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 78 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.
The role of CD4+ helper T cells in modulating the acquired immune response to herpes simplex virus type 1 (HSV-1) remains ill defined; in particular, it is unclear whether CD4+ T cells are needed for the generation of the protective HSV-1-specific CD8+-T-cell response. This study examined the contribution of CD4+ T cells in the generation of the primary CD8+-T-cell responses following acute infection with HSV-1. The results demonstrate that the CD8+-T-cell response generated in the draining lymph nodes of CD4+-T-cell-depleted C57BL/6 mice and B6-MHC-II−/− mice is quantitatively and qualitatively distinct from the CD8+ T cells generated in normal C57BL/6 mice. Phenotypic analyses show that virus-specific CD8+ T cells express comparable levels of the activation marker CD44 in mice lacking CD4+ T cells and normal mice. In contrast, CD8+ T cells generated in the absence of CD4+ T cells express the interleukin 2 receptor α-chain (CD25) at lower levels. Importantly, the CD8+ T cells in the CD4+-T-cell-deficient environment are functionally active with respect to the expression of cytolytic activity in vivo but exhibit a diminished capacity to produce gamma interferon and tumor necrosis factor alpha. Furthermore, the primary expansion of HSV-1-specific CD8+ T cells is diminished in the absence of CD4+-T-cell help. These results suggest that CD4+-T-cell help is essential for the generation of fully functional CD8+ T cells during the primary response to HSV-1 infection.Infection due to herpes simplex virus type 1 (HSV-1) results in a wide spectrum of clinical presentations depending on the host''s age, the host''s immune status, and the route of inoculation (47). HSV-1 typically causes mild and self-limited lesions on the orofacial areas or genital sites. However, the disease can be life-threatening, as in the case of neonatal and central nervous system infections (18). The host''s immune responses, particularly CD8+ T cells, play an important role in determining the outcome of HSV infections in both the natural human host (18, 19, 28) and experimental murine models (11, 43). Immunodepletion and adoptive transfer studies have demonstrated the role of CD8+ T cells in reducing viral replication, resolving cutaneous disease, and providing overall protection upon rechallenge (6, 25, 26). CD8+ T cells play a particularly important role in preventing infection of the peripheral nervous system (PNS) and the reactivation of latent virus from neurons in the sensory ganglia of infected mice (21, 24, 36). The mechanisms that CD8+ T cells employ include gamma interferon (IFN-γ) production and functions associated with cytolytic granule content at the sites of primary infection (23, 31, 38). In the PNS of infected mice, the mechanisms primarily involve IFN-γ secretion (16, 20, 29), particularly against infected neurons expressing surface Qa-1 (41). Histopathological evidence from HSV-1-infected human ganglion sections show a large CD8+-T-cell infiltrate and the presence of inflammatory cytokines, suggesting that the presence of activated, effector memory cells within the PNS is important for maintaining HSV-1 latency in the natural human host (10, 42).The generation of a robust CD8+-T-cell response is essential for the control of various infectious pathogens. Some studies suggest that a brief interaction with antigen-presenting cells (APCs) is sufficient for CD8+-T-cell activation and expansion into functional effectors (44). However, the magnitude and quality of the overall CD8+-T-cell response generated may be dependent on additional factors (49). Recent evidence suggests that CD4+ T cells facilitate the activation and development of CD8+-T-cell responses either directly through the provision of cytokines or indirectly by the conditioning of dendritic cells (DC) (8, 48, 51). Those studies suggested that the latter mechanism is the dominant pathway, wherein CD4+ T cells assist CD8+-T-cell priming via the engagement of CD40 ligand (CD154) on CD4+ T cells and CD40 expressed on DC (4, 30, 33). This interaction results in the activation and maturation of DC, making them competent to stimulate antigen-specific CD8+-T-cell responses (35, 37).The requirement for CD4+-T-cell help in the generation of primary and secondary CD8+-T-cell responses to antigen varies. Primary CD8+-T-cell responses to infectious pathogens, such as Listeria monocytogenes, lymphocytic choriomeningitis virus (LCMV), influenza virus, and vaccinia virus, can be mounted effectively independently of CD4+-T-cell help (3, 12, 22, 34). In contrast, primary CD8+-T-cell responses to nonmicrobial antigens display an absolute dependence on CD4+-T-cell help (4, 5, 30, 33, 46). This observed difference in the requirement for CD4+-T-cell help may ultimately be a product of the initial inflammatory stimulus generated following immunization (49). Microbial antigens trigger an inflammatory response that can lead to the direct activation and priming of APCs, such as DC, thereby bypassing the need for CD4+-T-cell help. Nonmicrobial antigens, however, trigger an attenuated inflammatory response that does not directly activate and prime DCs. In the absence of this inflammation, CD4+ T cells are thought to condition and license DC functions through CD154/CD40 interactions, which leads to the subsequent activation of antigen-specific CD8+-T-cell responses (5, 49). Even in the case of pathogens where primary CD8+-T-cell responses were independent of CD4+-T-cell help, the secondary responses to these pathogens were found to be defective in the absence of CD4+-T-cell help (3, 12, 34, 40).The requirement for CD4+-T-cell help in priming CD8+-T-cell responses against HSV-1 infection is not well defined. Earlier studies with HSV-1 suggested that CD4+ T cells play an important role in the generation of primary CD8+-T-cell responses, detected in vitro, to acute infection with HSV-1 (14), principally through the provision of interleukin 2 (IL-2) for optimal CD8+-T-cell differentiation and proliferation. Subsequent studies, utilizing an in vivo approach, indicated that CD4+ T cells were not required for CD8+-T-cell-mediated cytolytic function (23). CD4+ T cells are thought to provide help by conditioning DC in a cognate, antigen-specific manner, thereby making them competent to stimulate HSV-1-specific CD8+-T-cell responses (37). By contrast, findings from other studies show that CD4+-T-cell-depleted mice were able to fully recover from acute infection with HSV-1 (38). These studies imply that the absence of CD4+ T cells does not prevent priming of CD8+ T cells in vivo.Studies from this laboratory have identified two distinct HSV-1-specific CD8+-T-cell subpopulations generated during the primary response, based upon the ability to synthesize IFN-γ following antigenic stimulation in vitro (1). To better understand the need for CD4+-T-cell help, we examined the functional characteristics and phenotypes of these CD8+-T-cell populations generated during a primary response to acute infection with HSV-1 in mice lacking CD4+ T cells. Our findings show that primary CD8+-T-cell responses to HSV-1 are compromised in the absence of CD4+-T-cell help. Specifically, the HSV-1 gB-specific CD8+ T cells produced in the absence of CD4+ T cells were found to be active with regard to cytolysis in vivo but were functionally impaired in the production of IFN-γ and TNF-α compared with intact C57BL/6 mice. Virus-specific CD8+ T cells were also reduced in number in CD4-depleted mice and in B6 mice lacking major histocompatibility complex (MHC) class II expression (B6-MHC-II−/−) compared to wild-type (WT) mice. In addition, our data showed higher virus burdens in the infectious tissues obtained from mice lacking CD4+ T cells than in those from intact mice. Collectively, these findings demonstrate that CD4+-T-cell help is essential for the generation of primary CD8+-T-cell responses following acute cutaneous infection with HSV-1.  相似文献   

3.
4.
Autophagy is an important component of host innate and adaptive immunity to viruses. It is critical for the degradation of intracellular pathogens and for promoting antigen presentation. Herpes simplex virus type 1 (HSV-1) infection induces an autophagy response, but this response is antagonized by the HSV-1 neurovirulence gene product, ICP34.5. This is due, in part, to its interaction with the essential autophagy protein Beclin 1 (Atg6) via the Beclin-binding domain (BBD) of ICP34.5. Using a recombinant virus lacking the BBD, we examined pathogenesis and immune responses using mouse models of infection. The BBD-deficient virus (Δ68H) replicated equivalently to its marker-rescued counterpart (Δ68HR) at early times but was cleared more rapidly than Δ68HR from all tissues at late times following corneal infection. In addition, the infection of the cornea with Δ68H induced less ocular disease than Δ68HR. These results suggested that Δ68H was attenuated due to its failure to control adaptive rather than innate immunity. In support of this idea, Δ68H stimulated a significantly stronger CD4+ T-cell-mediated delayed-type hypersensitivity response and resulted in significantly more production of gamma interferon and interleukin-2 from HSV-specific CD4+ T cells than Δ68HR. Taken together, these data suggest a role for the BBD of ICP34.5 in precluding autophagy-mediated class II antigen presentation, thereby enhancing the virulence and pathogenesis of HSV-1.Autophagy is a conserved cellular pathway that eliminates defective proteins and organelles, prevents abnormal protein aggregate accumulation, and removes intracellular pathogens (11, 22, 32, 56). This process begins with the formation and elongation of a double membrane that fuses to form an autophagosome. The cytoplasmic contents are nonspecifically sequestered inside the autophagosome and then are degraded once the autophagosome fuses with the lysosome. Autophagy is upregulated during starvation, growth factor withdrawal, hypoxia, and infection (10). Following metabolic stress, autophagy can generate metabolic precursors that can be recycled for the de novo synthesis of proteins. The autophagic pathway has important roles in development, immune defense, apoptosis, tumor suppression, and the prevention of neuronal degeneration (reviewed in references 21 and 31).Autophagy is not limited to the degradation of self proteins; it also can engulf and break down invading microorganisms in a process termed xenophagy (26). Xenophagy can limit the replication of pathogens (3, 4, 29, 34, 42), but some infectious agents can exploit autophagy to enhance their replication (2, 18, 38). There are pathogens that actively inhibit autophagy through interaction with the essential autophagy-promoting protein, Beclin 1 (24, 35). Beclin 1 is the mammalian homolog of yeast Atg6 and is required for the formation of the autophagosome membrane through its interaction with VPS34, a class III phosphatidylinositol 3-kinase (19, 27). Autophagy/xenophagy is also an important process for the adaptive immune response to infection through the delivery of antigens for major histocompatibility complex class I and II (MHC-I and -II) presentation (7, 9, 12, 36, 45). The inhibition of autophagy by pathogens therefore would serve to block CD4+ and CD8+ cell responses and allow pathogens to remain underrecognized by the adaptive immune response.The interferon (IFN)-inducible double-strand RNA-inducible protein kinase (PKR) pathway is required for virus- and starvation-induced autophagy (50). The PKR-mediated induction of autophagy requires the phosphorylation of the translation initiation factor eIF2α (50). Herpes simplex virus type 1 (HSV-1) inhibits autophagy through at least two domains and the activities of the late protein ICP34.5, its C-terminal mediation of dephosphorylation of eIF2α, and its N-terminal binding to Beclin 1 (5, 16, 35). HSV-1 strains lacking ICP34.5 show significant attenuation in vivo, increased sensitivity to IFN (33), an inability to counteract the PKR-induced phosphorylation of eIF2α, and the induction of the generalized shutoff of protein synthesis in infected cells. These activities originally were ascribed solely to ICP34.5''s ability to recruit PP1α and redirect its activity to dephosphorylate eIF2α to counteract the general shutoff of protein synthesis mediated by PKR (17). This function is mediated by the C-terminal domain of ICP34.5 that contains homology to the growth arrest and DNA damage 34 (GADD34) gene (6). Independently of the role of ICP34.5 in countering the PKR-induced antiviral state, viruses lacking ICP34.5 also exhibit altered patterns of autophagy. This manifests as increased long-lived protein degradation, the increased formation of autophagosomes, increased autophagic vacuole volume density, and the enhanced xenophagic degradation of virions (35, 50, 51). Such alterations in the autophagy pathway now can be ascribed to ICP34.5''s ability to bind Beclin 1 in addition to its mediation of eIF2α dephosphorylation (35). The deletion of the Beclin 1-binding domain (BBD) of ICP34.5 renders HSV-1 less able to regulate autophagosome formation, and viruses lacking this domain are neuroattenuated (35).To determine the impact of Beclin 1-ICP34.5 interactions on the pathogenesis of HSV-1 during infection, we examined the ability of BBD-deficient virus to replicate, cause disease, and stimulate an immune response in mice. We determined that mice infected with HSV-1 lacking the ICP34.5 BBD were able to clear virus more efficiently than mice infected with wild-type virus. We observed the significantly enhanced stimulation of CD4+ T cells by the virus lacking the BBD compared to that of mice infected with wild-type virus. These data suggest that Beclin 1 binding and the inhibition of autophagy by ICP34.5 are important for HSV-1 pathogenesis through its ability to suppress autophagy and to dampen the activation of CD4+ T cells.  相似文献   

5.
6.
7.
8.
The kinetics of CD8+ T cell epitope presentation contribute to the antiviral efficacy of these cells yet remain poorly defined. Here, we demonstrate presentation of virion-derived Vpr peptide epitopes early after viral penetration and prior to presentation of Vif-derived epitopes, which required de novo Vif synthesis. Two Rev epitopes exhibited differential presentation kinetics, with one Rev epitope presented within 1 h of infection. We also demonstrate that cytolytic activity mirrors the recognition kinetics of infected cells. These studies show for the first time that Vpr- and Rev-specific CD8+ T cells recognize and kill simian immunodeficiency virus (SIV)-infected CD4+ T cells early after SIV infection.The antiviral activity of AIDS virus-specific CD8+ T cells is well documented in both in vivo (1, 4, 21) and in vitro (8, 24, 29) studies. Accordingly, human immunodeficiency virus (HIV) vaccine modalities that focus on engendering antiviral CD8+ T cells are being developed (13, 26, 28). Ideally, a CD8+ T cell-based vaccine would stimulate responses against epitopes that are presented by major histocompatibility complex class I (MHC-I) molecules early after infection of a target cell. However, successful selection of antigenic sequences for a CD8+ T cell-based vaccine has been frustrated in part by an incomplete understanding of the properties of effective CD8+ T cell responses (25).  相似文献   

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

14.
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   

15.
An N-carbamoyl-β-alanine amidohydrolase of industrial interest from Agrobacterium tumefaciens C58 (βcarAt) has been characterized. βcarAt is most active at 30°C and pH 8.0 with N-carbamoyl-β-alanine as a substrate. The purified enzyme is completely inactivated by the metal-chelating agent 8-hydroxyquinoline-5-sulfonic acid (HQSA), and activity is restored by the addition of divalent metal ions, such as Mn2+, Ni2+, and Co2+. The native enzyme is a homodimer with a molecular mass of 90 kDa from pH 5.5 to 9.0. The enzyme has a broad substrate spectrum and hydrolyzes nonsubstituted N-carbamoyl-α-, -β-, -γ-, and -δ-amino acids, with the greatest catalytic efficiency for N-carbamoyl-β-alanine. βcarAt also recognizes substrate analogues substituted with sulfonic and phosphonic acid groups to produce the β-amino acids taurine and ciliatine, respectively. βcarAt is able to produce monosubstituted β2- and β3-amino acids, showing better catalytic efficiency (kcat/Km) for the production of the former. For both types of monosubstituted substrates, the enzyme hydrolyzes N-carbamoyl-β-amino acids with a short aliphatic side chain better than those with aromatic rings. These properties make βcarAt an outstanding candidate for application in the biotechnology industry.N-Carbamoyl-β-alanine amidohydrolase (NCβAA) (EC 3.5.1.6), also known as β-alanine synthase or β-ureidopropionase, catalyzes the third and final step of reductive pyrimidine degradation. In this reaction, N-carbamoyl-β-alanine or N-carbamoyl-β-aminoisobutyric acid is irreversibly hydrolyzed to CO2, NH3, and β-alanine or β-aminoisobutyric acid, respectively (43). Eukaryotic NCβAAs have been purified from several sources (10, 25, 33, 39, 42, 44). Nevertheless, only two prokaryotic NCβAAs, belonging to the Clostridium and Pseudomonas genera (4, 29), have been purified to date, although this activity has been inferred for several microorganisms due to the appearance of the reductive pathway of pyrimidine degradation (38, 45). Pseudomonas NCβAA is also able to hydrolyze l-N-carbamoyl-α-amino acids, and indeed, this activity is widespread in the bacterial kingdom (3, 23, 26, 46).β-Amino acids have unique pharmacological properties, and their utility as building blocks of β-peptides, pharmaceutical compounds, and natural products is of growing interest (14). β-Alanine, a natural β-amino acid, is a precursor of coenzyme A and pantothenic acid in bacteria and fungi (vitamin B5) (7). β-Alanine is widely distributed in the central nervous systems of vertebrates and is a structural analogue of γ-amino-n-butyric acid and glycine, major inhibitory neurotransmitters, suggesting that it may be involved in synaptic transmissions (20). Another important natural β-amino acid is taurine (2-aminoethanesulfonic acid), which plays an important role in several essential processes, such as membrane stabilization, osmoregulation, glucose metabolism, antioxidation, and development of the central nervous system and the retina (9, 28, 33). 2-Aminoethylphosphonate, the most common naturally occurring phosphonate, also known as ciliatine, is an important precursor used in the biosynthesis of phosphonolipids, phosphonoproteins, and phosphonoglycans (5). β-Homoalanine (β-aminobutyric acid) has been used successfully for the design of nonnatural ligands for therapeutic application against autoimmune diseases such as rheumatoid arthritis, multiple sclerosis, or autoimmune uveitis (30). Substituted β-amino acids can be denominated β2, β3, and β2,3, depending on the position of the side chain(s) (R) on the amino acid skeleton (18). β2-Amino acids are not yet as readily available as their β3-counterparts, as they must be prepared using multistep procedures (17).We decided to characterize NCβAA (β-carbamoylase) from Agrobacterium tumefaciens C58 (βcarAt) after showing that some dihydropyrimidinases belonging to the Arthrobacter and Sinorhizobium genera are able to hydrolyze different 5- or 6-substituted dihydrouracils to the corresponding N-carbamoyl-β-amino acids (18, 22). If βcarAt could decarbamoylate the reaction products of dihydrouracils, different β-amino acids would be obtained enzymatically in the same way that α-amino acids are produced via the hydantoinase process (6, 21). We therefore describe the physical, biochemical, kinetic, and substrate specificity properties of recombinant βcarAt.  相似文献   

16.
17.
Highly pathogenic influenza A viruses cause acute severe pneumonia to which the occurrence of “cytokine storm” has been proposed to contribute. Here we show that interleukin-15 (IL-15) knockout (KO) mice exhibited reduced mortality after infection with influenza virus A/FM/1/47 (H1N1, a mouse-adapted strain) albeit the viral titers of these mice showed no difference from those of control mice. There were significantly fewer antigen-specific CD44+ CD8+ T cells in the lungs of infected IL-15 KO mice, and adoptive transfer of the CD8+ T cells caused reduced survival of IL-15 KO mice following influenza virus infection. Mice deficient in β2-microglobulin by gene targeting and those depleted of CD8+ T cells by in vivo administration of anti-CD8 monoclonal antibody displayed a reduced mortality rate after infection. These results indicate that IL-15-dependent CD8+ T cells are at least partly responsible for the pathogenesis of acute pneumonia caused by influenza A virus.Highly pathogenic influenza A viruses cause acute severe pneumonia that results in high morbidity and significant mortality (11, 12, 24, 26). Elevated levels of serum cytokines and chemokines accompany these clinical manifestations, and the possibility that this “cytokine storm” contributes to increased severity of the disease caused by avian H5N1 virus and by other strains of influenza A virus has been proposed (10, 21, 33). In fact, CCR2-deficient mice [CCR2 is chemokine (C-C motif) receptor 2] were protected from early pathological manifestations despite higher pulmonary titers of the influenza virus A/PR/8/34 (H1N1) strain (7). Tumor necrosis factor receptor 1 (TNFR-1)-deficient mice exhibited significantly reduced morbidity following challenge with H5N1 virus (31). Other cytokines or chemokines have also been investigated (8, 28, 34, 35, 38). Thus, at least some of the elevated proinflammatory cytokines may contribute to the pathogenesis of influenza A virus.Interleukin-15 (IL-15) is a pleiotropic cytokine involved in both innate and adaptive immune responses (20, 36). IL-15 utilizes the β-chain of the IL-2 receptor (IL-2R) (CD122) and the common cytokine receptor γ-chain (CD132) for signal transduction in lymphocytes and therefore shares many biological properties with IL-2 (3). Memory CD8+ T cells, natural killer (NK) cells, NKT cells, and intraepithelial lymphocyte (IEL) T cells (15, 23, 42) decrease in mice with defective IL-15 signaling, indicating the importance of IL-15 in their development and/or maintenance. IL-15 regulates not only the number of memory CD8+ T cells but also activation of their functions, including gamma interferon (IFN-γ) production and cytotoxic activity (40), which are important to target the virus (9). Therefore, it is possible that we may be able to use IL-15 as an immune-enhancing molecular adjuvant in vaccines for protection against various pathogens, including influenza A virus (37).In the present study, we demonstrate that IL-15 knockout (KO) mice exhibited high resistance against infection with mouse-adapted influenza virus A/FM/1/47 (H1N1) strain. We show for the first time that IL-15-dependent CD8+ T cells are at least partly responsible for the pathogenesis of acute pneumonia caused by influenza A virus. In addition, our observations are important in the light of recent research into the use of IL-15 as an adjuvant for vaccination.  相似文献   

18.
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