首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 46 毫秒
1.
2.
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   

3.
Like human immunodeficiency virus type 1 (HIV-1), most simian immunodeficiency virus (SIV) strains use CCR5 to establish infection. However, while HIV-1 can acquire the ability to use CXCR4, SIVs that utilize CXCR4 have rarely been reported. To explore possible barriers against SIV coreceptor switching, we derived an R5X4 variant, termed 239-ST1, from the R5 clone SIVmac239 by serially passaging virus in CD4+ CXCR4+ CCR5 SupT1 cells. A 239-ST1 env clone, designated 239-ST1.2-32, used CXCR4 and CCR5 in cell-cell fusion and reporter virus infection assays and conferred the ability for rapid, cytopathic infection of SupT1 cells to SIVmac239. Viral replication was inhibitable by the CXCR4-specific antagonist AMD3100, and replication was abrogated in a novel CXCR4 SupT1 line. Surprisingly, parental SIVmac239 exhibited low-level replication in SupT1 cells that was not observed in CXCR4 SupT1 cells. Only two mutations in the 239-ST1.2-32 Env, K47E in the C1 domain and L328W in the V3 loop, were required for CXCR4 use in cell-cell fusion assays, although two other V3 changes, N316K and I324M, improved CXCR4 use in infection assays. An Env cytoplasmic tail truncation, acquired during propagation of 239-ST1 in SupT1 cells, was not required. Compared with SIVmac239, 239-ST1.2-32 was more sensitive to neutralization by five of seven serum and plasma samples from SIVmac239-infected rhesus macaques and was approximately 50-fold more sensitive to soluble CD4. Thus, SIVmac239 can acquire the ability to use CXCR4 with high efficiency, but the changes required for this phenotype may be distinct from those for HIV-1 CXCR4 use. This finding, along with the increased neutralization sensitivity of this CXCR4-using SIV, suggests a mechanism that could select strongly against this phenotype in vivo.Simian immunodeficiency viruses (SIVs) share many structural and biological features with human immunodeficiency virus (HIV), including target cell entry via interactions of the viral envelope glycoprotein (Env) with CD4 and a chemokine coreceptor. For HIV, the most important coreceptors in vivo are CCR5 (2, 13, 19, 21, 22) and CXCR4 (30). HIV type 1 (HIV-1) strains that use only CCR5 (R5 viruses) predominate during the early stages of infection and are critical for transmission (84, 90), as evidenced by the finding that individuals lacking a functional CCR5 protein due to a homozygous 32-bp deletion in the CCR5 gene (ccr532) are largely resistant to HIV-1 infection (16, 54, 82). Although R5 viruses generally persist in late-stage disease, viruses that can use CXCR4, either exclusively (X4 viruses) or in addition to CCR5 (R5X4 viruses), emerge in approximately 50% of subtype B-infected individuals (15, 43). This coreceptor switch is associated with a more rapid decline in peripheral blood CD4+ T cells and a faster progression to AIDS (15, 43, 77), although it is unclear if CXCR4-using viruses are a cause or a consequence of progressing immunodeficiency. Like HIV, the vast majority of SIVs use CCR5 to establish infection (11, 12, 45). However, although CXCR4-using SIVs have been reported (47, 52, 65, 68, 69), their occurrence is rare, especially in models of pathogenic infection, where only one CXCR4-using SIV has been identified (17, 60, 71).This paucity of CXCR4-using SIVs is surprising for several reasons. First, SIV Envs tend to be more promiscuous than HIV-1 Envs and frequently use alternative coreceptors in addition to CCR5, including GPR1, GPR15, CXCR6, and CCR8 (20, 27, 29, 80, 81, 92) but not CXCR4. Second, HIV-2, which is more closely related to SIVmac than to HIV-1 (56, 57), commonly uses CXCR4 in vitro and in vivo (3, 28, 33, 58, 59, 67). Third, rhesus CXCR4 is ∼98% identical to human CXCR4 in amino acid sequence and can function as a coreceptor for HIV-1 in vitro (12). Finally, chimeric simian-human immunodeficiency viruses (SHIVs) that contain X4 HIV Envs on an SIV core can replicate to high levels in vivo and cause disease in rhesus macaques (39, 86). Moreover, it was recently shown that coreceptor switching can occur in rhesus macaques infected with an R5 SHIV (35). Thus, there does not appear to be any block per se against the use of rhesus CXCR4 as an entry coreceptor either in vitro or in vivo, suggesting that SIV is less capable of adapting to use CXCR4 and/or that mutations required for CXCR4 utilization may lead to a virus that is less fit and/or more susceptible to immune control in this host.For HIV-1, the Env determinants for CXCR4 use have been well documented and often involve the acquisition of positively charged amino acids in the V3 loop (18, 32, 87), particularly at positions 11, 24, and 25 (6, 18, 31, 32, 38, 75). Although the SIVmac239 V3 loop is a critical determinant for Env-coreceptor interactions (44, 63, 72), attempts to create an X4 SIVmac239 by introducing positively charged residues into the V3 loop (63) or by inserting a V3 loop from X4 HIV-1 (44) have been unsuccessful. SIVmac155T3, the only CXCR4-using variant of SIVmac that has been identified to date, was isolated from a rhesus macaque with advanced disease and contains additional positively charged residues in V3, although the determinants for CXCR4 use have not been determined (60, 71).Given questions concerning the possible determinants for and/or barriers to coreceptor switching in SIV, we sought to derive a CXCR4-using variant of the well-characterized pathogenic R5 SIV clone SIVmac239. Here we show that SIVmac239 could indeed acquire CXCR4 utilization when it was adapted in vitro for high-efficiency replication in the CXCR4+ CCR5 human SupT1 cell line. An env clone from this virus could use CXCR4 in cell-cell fusion and reporter virus infection assays and conferred CXCR4 tropism to a replication-competent SIV. Although V3 mutations were important for CXCR4 use, an L328W change at the V3 crown rather than the acquisition of positively charged residues was required, as was an unusual K47E mutation in the conserved C1 domain of gp120. These changes also caused the highly neutralization-resistant SIVmac239 strain to become more neutralization sensitive to sera and plasmas from SIVmac239-infected animals, and particularly to soluble CD4. These results indicate that mutations distinct from those typically seen for HIV-1 may be required for SIVmac to gain CXCR4 utilization and suggest that these changes render this virus more susceptible to humoral immune control. Collectively, our findings indicate that there are likely to be strong viral and host selection pressures against CXCR4 use that may contribute to the paucity of X4 coreceptor switching for SIVmac in vivo.  相似文献   

4.
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   

5.
6.
During the course of infection, transmitted HIV-1 isolates that initially use CCR5 can acquire the ability to use CXCR4, which is associated with an accelerated progression to AIDS. Although this coreceptor switch is often associated with mutations in the stem of the viral envelope (Env) V3 loop, domains outside V3 can also play a role, and the underlying mechanisms and structural basis for how X4 tropism is acquired remain unknown. In this study we used a V3 truncated R5-tropic Env as a starting point to derive two X4-tropic Envs, termed ΔV3-X4A.c5 and ΔV3-X4B.c7, which took distinct molecular pathways for this change. The ΔV3-X4A.c5 Env clone acquired a 7-amino-acid insertion in V3 that included three positively charged residues, reestablishing an interaction with the CXCR4 extracellular loops (ECLs) and rendering it highly susceptible to the CXCR4 antagonist AMD3100. In contrast, the ΔV3-X4B.c7 Env maintained the V3 truncation but acquired mutations outside V3 that were critical for X4 tropism. In contrast to ΔV3-X4A.c5, ΔV3-X4B.c7 showed increased dependence on the CXCR4 N terminus (NT) and was completely resistant to AMD3100. These results indicate that HIV-1 X4 coreceptor switching can involve (i) V3 loop mutations that establish interactions with the CXCR4 ECLs, and/or (ii) mutations outside V3 that enhance interactions with the CXCR4 NT. The cooperative contributions of CXCR4 NT and ECL interactions with gp120 in acquiring X4 tropism likely impart flexibility on pathways for viral evolution and suggest novel approaches to isolate these interactions for drug discovery.For human immunodeficiency virus type I (HIV-1) to enter a target cell, the gp120 subunit of the viral envelope glycoprotein (Env) must engage CD4 and a coreceptor on the cell surface. Although numerous coreceptors have been identified in vitro, the two most important coreceptors in vivo are the CCR5 (3, 11, 19, 22, 24) and CXCR4 (27) chemokine receptors. HIV-1 variants that can use only CCR5 (R5 viruses) are critical for HIV-1 transmission and predominate during the early stages of infection (86, 90). The importance of CCR5 for HIV-1 transmission is underscored by the fact that individuals bearing a homozygous 32-bp deletion in the CCR5 gene (ccr5-Δ32) are largely resistant to HIV-1 infection (15, 49, 84). Although R5 viruses typically persist into late disease stages, viruses that can use CXCR4, either alone (X4 viruses) or in addition to CCR5 (R5X4 viruses), emerge in approximately 50% of individuals infected with subtype B or D viruses (12, 39, 44). Although not required for disease progression, the appearance of X4 and/or R5X4 viruses is associated with a more rapid depletion of CD4+ cells in peripheral blood and faster progression to AIDS (12, 44, 77, 86). However, it remains unclear whether these viruses are a cause or a consequence of accelerated CD4+ T cell decline (57). The emergence of CXCR4-using viruses has also complicated the use of small-molecule CCR5 antagonists as anti-HIV-therapeutics as these compounds can select for the outgrowth of X4 or R5X4 escape variants (93).Following triggering by CD4, gp120 binds to a coreceptor via two principal interactions: (i) the bridging sheet, a four-stranded antiparallel beta sheet that connects the inner and outer domains of gp120, together with the base of the V3 loop, engages the coreceptor N terminus (NT); and (ii) more distal regions of V3 interact with the coreceptor extracellular loops (ECLs) (13, 14, 36-38, 43, 59, 60, 78, 79, 88). Although both the NT and ECL interactions are important for coreceptor binding and entry, their relative contributions vary among different HIV-1 strains (23). For example, V3 interactions with the ECLs, particularly ECL2, serve a dominant role in CXCR4 utilization (7, 21, 50, 63, 72), while R5 viruses exhibit a more variable use of CCR5 domains, with the NT interaction being particularly important (4, 6, 20, 67, 83). Although V3 is the primary determinant of coreceptor preference (34), it is unclear how specificity for CCR5 and/or CXCR4 is determined, and, in particular, it is unknown how X4 tropism is acquired. Several reports have shown that the emergence of X4 tropism correlates with the acquisition of positively charged residues in the V3 stem (17, 29, 87), particularly at positions 11, 24, and 25 (8, 17, 28, 29, 42, 75), raising the possibility that these mutations directly or indirectly mediate interactions with negatively charged residues in the CXCR4 ECLs. However, Env domains outside V3, including V1/V2 (9, 32, 45, 46, 61, 64, 65, 80, 95) and even gp41 (40), can also contribute to coreceptor switching, and it is unclear mechanistically or structurally how X4 tropism is determined.We previously derived a replication-competent variant of the R5X4 HIV-1 clone R3A that contained a markedly truncated V3 loop (47). This Env was generated by introducing a mutation termed ΔV3(9,9), which deleted the distal 15 amino acids of V3. The ΔV3(9,9) mutation selectively ablated X4 tropism but left R5 tropism intact, consistent with the view that an interaction between the distal half of V3 and the ECLs is critical for CXCR4 usage (7, 21, 43, 50, 59, 60, 63, 72). This V3-truncated virus provided a unique opportunity to address whether CXCR4 utilization could be regained on a background in which this critical V3-ECL interaction had been ablated and, if so, by what mechanism. Here, we characterize two novel X4 variants of R3A ΔV3(9,9) derived by adapting this virus to replicate in CXCR4+ CCR5 SupT1 cells. We show that R3A ΔV3(9,9) could indeed reacquire X4 tropism but through two markedly different mechanisms. One X4 variant, designated ΔV3-X4A, acquired changes in the V3 remnant that reestablished an interaction with the CXCR4 ECLs; the other, ΔV3-X4B, acquired changes outside V3 that engendered interactions with the CXCR4 NT. These divergent evolutionary pathways led to profound differences in sensitivity to the CXCR4 antagonist AMD3100, with ΔV3-X4A showing increased sensitivity relative to R3A and with ΔV3-X4B becoming completely resistant. These findings demonstrate the contributions that interactions with distinct coreceptor regions have in mediating tropism and drug sensitivity and illustrate how HIV''s remarkable evolutionary plasticity in adapting to selection pressures can be exploited to better understand its biological potential.  相似文献   

7.
8.
9.
The processing of lagging-strand intermediates has not been demonstrated in vitro for herpes simplex virus type 1 (HSV-1). Human flap endonuclease-1 (Fen-1) was examined for its ability to produce ligatable products with model lagging-strand intermediates in the presence of the wild-type or exonuclease-deficient (exo) HSV-1 DNA polymerase (pol). Primer/templates were composed of a minicircle single-stranded DNA template annealed to primers that contained 5′ DNA flaps or 5′ annealed DNA or RNA sequences. Gapped DNA primer/templates were extended but not significantly strand displaced by the wild-type HSV-1 pol, although significant strand displacement was observed with exo HSV-1 pol. Nevertheless, the incubation of primer/templates containing 5′ flaps with either wild-type or exo HSV-1 pol and Fen-1 led to the efficient production of nicks that could be sealed with DNA ligase I. Both polymerases stimulated the nick translation activity of Fen-1 on DNA- or RNA-containing primer/templates, indicating that the activities were coordinated. Further evidence for Fen-1 involvement in HSV-1 DNA synthesis is suggested by the ability of a transiently expressed green fluorescent protein fusion with Fen-1 to accumulate in viral DNA replication compartments in infected cells and by the ability of endogenous Fen-1 to coimmunoprecipitate with an essential viral DNA replication protein in HSV-1-infected cells.Herpes simplex virus type 1 (HSV-1), the prototypic member of the family of Herpesviridae and that of the alphaherpesviridae subfamily, has served as the model for understanding the replication of herpesvirus genomes during lytic virus replication (29). The 152-kbp genome of herpes simplex virus type 1 (HSV-1) possesses approximately 85 genes, 7 of which have been shown to be necessary and sufficient for viral DNA replication within host cells (reviewed in references 5 and 38). These seven genes encode a DNA polymerase (pol) and its processivity factor (UL42), a heterotrimeric complex containing a DNA helicase (UL5), primase (UL52), and noncatalytic accessory protein (UL8), a single-stranded DNA binding protein (infected cell protein 8 [ICP-8]), and an origin binding protein with DNA helicase activity (UL9). There is strong evidence in support of the circularization of the linear virion DNA shortly after entry, and DNA replication then is thought to initiate at one or more of the three redundant origins of replication (29, 38). At least in the earliest stages of viral DNA replication, UL9 protein is required, presumably to bind to and unwind the DNA and to attract the other DNA replication proteins (29, 38). The electron microscopic examination of pulse-labeled replicating HSV-1 DNA indicates the presence of lariats, eye-forms, and D-forms (21), which is consistent with bidirectional theta-like replication from origins. To date, however, no biochemical assay has demonstrated origin-dependent DNA replication in vitro. However, in the absence of UL9, the other six HSV DNA replication proteins can support initiation and replication from a circular single-stranded DNA (ssDNA) template in an origin-independent fashion (15, 26), resembling the rolling-circle mode of replication thought to occur during the later stages of viral replication.Although nicks and small gaps have been observed in isolated replicating and virion DNA (38), the evidence for bidirectional duplex synthesis, the rapid rate of viral DNA replication, and the absence of long stretches of ssDNA in replicating and mature DNA isolated from HSV-1-infected cells suggest that leading- and lagging-strand synthesis are closely coordinated in vivo. Falkenberg et al. (15) used a minicircle DNA template with a strand bias and the six essential HSV-1 DNA replication proteins needed for rolling circle replication to demonstrate lagging-strand synthesis in vitro. However, replication from the parental strand template (leading-strand synthesis) was more efficient than synthesis from the complementary-strand template (lagging-strand synthesis). These results suggest the possibility that one or more host functions required for efficient lagging-strand synthesis or for its close coordination with leading-strand synthesis is missing in such in vitro systems.Although leading- and lagging-strand syntheses share many of the same requirements for bulk DNA synthesis, lagging-strand synthesis is a more complex process. Because the direction of polymerization of lagging-strand intermediates is opposite the direction of replication fork movement, lagging-strand synthesis requires that priming and extension occur many times to produce discontinuous segments called Okazaki fragments (reviewed in reference 25). Okazaki fragments need to be processed to remove the RNA primer, to fill in the area previously occupied by the RNA, and to seal the remaining nick between fragments, all of which must occur efficiently, accurately, and completely. Failure to do so would result in the accumulation of DNA breaks, multiple mutations, delayed DNA replication, and/or cell death (16, 61).In eukaryotes, what is currently known regarding the process of lagging-strand synthesis is based on genetic and biochemical studies with Saccharomyces cerevisiae and on in vitro reconstitution studies to define the mammalian enzymes required for simian virus 40 (SV40) T-antigen-dependent DNA replication (17, 37, 44, 57, 58). These studies have revealed that the extension of a newly synthesized Okazaki fragment DNA with pol δ causes the strand displacement of the preceding fragment to produce a 5′ flap (25). Results suggest that flap endonuclease 1 (Fen-1) is the activity responsible for the removal of the bulk of the 5′ flaps generated (1, 44, 48), although dna2 protein may facilitate the removal of longer flaps coated with the ssDNA binding protein complex (2, 44). In addition, the overexpression of exonuclease I can partially compensate for the loss of Fen-1 function in yeast (24, 51). For the proper processing of lagging-strand intermediates, the entire 5′ flap and all of the RNA primer need to be removed, and the gap must be filled to achieve a ligatable nick. DNA ligase I has been shown to be the enzyme involved in sealing Okazaki fragments in yeast and in humans (3, 31, 50, 56, 57). DNA ligase I requires a nick in which there is a 5′ phosphate on one end and a 3′ hydroxyl linked to a deoxyribose sugar entity on the other, and it works poorly in the presence of mismatches (54). The close coordination of Fen-1 and DNA ligase I activities for Okazaki fragment processing is facilitated by the interactions of these proteins with proliferating cell nuclear antigen (PCNA), the processivity factor for pol δ and ɛ (6, 30, 32, 46, 52, 53).HSV-1 does not appear to encode a protein with DNA ligase activity or one that can specifically cleave 5′ flaps, although it does encode a 5′-to-3′ exonuclease activity (UL12 [10, 20]) and a 3′-to-5′ exonuclease activity that is part of the HSV-1 pol catalytic subunit (27). As for most eukaryotes, RNA primers are essential for HSV-1 DNA synthesis, as demonstrated by the presence of oligoribonucleotides in replicating DNA in vivo (4), by the well-characterized ability of the UL52 protein in complex with the UL5 helicase activity to synthesize oligoribonucleotide primers on ssDNA in vitro (11, 13), and by the requirement of the conserved catalytic residues in the UL52 primase in vitro and in HSV-1-infected cells (14, 26). It is the strand displacement activity of pol δ that produces the 5′ flaps that are key to the removal of RNA primers during Okazaki fragment processing (6, 25). However, we previously demonstrated that wild-type HSV-1 DNA polymerase possesses poor strand displacement activity (62), in contrast to mammalian DNA pol δ (25). Thus, it is not apparent how RNA primers would be removed when encountered by HSV-1 pol during HSV-1 lagging-strand synthesis or how such intermediates would be processed.We wished to test the hypothesis that the nick translation activity of mammalian Fen-1 could function in collaboration with HSV-1 pol to facilitate the proper removal of RNA primers and/or short flaps to produce the ligatable products required for Okazaki fragment processing. In this report, we have examined the ability of wild-type and exonuclease-deficient (exo) HSV-1 pol, which differ in their respective strand displacement activities, to extend model lagging-strand substrates in the presence or absence of mammalian Fen-1. Our results demonstrate that both wild-type and exo HSV-1 pol can cooperate with and enhance Fen-1 activity to achieve a ligatable nick in vitro. Moreover, colocalization and coimmunoprecipitation studies reveal a physical association of Fen-1 with HSV-1 DNA replication proteins, supporting a model for the involvement of Fen-1 in HSV-1 DNA replication.  相似文献   

10.
11.
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   

12.
PTP1B−/− mice are resistant to diet-induced obesity due to leptin hypersensitivity and consequent increased energy expenditure. We aimed to determine the cellular mechanisms underlying this metabolic state. AMPK is an important mediator of leptin''s metabolic effects. We find that α1 and α2 AMPK activity are elevated and acetyl-coenzyme A carboxylase activity is decreased in the muscle and brown adipose tissue (BAT) of PTP1B−/− mice. The effects of PTP1B deficiency on α2, but not α1, AMPK activity in BAT and muscle are neuronally mediated, as they are present in neuron- but not muscle-specific PTP1B−/− mice. In addition, AMPK activity is decreased in the hypothalamic nuclei of neuronal and whole-body PTP1B−/− mice, accompanied by alterations in neuropeptide expression that are indicative of enhanced leptin sensitivity. Furthermore, AMPK target genes regulating mitochondrial biogenesis, fatty acid oxidation, and energy expenditure are induced with PTP1B inhibition, resulting in increased mitochondrial content in BAT and conversion to a more oxidative muscle fiber type. Thus, neuronal PTP1B inhibition results in decreased hypothalamic AMPK activity, isoform-specific AMPK activation in peripheral tissues, and downstream gene expression changes that promote leanness and increased energy expenditure. Therefore, the mechanism by which PTP1B regulates adiposity and leptin sensitivity likely involves the coordinated regulation of AMPK in hypothalamus and peripheral tissues.Protein tyrosine phosphatase 1B (PTP1B) belongs to a family of tyrosine phosphatases with diverse roles in eukaryotes (2, 4). PTP1B attenuates insulin signaling by dephosphorylating the insulin receptor (19, 22, 61) and possibly IRS-1 (9, 23) and leptin signaling by dephosphorylating JAK2, which phosphorylates the leptin receptor and associated substrates (10, 45, 67). PTP1B-deficient mice are insulin hypersensitive, lean, and resistant to diet-induced obesity (20, 36) due, at least in part, to increased energy expenditure (36). The leanness can be explained by the absence of PTP1B in neurons, because neuron-specific PTP1B−/− mice also have reduced body weight and adiposity and increased energy expenditure (6). In contrast, muscle- and liver-specific PTP1B-deficient mice have normal body weight with improved insulin sensitivity, whereas adipose-PTP1B-deficient mice have increased body weight (6, 15, 16). These data suggest that PTP1B in peripheral tissues such as muscle and liver is an important mediator of peripheral insulin sensitivity, whereas PTP1B in the nervous system plays a critical role in regulating energy expenditure and adiposity (6).The adipocyte-derived hormone leptin plays an essential role in regulating energy homeostasis by acting on multiple tissues, most importantly the hypothalamus, to regulate food intake and energy expenditure (1). PTP1B−/− mice have enhanced basal and leptin-stimulated hypothalamic STAT3 phosphorylation and are hypersensitive to leptin''s effect on food intake and body weight (10, 67). The overexpression of PTP1B in heterologous cells dose dependently reduces the leptin-induced phosphorylation of JAK2 and STAT3 and inhibits leptin-stimulated STAT3-dependent reporter gene activation (10, 35, 39, 67). These and other data established that enhanced leptin sensitivity contributes to the leanness in PTP1B−/− mice. We sought to determine the cellular mechanisms underlying the altered energy homeostasis in the setting of PTP1B deficiency.AMP-activated protein kinase (AMPK) is a major mediator of leptin''s metabolic effects (43, 44). AMPK is a fuel-sensing enzyme complex activated by cellular stresses that increase AMP or deplete ATP, including hypoxia, ischemia, glucose deprivation, uncouplers of oxidative phosphorylation, exercise, and muscle contraction (66). AMPK also is activated by the antidiabetic drugs metformin (68) and the thiazolidinediones (21). Mechanisms involved in AMPK activation include (i) the binding of AMP to an allosteric site on the γ subunit, which renders the holoenzyme resistant to inactivating serine phosphatases and also may have direct allosteric effects on kinase activity (55), and (ii) phosphorylation by upstream AMPK kinases of the α (catalytic) subunits on Thr172, which is essential for kinase activity (29). Once activated, AMPK phosphorylates multiple downstream substrates, leading to the inhibition of ATP-utilizing pathways, such as fatty acid synthesis, and the activation of ATP-generating pathways, including fatty acid oxidation (34).The phosphorylation of acetyl coenzyme A (acetyl-CoA) carboxylase (ACC) by AMPK results in the inhibition of ACC activity, decreased malonyl-CoA content, and a subsequent increase in fatty acid oxidation in skeletal muscle caused by the disinhibition of carnitine palmitoyltransferase 1 (27, 52, 62). The leptin stimulation of muscle fatty acid oxidation is mediated by AMPK (44). AMPK also is an important regulator of muscle mitochondrial biogenesis and function (7, 37, 48, 58, 63). This may, in part, be mediated by peroxisome proliferator-activated receptor γ (PPARγ)-coactivator 1α (PGC-1α), because AMPK induces the expression and phosphorylation of PGC-1α, which regulates mitochondrial biogenesis and muscle fiber type (31).In addition to a role for AMPK in leptin action in peripheral tissues, the inhibition of hypothalamic AMPK activity by leptin plays an important role in mediating leptin''s effect on food intake and energy homeostasis (43). This appears to involve neurons that express neuropeptide Y (NPY) and agouti-related peptide (AgRP), since the expression of constitutively active AMPK in the basomedial hypothalamus augments NPY/AgRP expression (43). Furthermore, the deletion of the AMPK α2 catalytic subunit specifically in these neurons results in leanness, whereas deletion in proopiomelanocortin (POMC)-expressing neurons results in mild obesity (13).To determine whether alterations in AMPK contribute to increased energy expenditure and leanness in PTP1B−/− mice, we investigated the AMPK pathway in peripheral tissues and hypothalamus. We demonstrate that the global absence of PTP1B alters AMPK and downstream biological processes in multiple tissues, and that neuronal PTP1B regulates AMPK activity in peripheral tissues in an isoform-specific manner. Our data establish a novel link between PTP1B and AMPK, two signaling molecules that are critical in the regulation of energy homeostasis.  相似文献   

13.
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   

14.
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   

15.
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   

16.
17.
Intracranial (i.c.) infection of mice with lymphocytic choriomeningitis virus (LCMV) results in anorexic weight loss, mediated by T cells and gamma interferon (IFN-γ). Here, we assessed the role of CD4+ T cells and IFN-γ on immune cell recruitment and proinflammatory cytokine/chemokine production in the central nervous system (CNS) after i.c. LCMV infection. We found that T-cell-depleted mice had decreased recruitment of hematopoietic cells to the CNS and diminished levels of IFN-γ, CCL2 (MCP-1), CCL3 (MIP-1α), and CCL5 (RANTES) in the cerebrospinal fluid (CSF). Mice deficient in IFN-γ had decreased CSF levels of CCL3, CCL5, and CXCL10 (IP-10), and decreased activation of both resident CNS and infiltrating antigen-presenting cells (APCs). The effects of IFN-γ signaling on macrophage lineage cells was assessed using transgenic mice, called “macrophages insensitive to interferon gamma” (MIIG) mice, that express a dominant-negative IFN-γ receptor under the control of the CD68 promoter. MIIG mice had decreased levels of CCL2, CCL3, CCL5, and CXCL10 compared to controls despite having normal numbers of LCMV-specific CD4+ T cells in the CNS. MIIG mice also had decreased recruitment of infiltrating macrophages and decreased activation of both resident CNS and infiltrating APCs. Finally, MIIG mice were significantly protected from LCMV-induced anorexia and weight loss. Thus, these data suggest that CD4+ T-cell production of IFN-γ promotes signaling in macrophage lineage cells, which control (i) the production of proinflammatory cytokines and chemokines, (ii) the recruitment of macrophages to the CNS, (iii) the activation of resident CNS and infiltrating APC populations, and (iv) anorexic weight loss.Immune cell recruitment to and infiltration of the central nervous system (CNS) is central to the pathology of a variety of inflammatory neurological diseases, including infectious meningoencephalitis, multiple sclerosis, and cerebral ischemia (59, 60). Chemokines have been shown to be highly upregulated in both human diseases and animal models of neuroinflammation and are thought to be important mediators of immune cell entry into the CNS (59, 60). For example, during experimental autoimmune encephalomyelitis (EAE) and multiple sclerosis (MS), the chemokines CCL2 (monocyte chemoattractant protein 1 [MCP-1α]), CCL3 (macrophage inflammatory protein 1α [MIP-1α]), CCL5 (regulated upon activation, T-cell expressed and secreted [RANTES]), and CXCL10 (gamma interferon [IFN-γ]-inducible protein 10 [IP-10]) are produced by either resident CNS cells or infiltrating cells (27) and serve to amplify the ongoing inflammatory response (25, 28). However, in some EAE studies, neither CCL3 nor CXCL10 were required for disease (72, 73). During CNS viral infection, CXCL10 and CCL5 are highly produced in several models (2, 41, 48, 82). In addition, mice deficient in CCR5, which binds (among others) CCL3 and CCL5, do not display impaired CNS inflammation after certain viral infections (13). Thus, the role of chemokines in CNS inflammation is likely complex and dissimilar between autoimmune and viral infection models.IFN-γ is present in the CNS during autoimmunity and infection (7, 54, 69). Several studies suggest that IFN-γ can be a potent inducer of CNS chemokine expression. Adenoviral expression of IFN-γ in the CNS strongly induced CCL5 and CXCL10 mRNA and protein, and this induction was dependent on the presence of the IFN-γ receptor (50). In EAE and Toxoplasma infection, mice deficient in IFN-γ or the IFN-γ receptor demonstrated reduced expression of several chemokines, including CCL2, CCL3, CCL5, and CXCL10 (26, 69). However, given the near-ubiquitous expression of the IFN-γ receptor (44), the mechanisms by which IFN-γ regulates CNS chemokine production remain to be elucidated.We studied neuroinflammation and immune-mediated disease using a well-studied mouse model of infection with lymphocytic choriomeningitis virus (LCMV). Intracranial (i.c.) injection of mice with LCMV results in seizures and death 6 to 8 days after inoculation. The onset of symptoms is associated with a massive influx of mononuclear cells into the cerebrospinal fluid (CSF), meninges, choroid plexus, and ependymal membranes (6, 8, 18), as well as the presence of proinflammatory cytokines (7, 38). The immune response is critical for disease, since infection of irradiated or T-cell-depleted mice leads to persistent infection with very high levels of virus in multiple tissues without the development of lethal meningitis (18, 34, 64). i.c. LCMV infection of β2-microglobulin-deficient mice (β2m−/− mice) also results in meningitis and production of proinflammatory cytokines and chemokines; however, meningitis occurs with a later onset and lower severity compared to wild-type mice (17, 24, 53, 57). Interestingly, i.c. LCMV infection of these mice also causes severe anorexia and weight loss (33, 38, 46, 52, 57) that is mediated by major histocompatibility complex (MHC) class II-restricted, CD4+ T cells (17, 46, 53, 57). Anorexia and weight loss are also observed in wild-type mice, but they succumb to lethal meningitis shortly thereafter (33), making study of this particular aspect of disease difficult. LCMV-induced weight loss, similar to what we have observed in β2m−/− mice also occurs in perforin-deficient mice, which possess CD8+ T cells (37). Although some reports have observed weight loss after peripheral LCMV infection (11, 45), we note that these studies used high doses of the clone 13 strain of LCMV, in contrast to our studies which have used the Armstrong strain of LCMV and orders of magnitude less virus (33, 38, 46, 52, 57). Although we cannot exclude a contribution of peripheral cells to weight loss in our i.c. Armstrong infection model, we previously showed that this weight loss does not occur with peripheral infection with LCMV Armstrong (33, 38), indicating that interactions between the CNS and the immune system are contribute substantially to disease.During LCMV infection, there is biphasic production of IFN-γ: a small, early peak of IFN-γ (most likely produced by NK or NKT cells), followed by T-cell-mediated production of IFN-γ (23, 75). Further, both CD4+ T cells and CD8+ T cells produce large amounts of IFN-γ after LCMV infection and T-cell production of IFN-γ is critical for LCMV-induced weight loss (35). Chemokines, especially CXCL10, CCL5, and CCL2, and their receptors, are upregulated in the brain after i.c. LCMV infection (2, 13). Brain chemokine mRNA expression after i.c. LCMV infection is reduced in IFN-γ-deficient mice and relatively absent in athymic mice (2). However, the mechanism(s) by which T cells and IFN-γ mediate the effects on CNS chemokine expression, cellular infiltration into the CNS, and LCMV-induced anorexic weight loss remain unclear.In the present study, we focused on two major questions. The first question concerned the role of IFN-γ on immune cell recruitment to and chemokine/cytokine production within the CNS? We found that macrophages and myeloid dendritic cells (DCs) require IFN-γ for their accumulation within the CNS. Second, since macrophages and myeloid DCs are the predominant cellular infiltrate, we sought to determine whether IFN-γ signaling on these cells was direct with regard to their recruitment and to chemokine/cytokine production. We found that IFN-γ signaling in macrophage lineage cells contributes significantly to their recruitment, to chemokine production in the CNS, and to anorexic weight loss. Together, these data suggest that much of the proinflammatory effects of IFN-γ in the CNS are mediated by the effects of IFN-γ on CD68-bearing cells.  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