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 546 毫秒
1.
Genital herpes, caused by herpes simplex virus type 2 (HSV-2), is one of the most prevalent sexually transmitted diseases worldwide and a risk factor for acquiring human immunodeficiency virus. Although many vaccine candidates have shown promising results in animal models, they have failed to be effective in human trials. In this study, a humanized mouse strain was evaluated as a potential preclinical model for studying human immune responses to HSV-2 infection and vaccination. Immunodeficient mouse strains were examined for their abilities to develop human innate and adaptive immune cells after transplantation of human umbilical cord stem cells. A RAG2−/− γc−/− mouse strain with a BALB/c background was chosen as the most appropriate model and was then examined for its ability to mount innate and adaptive immune responses to intravaginal HSV-2 infection and immunization. After primary infection, human cells in the lymph nodes were able to generate a protective innate immune response and produce gamma interferon (IFN-γ). After intravaginal immunization and infection, human T cells and NK cells were found in the genital tract and iliac lymph nodes. In addition, human T cells in the spleen, lymph nodes, and vaginal tract were able to respond to stimulation with HSV-2 antigens by replicating and producing IFN-γ. Human B cells were also able to produce HSV-2-specific immunoglobulin G. These adaptive responses were also shown to be protective and reduce local viral replication in the genital tract. This approach provides a means for studying human immune responses in vivo using a small-animal model and may become an important preclinical tool.Genital herpes, caused primarily by herpes simplex virus type 2 (HSV-2), is one of the most prevalent sexually transmitted diseases in the world and is associated with substantial morbidity (13). After initial infection of the genital tract, the virus establishes latency within the nervous system and thus maintains lifelong infection in humans. Latent virus can reactivate and cause recurrent symptoms, including genital lesions; however, subclinical infection and asymptomatic viral shedding also occur (11, 35, 40, 53). HSV-2 has gained increasing interest in the light of evidence that it is a major risk factor for human immunodeficiency virus type 1 (HIV-1) acquisition and transmission and for the progression of HIV-1 infection (8, 9, 17, 25, 37, 55, 56). In addition, there is evidence that anti-HSV therapy can reduce the amount of infectious HIV-1 in the genital tracts of women (9, 45). Although antiviral treatment is available and can reduce the severity of the infection, compliance problems, as well as difficulty in diagnosing infection in patients, have hampered efforts to control the disease. A vaccine would provide a more effective way of preventing or limiting infection and would therefore greatly reduce the social and economic burdens caused by HSV-2 infection.Several vaccine candidates exist; however, they have proven to be less successful in clinical trials than anticipated, and new strategies may need to be developed (24, 61). A key concern is that preclinical vaccine strategies have been evaluated largely by using studies performed with mouse models of HSV-2 infection and, thus, the immune responses observed were mediated by murine cells. As a consequence, the results of these studies may not accurately represent the human immune response to infection. In order to develop an effective vaccine and/or treatment, it is necessary to understand which immune mechanisms provide protection against infection at the site of viral entry, the vaginal tract, and how these immune responses can be induced in humans.Innate and adaptive immune responses are both important for controlling HSV-2 infection. Innate immune cells such as NK and NKT cells are required for protection against genital HSV-2 infection in mice (1) and in humans; NK cells accumulate at sites of HSV-2 infection and can lyse HSV-infected cells (30, 67). Adaptive immune responses to HSV-2 include the cellular response mediated by CD4+ and CD8+ T cells and the humoral response mediated by B cells and antibodies. There is much evidence that T cells play a crucial role in protection against HSV-2 in mice and humans (28). T cells are present in herpes lesions, and depletion of T cells in mice greatly reduces protection (16, 27, 29, 30, 44, 51, 70). Gamma interferon (IFN-γ), which is produced early after infection by NK cells and later by CD4+ T cells, has been shown to be a crucial cytokine for the control of HSV (43, 52, 58, 63). Although HSV-2-specific antibodies are produced in response to infection and vaccination, a correlation with protection in humans has not been established (2, 3, 7, 10, 11, 48). In mice, a role for antibodies early after infection has been shown; however, if B cells are knocked out, mice are still able to eventually clear the virus (16, 50). Although we do not have a complete understanding of the components that are necessary for protection, it appears that both innate and adaptive immune responses will be required and that it will be important to elicit these responses at the site of infection in the genital tract.The lack of an effective vaccine and accurate translation of results obtained with mice to humans indicates a need for a more relevant preclinical model to study human immune responses and disease. Substantial improvements in the development of humanized mice have made them a novel tool for the study of human diseases (69). Human CD34+ stem cells have been injected into several immunodeficient mouse strains, such as NOD/SCID/γc−/− and RAG2−/− γc−/− mice, in which superior engraftment has resulted in multilineage differentiation of the human cells (23, 64). These novel humanized mice have been shown to develop human immune responses to pathogens such as Epstein-Barr virus, dengue virus, and influenza virus and to immunization with cholera toxin (33, 64, 66, 68). In addition, humanized mice can support infection with HIV after systemic or mucosal challenge in the vaginal tract and rectum (4-6, 62, 65). HSV-2 infection in humanized mice has not been examined, and mucosal immunization that can provide protection from infection with wild-type virus has also not been demonstrated. In addition, although it is clear that adaptive immune responses can be generated in humanized mice, innate responses to viral infection have not been extensively examined.In this study, we evaluated three immunodeficient mouse strains for their abilities to engraft human umbilical cord-derived stem cells and support the differentiation of these cells into important innate and adaptive immune cells. The most appropriate model was then used to examine mucosal immune responses following primary HSV-2 infection, immunization, and secondary HSV-2 challenge. We show for the first time that the humanized mice can mount protective human NK cell-mediated innate immune responses to primary mucosal infection with HSV-2. In addition, mucosal immunization and infection can induce HSV-2-specific antibody production and, to a greater extent, T-cell-mediated responses both systemically and locally in the genital tracts of humanized mice. We further show that mucosal immunization can provide protection against a lethal intravaginal (IVAG) challenge with HSV-2.  相似文献   

2.
3.
Rag2−/− γC−/− mice transplanted with human hematopoietic stem cells (DKO-hu-HSC mice) mimic aspects of human infection with human immunodeficiency virus type 1 (HIV-1), including sustained viral replication and CD4+ T-cell decline. However, the extent of HIV-1 evolution during long-term infection in these humanized mice, a key feature of the natural infection, has not been assessed fully. In this study, we examined the types of genotypic and phenotypic changes in the viral env gene that occur in the viral populations of DKO-hu-HSC mice infected with the CCR5-tropic isolate HIV-1JRCSF for up to 44 weeks. The mean rate of divergence of viral populations in mice was similar to that observed in a cohort of humans during a similar period of infection. Many amino acid substitutions were common across mice, including losses of N-linked glycosylation sites and substitutions in the CD4 binding site and in CD4-induced epitopes, indicating common selective pressures between mice. In addition, env variants evolved sensitivity to antibodies directed at V3, suggesting a more open conformation for Env. This phenotypic change was associated with increased CD4 binding efficiency and was attributed to specific amino acid substitutions. In one mouse, env variants emerged that exhibited a CXCR4-tropic phenotype. These sequences were compartmentalized in the mesenteric lymph node. In summary, viral populations in these mice exhibited dynamic behavior that included sequence evolution, compartmentalization, and the appearance of distinct phenotypic changes. Thus, humanized mice offer a useful model for studying evolutionary processes of HIV-1 in a complex host environment.Animal models of HIV-1 infection are important tools for studying transmission, replication, and pathogenesis, as well as therapeutic intervention, of HIV-1 infection. Nonhuman primates such as rhesus macaques, infected with simian or chimeric simian/human immunodeficiency viruses (SIV or SHIV, respectively), represent well-characterized and highly relevant models; however, key limitations include expense, genetic variability of the host animals, and the fact that SIV, while closely related, is distinct from HIV-1. Therefore, small animal models that support HIV-1 infection and recapitulate many aspects of the human infection have been sought using several approaches.Recent approaches have involved the use of genetically immunodeficient mice that have been reconstituted using human-derived hematopoietic stem cells (HSC) (known as humanized mice). Several models have been developed based on this approach, including Rag2−/− γC−/− (DKO) and NOD/SCID/γC−/− (NOG or NSG) mice transplanted with human HSC (DKO-hu-HSC or NOG-hu-HSC mice) (40, 92) and the NOD/SCID mouse with transplanted human fetal thymus and liver tissue in addition to HSC (62). These models all support HIV-1 infection (1, 3, 6, 30, 87, 96, 102; for a review of these models, see the work of Denton and Garcia [22]). The DKO-hu-HSC mouse lacks both recombination activating gene 2 (Rag2) and the cytokine receptor common gamma chain (γC), and as a result, it does not generate murine T, B, and natural killer (NK) cells but supports engraftment of HSC and differentiation of human myeloid and lymphoid lineages. Immune reconstitution in this model likely involves education of human T cells in the mouse thymus and dissemination of differentiated human lymphoid subsets into the peripheral blood and to multiple lymphoid tissues, including lymph nodes, spleen, and bone marrow (92). The DKO-hu-HSC mouse, along with the other humanized mouse models, has been used in studies of transmission (5, 21), pathogenesis (43), and viral inhibition (16, 21, 53, 88, 94).One important feature of HIV-1 infection is the diversification and evolution of the viral genome over the course of infection. Diversification occurs most prominently in the envelope (env) gene, which encodes the viral surface glycoprotein (Env). Env mediates viral entry into cells through attachment to the primary receptor CD4, which primes Env for engagement with a coreceptor, either CCR5 or CXCR4, triggering virion fusion with the cellular plasma membrane (54). HIV-1 infection is typically established by one or a few CCR5-tropic (R5) variants that give rise to an initially homogenous viral population, which then diversifies over the course of chronic infection (45, 84). Diversification of Env results from immune selective pressures (27), isolation in or adaptation to different cellular and anatomical compartments (20, 28, 33, 46, 51), and selection for altered CD4 affinity (72, 90, 95) and coreceptor tropism (26, 39). In many cases, during late-stage infection, variants emerge from the R5 virus population that are CXCR4 tropic (X4), an event that is often associated with accelerated CD4 T-cell loss and progression to AIDS (9, 18, 89). In an effort to determine if any of these aspects of HIV-1 evolution are exhibited in the humanized mouse model, we examined the extent of HIV-1 diversification and the types of evolutionary changes that occur in env in mice infected with CCR5-tropic HIV-1 for up to 44 weeks.Sampling of viral env variants from the peripheral blood plasma over the course of the infection revealed increasing diversity and divergence of the viral population at rates similar to those observed in natural infection. Mutations were identified that affected Env conformation and sensitivity to neutralizing antibodies, CXCR4 coreceptor use, and potential N-linked glycosylation sites. Other mutations potentially affecting the Env phenotype were identified in CD4 binding sites and CD4-induced epitopes. The patterns of substitutions indicated that certain sites were under selection, particularly in cases where the same substitution was identified in multiple mice.This study demonstrates the potential for studying HIV-1 evolution in the DKO-hu-HSC mouse model and also gives insight into the types of selective pressures driving HIV-1 env evolution in this host environment. These findings, while highlighting some of the limitations of this model, will help to inform its appropriate use for studying different aspects of HIV-1 infection, such as the evolutionary constraints placed on HIV-1 during natural infection and in the face of pharmacological and immunological inhibition.  相似文献   

4.
5.
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   

6.
CD4C/HIVnef transgenic (Tg) mice express Nef in CD4+ T cells and in the cells of the macrophage/monocyte/dendritic lineage, and they develop an AIDS-like disease similar to human AIDS. In these mice, Nef is constitutively expressed throughout life. To rule out the contribution of any developmental defects caused by early expression of Nef, we generated inducible human immunodeficiency virus type 1 (HIV-1) Nef Tg mice by using the tetracycline-inducible system. Faithful expression of the Nef transgene was induced in (CD4C/rtTA × TRE/HIVNef) or (CD4C/rtTA2S-M2 × TRE/HIVNef) double-Tg mice upon doxycycline (DOX) treatment in drinking water. Long-term treatment of these mice with DOX also led to loss, apoptosis, and activation of CD4+ T cells, this latter phenotype being observed even with low levels of Nef. These phenotypes could be transferred by bone marrow (BM) transplantation, indicating a hematopoietic cell autonomous effect. In addition, in mixed Tg:non-Tg BM chimeras, only Tg and not non-Tg CD4+ T cells exhibited an effector/memory phenotype in the absence of lymphopenia. Finally, the DOX-induced double-Tg mice developed nonlymphoid organ diseases similar to those of CD4C/HIVNef Tg mice and of humans infected with HIV-1. These results show for the first time that adult mice are susceptible to the detrimental action of Nef and that Nef-mediated T-cell activation can be independent of lymphopenia. These Tg mice represent a unique model which is likely to be instrumental for understanding the cellular and molecular pathways of Nef action as well as the main characteristics of immune reconstitution following DOX withdrawal.Small animal models able to express the entire human immunodeficiency virus (HIV) genome or selected HIV genes have provided useful information on the pathogenesis of AIDS and still represent important research tools toward this goal. Among these models, transgenic (Tg) mice containing intact copies of HIV DNA, defective provirus with the gag and pol genes deleted, or individual HIV-1 genes have been reported to develop various pathologies, some of which resemble those found in human AIDS (2, 3, 8, 9, 16, 17, 18, 24, 27, 29, 30, 38, 44, 45, 46, 49, 51, 52). The cell type context in which the HIV-1 transgene is expressed in these Tg mice appears to play an important role in determining the type of pathological lesions. Tg mice generated in our laboratory and expressing the entire coding sequence of HIV-1 (CD4C/HIVWT) or HIV-1 Nef alone (CD4C/HIVNef) in the relevant target cells of HIV-1, namely, CD4+ T cells, macrophages, and dendritic cells, develop pathologies very similar to those in human AIDS (17, 18). The AIDS-like disease of CD4C/HIVNef Tg mice is characterized by immunodeficiency, loss of CD4+ T cells, thymic atrophy, activation of T cells and pathologies in heart, lungs, and kidneys (18, 53). Similarly, expression of simian immunodeficiency virus (SIV) Nef in Tg mice under the control of the same promoter sequences (CD4C) results in an AIDS-like disease (42). These studies demonstrated that Nef plays an important role in the development of the AIDS-like disease induced by HIV-1 or SIV in Tg mice.Among the AIDS-like phenotypes of these models, the T-cell activation observed by a number of groups in Tg mice expressing Nef (3, 33, 44, 53) may be of special interest for its resemblance to that of humans or macaques infected with HIV-1 or SIV, respectively. HIV infection results in a state of chronic immune activation which correlates very closely with disease progression in humans (11, 14, 23). Similarly, SIV-infected macaques which develop AIDS show aberrant immune activation (35), while SIV-infected sooty mangabey monkeys, natural hosts of SIV, do not develop immunopathologies and do not show immune activation either (41). Various factors may contribute to this immune activation, including increased plasma lipopolysaccharide levels due to microbial translocation from the gut (4), impaired regulatory T cell function (32), or the action of the HIV-1 gene products themselves, such as Env gp120 and Nef (10, 12, 43). Consistent with this latter scenario, we reported that in CD4C/HIVNef Tg mice the extent of T-cell activation correlates with levels of Nef expression in CD4+ T cells, thus suggesting a direct involvement of Nef in this activation (53). In contrast, Koenen and coworkers reported that T-cell activation in CD2/Nef Tg mice is induced indirectly by lymphophenia (26). In that study, chimeric mice, which were generated from a mixture of non-Tg and Nef Tg bone marrow (BM) cells, were not lymphopenic, and the donor-derived Nef-expressing Tg T cells did not show an activated phenotype. However, the donor Nef Tg T cells constituted only 1 to 2% of peripheral T cells of these chimeric mice (26). Clearly, alternative experimental approaches are needed to study this phenotype in a more physiological context.In the previously described CD4C/HIVNef Tg mice (18), Nef expression begins early in life and is constitutively expressed throughout the life of the animal. The AIDS-like disease caused by this early expression of Nef best represents a model for pediatric AIDS. However, in these Tg mice, Nef may interfere with normal developmental processes and these latter defects may contribute to some of the phenotypes observed. To assess the effects of Nef in fully mature adult animals, and thus develop a model of adult AIDS, temporal regulation of Nef expression in adult mice using an inducible system is required.In the present study, we chose the tet-On (rtTA and rtTA2S-M2) system (13, 15, 25, 48) to induce expression of HIV-1 Nef in CD4+ T cells and cells of the macrophage/dendritic lineage of mice using the CD4C tissue-specific regulatory elements. These CD4C sequences were previously used to generate the constitutively Nef-expressing CD4C/HIVNef Tg mice (18). These inducible adult (TRE/HIVNef × CD4C/rtTA) and (TRE/HIVNef × CD4C/rtTA2S-M2) double-Tg (DTg) mice express Nef when treated with doxycycline (DOX) and develop an AIDS-like disease very similar to that seen in constitutively Nef-expressing CD4C/HIVNef Tg mice. We took advantage of this novel biological system to reassess the role of Nef in T-cell activation. Using a mixed chimera made with BM cells from these inducible Nef Tg mice and from non-Tg mice, we could document CD4+ T-cell activation only in donor-derived Nef-expressing Tg cells, but not in non-Tg cells, in the absence of lymphopenia. This result strongly suggests that this CD4+ T-cell activation phenotype is most likely driven by expression of Nef in these cells.  相似文献   

7.
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   

8.
9.
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   

10.
11.
12.
13.
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   

14.
The human scavenger receptor gp340 has been identified as a binding protein for the human immunodeficiency virus type 1 (HIV-1) envelope that is expressed on the cell surface of female genital tract epithelial cells. This interaction allows such epithelial cells to efficiently transmit infective virus to susceptible targets and maintain viral infectivity for several days. Within the context of vaginal transmission, HIV must first traverse a normally protective mucosa containing a cell barrier to reach the underlying T cells and dendritic cells, which propagate and spread the infection. The mechanism by which HIV-1 can bypass an otherwise healthy cellular barrier remains an important area of study. Here, we demonstrate that genital tract-derived cell lines and primary human endocervical tissue can support direct transcytosis of cell-free virus from the apical to basolateral surfaces. Further, this transport of virus can be blocked through the addition of antibodies or peptides that directly block the interaction of gp340 with the HIV-1 envelope, if added prior to viral pulsing on the apical side of the cell or tissue barrier. Our data support a role for the previously described heparan sulfate moieties in mediating this transcytosis but add gp340 as an important facilitator of HIV-1 transcytosis across genital tract tissue. This study demonstrates that HIV-1 actively traverses the protective barriers of the human genital tract and presents a second mechanism whereby gp340 can promote heterosexual transmission.Through correlative studies with macaques challenged with simian immunodeficiency virus (SIV), the initial targets of infection in nontraumatic vaginal exposure to human immunodeficiency virus type 1 (HIV-1) have been identified as subepithelial T cells and dendritic cells (DCs) (18, 23, 31, 36-38). While human transmission may differ from macaque transmission, the existing models of human transmission remain controversial. For the virus to successfully reach its CD4+ targets, HIV must first traverse the columnar mucosal epithelial cell barrier of the endocervix or uterus or the stratified squamous barrier of the vagina or ectocervix, whose normal functions include protection of underlying tissue from pathogens. This portion of the human innate immune defense system represents a significant impediment to transmission. Studies have placed the natural transmission rate of HIV per sexual act between 0.005 and 0.3% (17, 45). Breaks in the epithelial barrier caused by secondary infection with other sexual transmitted diseases or the normal physical trauma often associated with vaginal intercourse represent one potential means for viral exposure to submucosal cells and have been shown to significantly increase transmission (reviewed in reference 11). However, studies of nontraumatic exposure to SIV in macaques demonstrate that these disruptions are not necessary for successful transmission to healthy females. This disparity indicates that multiple mechanisms by which HIV-1 can pass through mucosal epithelium might exist in vivo. Identifying these mechanisms represents an important obstacle to understanding and ultimately preventing HIV transmission.Several host cellular receptors, including DC-specific intercellular adhesion molecule-grabbing integrin, galactosyl ceramide, mannose receptor, langerin, heparan sulfate proteoglycans (HSPGs), and chondroitin sulfate proteoglycans, have been identified that facilitate disease progression through binding of HIV virions without being required for fusion and infection (2, 3, 12, 14, 16, 25, 29, 30, 43, 46, 50). These host accessory proteins act predominately through glycosylation-based interactions between HIV envelope (Env) and the host cellular receptors. These different host accessory factors can lead to increased infectivity in cis and trans or can serve to concentrate and expose virus at sites relevant to furthering its spread within the body. The direct transcytosis of cell-free virus through primary genital epithelial cells and the human endometrial carcinoma cell line HEC1A has been described (7, 9); this is, in part, mediated by HSPGs (7). Within the HSPG family, the syndecans have been previously shown to facilitate trans infection of HIV in vitro through binding of a specific region of Env that is moderately conserved (7, 8). This report also demonstrates that while HSPGs mediate a portion of the viral transcytosis that occurs in these two cell types, a significant portion of the observed transport occurs through an HSPG-independent mechanism. Other host cell factors likely provide alternatives to HSPGs for HIV-1 to use in subverting the mucosal epithelial barrier.gp340 is a member of the scavenger receptor cysteine-rich (SRCR) family of innate immune receptors. Its numerous splice variants can be found as a secreted component of human saliva (34, 41, 42) and as a membrane-associated receptor in a large number of epithelial cell lineages (22, 32, 40). Its normal cellular function includes immune surveillance of bacteria (4-6, 44), interaction with influenza A virus (19, 20, 32, 51) and surfactant proteins in the lung (20, 22, 33), and facilitating epithelial cell regeneration at sites of cellular inflammation and damage (27, 32). The secreted form of gp340, salivary agglutinin (SAG), was identified as a component of saliva that inhibits HIV-1 transmission in the oral pharynx through a specific interaction with the viral envelope protein that serves to agglutinate the virus and target it for degradation (34, 35, 41). Interestingly, SAG was demonstrated to form a direct protein-protein interaction with HIV Env (53, 54). Later, a cell surface-associated variant of SAG called gp340 was characterized as a binding partner for HIV-1 in the female genital tract that could facilitate virus transmission to susceptible targets of infection (47) and as a macrophage-expressed enhancer of infection (10).  相似文献   

15.
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   

16.
17.
18.
Using FrCasE retrovirus-infected newborn mice as a model system, we have shown recently that a long-lasting antiviral immune response essential for healthy survival emerges after a short treatment with a neutralizing (667) IgG2a isotype monoclonal antibody (MAb). This suggested that the mobilization of adaptive immunity by administered MAbs is key for the success in the long term for the MAb-based passive immunotherapy of chronic viral infections. We have addressed here whether the anti-FrCasE protective endogenous immunity is the mere consequence of viral propagation blunting, which would simply give time to the immune system to react, and/or to actual immunomodulation by the MAb during the treatment. To this aim, we have compared viral replication, disease progression, and antiviral immune responses between different groups of infected mice: (i) mice treated with either the 667 MAb, its F(ab′)2 fragment, or an IgM (672) with epitopic specificity similar to that of 667 but displaying different effector functions, and (ii) mice receiving no treatment but infected with a low viral inoculum reproducing the initial viral expansion observed in their infected/667 MAb-treated counterparts. Our data show that the reduction of FrCasE propagation is insufficient on its own to induce protective immunity and support a direct immunomodulatory action of the 667 MAb. Interestingly, they also point to sequential actions of the administered MAb. In a first step, viral propagation is exclusively controlled by 667 neutralizing activity, and in a second one, this action is complemented by FcγR-binding-dependent mechanisms, which most likely combine infected cell cytolysis and the modulation of the antiviral endogenous immune response. Such complementary effects of administered MAbs must be taken into consideration for the improvement of future antiviral MAb-based immunotherapies.Although monoclonal antibodies (MAbs) principally have been considered for anticancer applications heretofore (62, 64), they now are increasingly being considered to treat severe acute and chronic viral infections (43, 63, 83). The best-studied antiviral MAbs are (i) pavalizumab, a humanized anti-respiratory syncytial virus (RSV) MAb approved by the FDA in 1998 for treating severe lower-respiratory-tract diseases in infants (45); (ii) several anti-human immunodeficiency virus (HIV) MAbs, which have been used in macaque preclinical infection models and in several human trials (4, 5, 19, 27-30, 32, 42, 50, 55, 57, 76-79); and (iii) a few anti-hepatitis C virus (HCV) MAbs, some of which currently are being tested in humans (9, 22, 40). However, other MAbs, some of them of human origin, also have been generated against other human viruses in recent years. Among them are antibodies against Ebola virus (75), West Nile virus (WNV) (48, 53, 54), cytomegalovirus (CMV) (11), avian and human influenza viruses (59, 60, 73, 74), severe acute respiratory syndrome coronavirus (SARS CoV) (81), hepatitis B virus (HBV) (31, 35), Hanta virus (80, 82), and Nipah virus (80, 82). These antiviral MAbs all have been selected on the basis of their neutralizing activity and the possibility that they interfere with the antiviral immune response of treated hosts, because their effector functions have been considered surprisingly little so far. Addressing this question in clinical settings currently is not possible for a variety of reasons that include ethical, technical, and cost concerns. Therefore, we have turned to the neonatal infection of mice by the lethal FrCasE retrovirus as a model system. This model allowed us to show that a very short immunotherapy by a neutralizing MAb of the IgG2a isotype (667 MAb) can permit, in addition to an immediate direct effect on the viral load, the mounting of a long-lasting endogenous antiviral immunity, which is essential for viral control and healthy survival (23-25). Because of the broad therapeutic perspectives opened by this observation, it now is essential to elucidate the molecular and cellular mechanisms underlying this effect.FrCasE is a simple chimeric mouse retrovirus in which the env gene of the leukemogenic Friend murine leukemia virus (F-MuLV) was replaced by that of the neurodegeneration-inducing CasBr retrovirus (58). When 5 × 104 infectious particles are inoculated into newborn mice under the age of 5 to 6 days, FrCasE can enter the central nervous system (CNS) and induces a neurodegeneration fatal within 1 to 2 months with 100% incidence (15, 23, 41, 58). However, upon infection at a later time, FrCasE can no longer enter the CNS. Instead, it replicates only in the periphery and gives rise to a fatal erythroleukemia preceded by spleen enlargement and a dramatic drop of the hematocrit. Erythroleukemia incidence and incubation period, however, are variable, depending on the inoculum and the date of infection (46).667 is an IgG2a/κ (44) directed to the main viral receptor-binding site of CasBr Env (16). It displays both in vitro (44) and in vivo (56) neutralizing activities. When rapidly (<2 days) administered for a few days to neonatally FrCasE-infected pups, viral propagation is rapidly blunted, which prevents virus entry in the brain and subsequent neurodegeneration (23). Moreover, all 667-treated mice develop a strong, long-lasting antiviral immune response, which is necessary for them to survive healthy and with no sign of neurodegeneration or of erythroleukemia (23-25) and to resist viral challenges carried out as long as 14 months after first infection (23). Protective antiviral immunity is of a typical TH1 type with humoral and cytotoxic T-cell (CTL) contributions. The anti-FrCasE humoral contribution is high, sustained, and principally of the IgG2a type with both in vitro neutralization- and complement-dependent cytolysis activities (23). Interestingly, it shows typical secondary response characteristics in viral challenge experiments (23), and anti-FrCasE antibodies are transmitted transplacentally and through breastfeeding by mothers to children, where they manifest the same properties as those of 667 in the perinatal infection setting, i.e., they prevent mice from developing neurodegeneration and permit the induction of an endogenous protective antiviral immune response (25). Finally, the CTL response directed to infected cells was shown to be necessary for the protection of FrCasE-infected, 667-treated mice, as the depletion of CD8+ T cells leads to death by retrovirally induced erythroleukemia (24).At this stage, an important issue is the clarification of whether the anti-FrCasE protective immunity seen in 667 MAb-treated mice is due to actual immunomodulation by the MAb owing to its effector function(s) and/or is the consequence of viral propagation blunting, which would prevent the immune system from being overwhelmed by an excess of antigen and, hence, would give it time to react optimally. To address these two nonexclusive possibilities, we compared here viral propagations, health statuses, and endogenous immune responses in four groups of mice. The first three groups were mice neonatally infected under standard conditions (5 × 104 infectious particles) and treated with either the natural 667 MAb, the antibody effector function-lacking F(ab′)2 fragment of 667, or a neutralizing IgM (672) with effector functions inherently different from those of 667. The last group consisted of mice neonatally infected with a low FrCasE inoculum but not subjected to immunotherapy, which is a condition permitting early viral propagation kinetics similar to those of animals infected and 667 MAb treated under standard conditions. Taken together, our data indicate that the drastic reduction of viral propagation shortly after infection is not sufficient for the induction of protective adaptive immunity and, thereby, point to an immunomodulatory action of 667. Interestingly, they also point to two sequential actions of the administered MAb. In the immediate postinfection period, viral spread is controlled exclusively by 667 neutralizing activity, and later it involves the cytolysis of infected cells owing to FcγR-binding-dependent mechanisms. Finally, our work shows that not all antibody isotypes are equally efficient at protecting infected mice and favoring the mounting of protective immunity, as 672 IgM immunotherapy-treated animals died of erythroleukemia.  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