subsequently delivered to endosomes and then to lysosomes for degradation (3, 23, 59, 64).Despite progress in the understanding of the mechanism of Nef-induced CD4 downregulation, several important aspects remain to be elucidated. Previous studies have shown that the Nef dileucine and diacidic motifs interact with a combination of the α and σ2 subunits of AP-2 (referred to as the α-σ2 hemicomplex) (12, 20, 40, 49), but the precise location of the Nef binding sites is unknown. It also remains to be determined whether Nef can actually bind CD4 and AP-2 at the same time. Indeed, the formation of a tripartite CD4-Nef-AP-2 complex in which Nef links the cytosolic tail of CD4 to AP-2 has long been hypothesized but has never been demonstrated experimentally. Given the relatively weak affinity of Nef for the CD4 tail (28) and AP-2 (12, 40), it is unclear how such a complex could assemble and function in CD4 downregulation.In this study, we have addressed these issues by using a combination of yeast hybrid, in vitro binding, and in vivo CD4 downregulation assays. We report the identification of a candidate binding site for the Nef diacidic motif on the AP-2 complex. This site, a basic patch comprising K297 and R340 on α-adaptin, is specifically required for Nef binding and Nef-induced CD4 downregulation. We also show that the Nef diacidic motif and the α-adaptin basic patch are required for the cooperative assembly of a tripartite complex composed of the CD4 cytosolic tail, Nef, and the α-σ2 hemicomplex. The cooperative manner in which this complex is formed explains how Nef is able to efficiently downregulate CD4 from the plasma membrane despite weak binary interactions between the components of this complex.  相似文献   

5.
6.
7.
Human immunodeficiency virus type 1 (HIV-1) Nef interferes with the endocytic machinery to modulate the cell surface expression of CD4. However, the basal trafficking of CD4 is governed by different rules in the target cells of HIV-1: whereas CD4 is rapidly internalized from the cell surface in myeloid cells, CD4 is stabilized at the plasma membrane through its interaction with the p56lck kinase in lymphoid cells. In this study, we showed that Nef was able to downregulate CD4 in both lymphoid and myeloid cell lines but that an increase in the internalization rate of CD4 could be observed only in lymphoid cells. Expression of p56lck in nonlymphoid CD4-expressing cells restores the ability of Nef in order to increase the internalization rate of CD4. Concurrent with this observation, the expression of a p56lck-binding-deficient mutant of CD4 in lymphoid cells abrogates the Nef-induced acceleration of CD4 internalization. We also show that the expression of Nef causes a decrease in the association of p56lck with cell surface-expressed CD4. Regardless of the presence of p56lck, the downregulation of CD4 by Nef was followed by CD4 degradation. Our results imply that Nef uses distinct mechanisms to downregulate the cell surface expression levels of CD4 in either lymphoid or myeloid target cells of HIV-1.Besides proteins that are essential for proper virus processing and assembly, the genomes of primate lentiviruses such as human immunodeficiency virus type 1 (HIV-1) encode auxiliary proteins that modulate viral infectivity. The 27-kDa auxiliary protein Nef is a key element in the progression of primary HIV-1 infection toward AIDS. Cases of patients infected with HIV-1 strains harboring a deletion in the nef gene or a defective nef allele have been reported. Some of these patients exhibit asymptomatic or slow progression toward the disease (6, 17, 37). In vitro, Nef facilitates viral replication and enhances the infectivity of viral particles (13, 47, 69). The mechanisms involved in the Nef-induced increase of viral infectivity remain elusive; however, it is a multifactorial process related to the ability of Nef to alter the trafficking of host cell proteins.Indeed, the most documented effect of Nef during the course of viral infection is its ability to disturb the clathrin-dependent trafficking machinery involved in the transport of transmembrane proteins through endosomal compartments. This leads to the modulation of the level of cell surface expression for some receptors, including CD4, which is the primary receptor of HIV-1 (35) and major histocompatibility complex class I (reviewed in references 22 and 27). The downregulation of CD4, which results in the impairment of the immunological synapse (72) and the downregulation of major histocompatibility complex class I molecules (reviewed in reference 16), is believed to contribute to the escape of HIV-1-infected cells from immunosurveillance. Moreover, the downregulation of CD4 helps avoid superinfection of cells, which would be deleterious to the virus (reviewed in reference 21), and has a direct impact on viral fitness by allowing better incorporation of the functional envelope in viral particles produced from CD4-expressing cells (3, 36, 53).Nef-induced cell surface downregulation of CD4 is efficient in all CD4-expressing cells and depends on the integrity of a di-Leu motif at position 164/165 of the C-terminal flexible loop of HIV-1 Nef (2, 9, 25). This di-Leu motif allows for the interaction with clathrin-associated adaptor protein (AP) complexes that participate in the clathrin-dependent vesicular transport within the endocytic pathway. The AP type 2 (AP-2) complex is localized at the plasma membrane and is essential to the assembly and function of clathrin-coated pits involved in the internalization of receptors from the cell surface (59). The interaction of Nef with AP-2 is well delineated and has been proposed to enhance the targeting of CD4 to clathrin-coated pits and its internalization (10, 12, 26, 32, 39).Helper T lymphocytes are the predominant cell type that expresses CD4; however, CD4 is also present at the surfaces of monocytes and macrophages (70), where its function is yet to be elucidated. Whereas cell surface CD4 is rapidly internalized in myeloid cells, CD4 is stabilized at the plasma membrane in lymphoid cells through its interaction with the Src family protein tyrosine kinase p56lck. Cys residues located at positions 420/422 in the CD4 cytoplasmic tail are essential to the constitutive association with p56lck (73). Besides its role in signal transduction, this interaction also correlates with an accumulation of CD4 in lipid rafts and enhanced exclusion of CD4 from clathrin-coated pits (50).In T cells, treatment with phorbol esters such as phorbol 12-myristate 13-acetate (PMA) provokes the phosphorylation of Ser residues found in the cytoplasmic tail of CD4. This correlates with a decreased association of p56lck with CD4 and the internalization of the receptor (24, 32-34, 41, 45, 48, 52, 56, 61, 66-68). Nef-induced CD4 downregulation is known to be independent of Ser phosphorylation (20) and is therefore governed by mechanisms different from those involved in PMA-induced CD4 downregulation. However, the Leu-based sorting motif in the CD4 cytoplasmic tail is critical for both PMA and Nef-induced CD4 downregulation (2, 5, 24, 31, 56, 60, 68), thus indicating that despite being different, the mechanisms involved in Nef- and PMA-induced CD4 downregulation partially overlap.In the present study, we investigated whether the mechanisms used by Nef to downregulate CD4 are cell type-dependent processes. We looked at the trafficking and steady-state expression of CD4 in the main target cells of HIV-1, CD4-positive T lymphocytes, and cells of the monocyte/macrophage lineage. Our results demonstrate that the presence of p56lck has a direct impact on the mechanisms used by Nef to downregulate CD4 from the cell surface of T lymphocytes. They also reveal that Nef uses distinct pathways to decrease levels of cell surface expression of CD4 in lymphoid or myeloid target cells of HIV-1.  相似文献   

8.
The HIV-1 Nef protein is a pathogenic factor modulating the behavior of infected cells. Nef induces actin cytoskeleton changes and impairs cell migration toward chemokines. We further characterized the morphology, cytoskeleton dynamics, and motility of HIV-1-infected lymphocytes. By using scanning electron microscopy, confocal immunofluorescence microscopy, and ImageStream technology, which combines flow cytometry and automated imaging, we report that HIV-1 induces a characteristic remodeling of the actin cytoskeleton. In infected lymphocytes, ruffle formation is inhibited, whereas long, thin filopodium-like protrusions are induced. Cells infected with HIV with nef deleted display a normal phenotype, and Nef expression alone, in the absence of other viral proteins, induces morphological changes. We also used an innovative imaging system to immobilize and visualize living individual cells in suspension. When combined with confocal “axial tomography,” this technique greatly enhances three-dimensional optical resolution. With this technique, we confirmed the induction of long filopodium-like structures in unfixed Nef-expressing lymphocytes. The cytoskeleton reorganization induced by Nef is associated with an important impairment of cell movements. The adhesion and spreading of infected cells to fibronectin, their spontaneous motility, and their migration toward chemokines (CXCL12, CCL3, and CCL19) were all significantly decreased. Therefore, Nef induces complex effects on the lymphocyte actin cytoskeleton and cellular morphology, which likely impacts the capacity of infected cells to circulate and to encounter and communicate with bystander cells.Human immunodeficiency virus type 1 (HIV-1) mostly replicates in T-cell areas of secondary lymphoid organs (SLOs) and induces pathological changes in their architecture. Such changes are likely due to a combination of events, including destruction of T cells, chronic immune activation, and alteration of T-cell motility toward and inside the SLOs (27, 37, 50, 53). Indeed, to fulfill their immune surveillance role, T cells continuously circulate in and out of blood, lymph nodes (LNs), and tissues (60).Lymphocyte recruitment from the bloodstream into LNs depends on three distinct processes, i.e., attachment to high endothelial venules (HEVs), extravasation, and cell migration (10, 60). Adhesion to the endothelium and extracellular matrix (ECM) is a crucial step, regulated in part by β1 integrins, α4β1 (VLA-4) and α5β1, that bind VCAM-1 and/or fibronectin (56). Chemokines and their Gαi-protein-coupled receptors are key regulators of lymphocyte trafficking (32). For instance, CCL19 and CCL21 are constitutively produced by HEVs and by fibroblastic reticular cells of T-cell areas of LNs (21, 28, 29). These two chemokines share the receptor CCR7, expressed by naïve T cells and a fraction of memory T cells (47). They play a major role in lymphocyte homing to LNs, in steady state as well as under conditions of inflammation, and may control T-cell positioning within defined functional compartments (1, 17, 18, 47). CXCR4 and its ligand CXCL12/SDF-1 also contribute to T-cell entry into LNs (5, 23, 40). In addition, effector and memory T cells express a broad range of receptors binding inflammatory chemokines, such as the CCR5 ligands CCL3 (MIP1α), CCL4 (MIP1β), and CCL5 (Rantes).Efficient accomplishment of lymphocyte migration and immune functions requires tight regulation of the cellular cytoskeleton (59). This is mediated by the small GTPases of the Rho subfamily, such as Rho, Rac, and Cdc42 (11, 58). They activate specific actin filament assembly factors to generate sheet-like protrusive structures (such as lamellipodia and ruffles) and finger-like protrusions (such as filopodia and microvilli) (6). These structures have different functions. Lamellipodia and ruffles are formed during crawling cell motility and spreading. Filopodia protrude from the leading edges of many motile cells. They appear to perform sensory and exploratory functions to steer cells, depending on cues from the environment (42). Moreover, filopodia, or other thin structures called tunneling nanotubes, have been shown to form intercellular bridges, allowing viruses to spread through remote contacts between infected cells and targets (44, 48, 49, 52).HIV-1 hijacks cytoskeleton dynamics in order to ensure viral entry and transport within and egress from target cells (34; reviewed in reference 13). In particular, the viral protein Nef modifies actin remodeling in various cell systems. In T cells, Nef alters actin rearrangements triggered by activation of T-cell (TCR) or chemokine receptors (22, 54). Nef inhibits immunological synapse formation, a dynamic process involving rapid actin modifications (57). Nef also affects plasma membrane plasticity, inducing secretion of microvesicle clusters (33). In macrophages, Nef induces the extension of long intercellular conduits allowing its own transfer to B cells (61). A number of studies have reported that Nef affects T-cell chemotaxis (generally to CXCL12) through the modulation of Rho-GTPase-regulated signaling pathways (7, 24, 39, 54). Migration studies have generally been performed using Nef-expressing cells, and rarely in the context of HIV-1 infection (54). From a molecular standpoint, it has recently been proposed that Nef acts in part by deregulating cofilin, an actin-depolymerizing factor that promotes actin turnover and subsequent cell motility (54).In the present study, our goal was to gain further insights into the effect of HIV-1 infection on cytoskeleton dynamics. We used a panel of innovative techniques allowing analysis of cell shape, adhesion, and motility. We report that in HIV-infected lymphocytes, Nef promotes filopodium-like formation while it inhibits membrane ruffling. Nef impairs cell adhesion on the extracellular matrix and decreases intrinsic cell motility. Lymphocyte migration toward various chemokines (CXCL12, CCL3, and CCL19) is also inhibited. Our results suggest that Nef may facilitate viral spread and contribute to AIDS pathogenesis by manipulating the migration of lymphocytes.  相似文献   

9.
Chronic immune activation is thought to play a major role in human immunodeficiency virus (HIV) pathogenesis, but the relative contributions of multiple factors to immune activation are not known. One proposed mechanism to protect against immune activation is the ability of Nef proteins from some HIV and simian immunodeficiency virus strains to downregulate the T-cell receptor (TCR)-CD3 complex of the infected cell, thereby reducing the potential for deleterious activation. HIV type 1 (HIV-1) Nef has lost this property. In contrast to HIV-1, HIV-2 infection is characterized by a marked disparity in the disease course, with most individuals maintaining a normal life span. In this study, we examined the relationship between the ability of HIV-2 Nef proteins to downregulate the TCR and immune activation, comparing progressors and nonprogressors. Representative Nef variants were isolated from 28 HIV-2-infected individuals. We assessed their abilities to downregulate the TCR from the surfaces of CD4 T cells. In the same individuals, the activation of peripheral lymphocytes was evaluated by measurement of the expression levels of HLA-DR and CD38. We observed a striking correlation of the TCR downregulation efficiency of HIV-2 Nef variants with immune activation in individuals with a low viral load. This strongly suggests that Nef expression can influence the activation state of the immune systems of infected individuals. However, the efficiency of TCR downregulation by Nef was not reduced in progressing individuals, showing that TCR downregulation does not protect against progression in HIV-2 infection.The majority of humans infected with human immunodeficiency virus type 1 (HIV-1) progress relentlessly toward immunodeficiency, whereas simian immunodeficiency virus (SIV) infection in the natural hosts, Old World monkeys, rarely causes disease (9). It was recently shown that HIV-1 and its simian ancestor, SIVcpz, have one distinctive characteristic that may contribute to pathogenesis. In contrast to the Nef proteins of other immunodeficiency viruses, HIV-1 and SIVcpz Nef proteins are unable to downregulate the T-cell receptor (TCR) from the surfaces of infected cells (1, 22). Schindler and colleagues proposed that TCR downregulation protects the host from the impact of chronic immune activation (22), which is increasingly thought to play a major role in HIV-1 disease progression (7). In most cases, SIVsmm infection of sooty mangabeys leads to high viral loads without evidence of immunodeficiency or CD4 depletion, and this is associated with very low levels of immune activation (25). CD4 depletion without immunodeficiency has been reported in a minority of SIVsmm-infected sooty mangabeys. However, this CD4 depletion is not associated with major immune activation or viral-load increase (26). Immunodeficiency associated with CD4 depletion was reported in only one case (18). Schindler et al. discovered that in sooty mangabeys showing a loss of CD4+ T cells, the Nef protein of the infecting SIVsmm was less efficient at TCR downregulation (22), suggesting that the CD4 depletion in sooty mangabeys is linked to the loss of this function, together with a loss of major histocompatibility complex class I downregulation (23). Following transmission to humans in West Africa, SIVsmm zoonosis gave rise to HIV-2 infection, identified in patients with AIDS in 1986 (10). HIV-2 infection can lead to a clinical picture indistinguishable from AIDS caused by HIV-1, but in general, the progress to clinical immunodeficiency is slower than in HIV-1 infection: this appears to be due to an unusually high proportion of HIV-2-infected long-term nonprogressors (8, 21). Although the few HIV-2 nef alleles that have been studied so far are capable of TCR downregulation, this has not been systematically evaluated in relation to disease progression. Here, we present data from a well-characterized community cohort followed in Caio in Guinea-Bissau since 1989 (27), in which the abilities of nef alleles from the infecting HIV-2 strains to downregulate the TCR could be studied in relation to immune activation and disease status.  相似文献   

10.
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   

11.
12.
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   

13.
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   

14.
15.
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   

16.
17.
The generation of humanized BLT mice by the cotransplantation of human fetal thymus and liver tissues and CD34+ fetal liver cells into nonobese diabetic/severe combined immunodeficiency mice allows for the long-term reconstitution of a functional human immune system, with human T cells, B cells, dendritic cells, and monocytes/macrophages repopulating mouse tissues. Here, we show that humanized BLT mice sustained high-level disseminated human immunodeficiency virus (HIV) infection, resulting in CD4+ T-cell depletion and generalized immune activation. Following infection, HIV-specific humoral responses were present in all mice by 3 months, and HIV-specific CD4+ and CD8+ T-cell responses were detected in the majority of mice tested after 9 weeks of infection. Despite robust HIV-specific responses, however, viral loads remained elevated in infected BLT mice, raising the possibility that these responses are dysfunctional. The increased T-cell expression of the negative costimulator PD-1 recently has been postulated to contribute to T-cell dysfunction in chronic HIV infection. As seen in human infection, both CD4+ and CD8+ T cells demonstrated increased PD-1 expression in HIV-infected BLT mice, and PD-1 levels in these cells correlated positively with viral load and inversely with CD4+ cell levels. The ability of humanized BLT mice to generate both cellular and humoral immune responses to HIV will allow the further investigation of human HIV-specific immune responses in vivo and suggests that these mice are able to provide a platform to assess candidate HIV vaccines and other immunotherapeutic strategies.An ideal animal model of human immunodeficiency virus (HIV) infection remains elusive. Nonhuman primates that are susceptible to HIV infection typically do not develop immunodeficiency (63), and although the simian immunodeficiency virus (SIV) infection of rhesus macaques has provided many critically important insights into retroviral pathogenesis (30), biological and financial considerations have created some limitations to the wide dissemination of this model. The great need for an improved animal model of HIV itself recently has been underscored by the disappointing results of human trials of MRKAd5, an adenovirus-based HIV type 1 (HIV-1) vaccine. This vaccine was not effective and actually may have increased some subjects'' risk of acquiring HIV (53). In the wake of these disappointing results, there has been increased interest in humanized mouse models of HIV infection (54). The ability of humanized mouse models to test candidate vaccines or other immunomodulatory strategies will depend critically on the ability of these mice to generate robust anti-HIV human immune responses.Mice have provided important model systems for the study of many human diseases, but they are unable to support productive HIV infection, even when made to express human coreceptors for the virus (7, 37, 52). A more successful strategy to humanize mice has been to engraft human immune cells and/or tissues into immunodeficient severe combined immunodeficiency (SCID) or nonobese diabetic (NOD)/SCID mice that are unable to reject xenogeneic grafts (39, 42, 57). Early versions of humanized mice supported productive HIV infection and allowed investigators to begin to address important questions in HIV biology in vivo (23, 40, 43-45). More recently, human cord blood or fetal liver CD34+ cells have been used to reconstitute Rag2−/− interleukin-2 receptor γ chain-deficient (γc−/−) and NOD/SCID/γc−/− mice, resulting in higher levels of sustained human immune cell engraftment (27, 29, 61). These mice have allowed for stable, disseminated HIV infection (2, 4, 24, 65, 67), including mucosal transmission via vaginal and rectal routes (3). These mice recently have been used to demonstrate an important role for Treg cells in acute HIV infection (29) and to demonstrate that the T-cell-specific delivery of antiviral small interfering RNA is able to suppress HIV replication in vivo (31). These mice also have demonstrated some evidence of adaptive human immune responses, including the generation of HIV-specific antibody responses in some infected mice (2, 65), and some evidence of humoral and cell-mediated responses to non-HIV antigens or pathogens (24, 61). Most impressively, Rag2−/− γc−/− mice reconstituted with human fetal liver-derived CD34+ cells have generated humoral responses to dengue virus infection that demonstrated both class switching and neutralizing capacity (32). In spite of these advances, however, these models have not yet been reported to generate de novo HIV-specific cell-mediated immune responses, which are considered to be a crucial arm of host defense against HIV infection in humans.In contrast to humanized mouse models in which only human hematopoietic cells are transferred into immunodeficient mice, the surgical implantation of human fetal thymic and liver tissue has been performed in addition to the transfer of human hematopoietic stem cells (HSC) to generate mice in which human T cells are educated by autologous human thymic tissue rather than by the xenogeneic mouse thymus. Melkus and colleagues refer to mice they have reconstituted in this way as NOD/SCID-hu BLT (for bone marrow, liver, and thymus), or simply BLT, mice (41). We previously referred to mice that we have humanized in a similar way as NOD/SCID mice cotransplanted with human fetal thymic and liver tissues (Thy/Liv) and CD34+ fetal liver cells (FLC) (33, 60) but now adopt the designation BLT mice as well. BLT mice demonstrate the robust repopulation of mouse lymphoid tissues with functional human T lymphocytes (33, 41, 60) and can support the rectal and vaginal transmission of HIV (13, 59). Further, BLT mice demonstrate antigen-specific human immune responses against non-HIV antigens and/or pathogens (41, 60). The ability of these mice to generate human immune responses against HIV, however, has not yet been reported. In this study, we investigated whether the provision of autologous human thymic tissue in BLT mice generated by the cotransplantion of human fetal Thy/Liv tissues and CD34+ FLC would allow for the maturation of human T cells in humanized mice capable of providing improved cellular responses to HIV as well as providing adequate help for improved humoral responses. To describe the cells contributing to human immune responses in BLT mice, we also characterized the phenotypes of multiple subsets of T cells, B cells, dendritic cells (DCs), and monocytes/macrophages present in uninfected humanized mice. The generation of robust HIV-directed human cellular and humoral immune responses in these mice would further demonstrate the ability of humanized mice to provide a much needed platform for the evaluation of HIV vaccines and other novel immunomodulatory strategies.  相似文献   

18.
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   

19.
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