首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
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   

2.
3.
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   

4.
The Na+-dependent K+ uptake KtrABE system is essential for the adaptation of Synechocystis to salinity stress and high osmolality. While KtrB forms the K+-translocating pore, the role of the subunits KtrA and KtrE for Ktr function remains elusive. Here, we characterized the role of KtrA and KtrE in Ktr-mediated K+ uptake and in modulating Na+ dependency. Expression of KtrB alone in a K+ uptake-deficient Escherichia coli strain conferred low K+ uptake activity that was not stimulated by Na+. Coexpression of both KtrA and KtrE with KtrB increased the K+ transport activity in a Na+-dependent manner. KtrA and KtrE were found to be localized to the plasma membrane in Synechocystis. Site-directed mutagenesis was used to analyze the role of single charged residues in KtrB for Ktr function. Replacing negatively charged residues facing the extracellular space with residues of the opposite charge increased the apparent Km for K+ in all cases. However, none of the mutations eliminated the Na+ dependency of Ktr-mediated K+ transport. Mutations of residues on the cytoplasmic side had larger effects on K+ uptake activity than those of residues on the extracellular side. Further analysis revealed that replacement of R262, which is well conserved among Ktr/Trk/HKT transporters in the third extracellular loop, by Glu abolished transport activity. The atomic-scale homology model indicated that R262 might interact with E247 and D261. Based on these data, interaction of KtrA and KtrE with KtrB increased the K+ uptake rate and conferred Na+ dependency.Cyanobacterium Synechocystis sp. strain PCC 6803 contains a number of different K+ uptake systems that may contribute to satisfying its requirement of K+ (3, 19, 36). Among these systems, Ktr has been shown to have a major role not only in K+ uptake but also in adaptation against high-osmolarity stress (3, 19). Inactivation of the ktr gene renders the cells hypersensitive to high concentrations of NaCl and the nonionic compound sorbitol. Ktr-mediated K+ uptake depends on the presence of Na+ in the medium, which is likely to be an adaptation to salinity stress. A requirement of Na+ for K+ transport activity has also been found in the homologous protein from Vibrio alginolyticus (21). This dependency on Na+ is a unique property of Ktr-type transporters and has not been found in other types of K+ transporters or channels (32). The structure and function of Ktr-type transporters have been studied in a number of organisms (3, 6, 7, 9, 11-14, 18-20, 30, 32-34). The Ktr system from Synechocystis consists of three subunits, KtrA, KtrB, and KtrE (19). The KtrE gene and the KtrB gene form a cistron, whereas the KtrA gene resides at a site distant from the KtrEB genes in the Synechocystis genome (19). KtrB, the K+-translocating subunit, is a member of the Ktr/Trk/HKT family of K+ transporters. These transporters have been proposed to have evolved from two membrane-spanning K+ channels (6, 7). According to the model, this type of transporter contains eight transmembrane domains, which consist of a 4-fold-repeated membrane-pore-membrane (M1-P-M2) motif (6, 7, 13, 18). An intramolecular electrostatic interaction of Synechocystis KtrB has been proposed to stabilize the protein in its active configuration (12). In addition, a conserved His in the external region in Synechocystis KtrB has been shown to be crucial for KtrB function (39). The region of the Vibrio Ktr protein responsible for gating of ion permeation has been identified (9). However, not much is known about the mechanism of Na+ binding to KtrB in Synechocystis.The KtrA subunit belongs to the family of KTR (K+-transport nucleotide binding)/RCK (regulating the conductance of K+ channels) proteins, which contain a Rossmann-fold sequence encoding β-α protein structure for NAD+/NADH binding (17). Accordingly KtrA has been proposed to regulate the K+ transport activity of KtrB by changing its binding from NAD+ to NADH through a ligand-mediated conformational switch mechanism (25). It has also been shown that ATP promotes complex formation between KtrA and KtrB and that KtrAB from V. alginolyticus when expressed in Escherichia coli cells requires both ATP and the membrane potential for its activity (17).KtrE is a unique subunit found only in Synechocystis; it is not involved in KtrB-mediated K+ transport in V. alginolyticus and Bacillus subtilis (11, 32). The termination codon of ktrE overlaps the initiation codon of ktrB in the same cistron, which has not been found in other bacterial ktrB-related genes. Coexpression of KtrA with KtrB alone does not complement the growth defect of an E. coli K+ uptake mutant. However, introduction of KtrE into the same mutant background in addition to KtrA and KtrB complements the mutation of the K+ uptake system (19). Interestingly, the KtrE protein has been shown to function as a digalactosyldiacylglycerol (DGDG) synthase (EC 2.4.6.241), an enzyme that produces DGDG from monogalactosyldiacylglycerol (MGDG). KtrE has therefore also been designated DgdA (1). Under nonstress conditions, DGDG is found in the thylakoid membranes, which helps stabilize the photosystem II complex in Synechocystis (29). Under phosphate-limited conditions, DGDG is synthesized instead of phospholipids in Synechocystis (1). However, KtrB functions as a major K+-conducting transport pore in the Synechocystis plasma membrane. The subcellular localization of KtrE has not been identified directly. Inactivation of ktrE (also called dgdA) in Synechocystis does not result in sensitivity to osmotic stress imposed by 300 mM sorbitol (1). This may be inconsistent with the requirement of KtrE for KtrB-mediated K+ uptake in the presence of KtrA in the E. coli expression system (19).Because of these uncertainties about the roles of the KtrA and KtrE subunits in K+ uptake by KtrB in Synechocystis and about the identity of the Na+ binding site in KtrB, we examined the subcellular localization and membrane association of KtrA and KtrE, the requirement of these subunits for KtrB-mediated K+ uptake, and the primary target for Na+ binding in KtrB.  相似文献   

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

6.
Candidate HIV-1 vaccine regimens utilizing intramuscularly (i.m.) administered recombinant adenovirus (rAd)-based vectors can induce potent mucosal cellular immunity. However, the degree to which mucosal rAd vaccine routing might alter the quality and anatomic distribution of vaccine-elicited CD8+ T lymphocytes remains unclear. We show that the route of vaccination critically impacts not only the magnitude but also the phenotype and trafficking of antigen-specific CD8+ T lymphocytes in mice. I.m. rAd immunization induced robust local transgene expression and elicited high-frequency, polyfunctional CD8+ T lymphocytes that trafficked broadly to both systemic and mucosal compartments. In contrast, intranasal (i.n.) rAd immunization led to similarly robust local transgene expression but generated low-frequency, monofunctional CD8+ T lymphocytes with restricted anatomic trafficking patterns. Respiratory rAd immunization elicited systemic and mucosal CD8+ T lymphocytes with phenotypes and trafficking properties distinct from those elicited by i.m. or i.n. rAd immunization. Our findings indicate that the anatomic microenvironment of antigen expression critically impacts the phenotype and trafficking of antigen-specific CD8+ T lymphocytes.Acute human immunodeficiency virus type 1 (HIV-1) infection is accompanied by a massive, irreversible destruction of memory CD4+ T lymphocytes, particularly within the intestinal mucosa (11, 26, 30, 42), as a result of the high proportion of effector/memory target cells within the intestinal lamina propria. Chronic HIV-1 infection is characterized by inflammation within the intestinal mucosa, breakdown of epithelial-barrier integrity, and translocation of gut microflora from the intestinal lumen (10, 24). These processes may drive systemic inflammation and contribute to HIV-1 disease progression. Therefore, vaccination strategies that enhance mucosal cellular immunity and attenuate the mucosal immunopathology of HIV-1 infection would be desirable.Recombinant adenovirus (rAd) vectors are potent inducers of cellular immunity (3, 12, 25), and we have recently demonstrated that intramuscular (i.m.) rAd immunization transiently activates peripheral antigen-specific CD8+ T lymphocytes and allows them to migrate to mucosal surfaces and establish potent, durable mucosal cellular immunity (22). Moreover, we have shown that an i.m. delivered heterologous rAd prime-boost regimen prevented the destruction of CD4+ T lymphocytes within the intestinal mucosa and attenuated disease progression following simian immunodeficiency virus (SIV) challenge (29). Notably, this vaccine regimen did not contain the SIV Env protein, indicating that cellular mucosal immunity likely played a critical role in abrogating mucosal CD4+ T-lymphocyte destruction.While our laboratory and others have observed potent mucosal CD8+ T-lymphocyte responses after i.m. immunization with rAd vectors (2, 21, 28, 41) and other vaccine modalities (40-41), other studies have suggested that mucosal routing of vaccine vectors may optimize mucosal cellular immunity (4-7, 16, 33, 36, 46). We therefore assessed the phenotype and anatomic trafficking patterns of antigen-specific CD8+ T-lymphocyte responses following i.m. and mucosal rAd immunization in mice. We found that the immunization route dramatically impacted the phenotype of vaccine-elicited systemic and mucosal CD8+ T lymphocytes. In particular, while both i.m. and intranasal (i.n.) rAd immunization resulted in efficient local transgene expression, only i.m. immunization induced potent, polyfunctional cellular immune memory in both systemic and mucosal anatomic compartments, while i.n. immunization elicited lower-frequency cellular immune responses that were restricted to mucosal surfaces and characterized by monofunctional gamma interferon (IFN-γ) secretion. Our data highlight the critical impact of the route of antigen delivery and the anatomic microenvironment of transgene expression on the quality and distribution of vaccine-elicited CD8+ T-lymphocyte responses.  相似文献   

7.
8.
9.
A restricted number of studies have shown that human immunodeficiency virus type 1 (HIV-1)-specific cytotoxic CD4+ T cells are present in HIV-1-infected individuals. However, the roles of this type of CD4+ T cell in the immune responses against an HIV-1 infection remain unclear. In this study, we identified novel Nef epitope-specific HLA-DRB1*0803-restricted cytotoxic CD4+ T cells. The CD4+ T-cell clones specific for Nef187-203 showed strong gamma interferon production after having been stimulated with autologous B-lymphoblastoid cells infected with recombinant vaccinia virus expressing Nef or pulsed with heat-inactivated virus particles, indicating the presentation of the epitope antigen through both exogenous and endogenous major histocompatibility complex class II processing pathways. Nef187-203-specific CD4+ T-cell clones exhibited strong cytotoxic activity against both HIV-1-infected macrophages and CD4+ T cells from an HLA-DRB1*0803+ donor. In addition, these Nef-specific cytotoxic CD4+ T-cell clones exhibited strong ability to suppress HIV-1 replication in both macrophages and CD4+ T cells in vitro. Nef187-203-specific cytotoxic CD4+ T cells were detected in cultures of peptide-stimulated peripheral blood mononuclear cells (PBMCs) and in ex vivo PBMCs from 40% and 20% of DRB1*0803+ donors, respectively. These results suggest that HIV-1-specific CD4+ T cells may directly control HIV-1 infection in vivo by suppressing virus replication in HIV-1 natural host cells.Human immunodeficiency virus (HIV)-specific CD8+ cytotoxic T cells (CTLs) play a central role in the control of HIV type 1 (HIV-1) during acute and chronic phases of an HIV-1 infection (5, 29, 34). However, HIV-1 escapes from the immune surveillance of CD8+ CTLs by mechanisms such as mutations of immunodominant CTL epitopes and downregulation of major histocompatibility complex class I (MHC-I) molecules on the infected cells (9, 11, 12, 49). Therefore, most HIV-1-infected patients without highly active antiretroviral therapy (HAART) develop AIDS eventually.HIV-1-specific CD4+ T cells also play an important role in host immune responses against HIV-1 infections. An inverse association of CD4+ T-cell responses with viral load in chronically HIV-1-infected patients was documented in a series of earlier studies (8, 36, 39, 41, 48), although the causal relationship between them still remains unclear (23). Classically, CD4+ T cells help the expansion of CD8+ CTLs by producing growth factors such as interleukin-2 (IL-2) or by their CD40 ligand interaction with antigen-processing cells and CD8+ CTLs. In addition, CD4+ T cells provide activation of macrophages, which can professionally maintain CD8+ T-cell memory (17). On the other hand, the direct ability of virus-specific cytotoxic CD4+ T cells (CD4+ CTLs) to kill target cells has been widely observed in human virus infections such as those by human cytomegalovirus, Epstein-Barr virus (EBV), hepatitis B virus, Dengue virus, and HIV-1 (2, 4, 10, 19, 30, 31, 38, 50). Furthermore, one study showed that mouse CD4+ T cells specific for lymphocytic choriomeningitis virus have cytotoxic activity in vivo (25). These results, taken together, indicate that a subset of effector CD4+ T cells develops cytolytic activity in response to virus infections.HIV-1-specific CD4+ CTLs were found to be prevalent in HIV-1 infections, as Gag-specific cytotoxic CD4+ T cells were detected directly ex vivo among peripheral blood mononuclear cells (PBMCs) from an HIV-1-infected long-term nonprogressor (31). Other studies showed that up to 50% of the CD4+ T cells in some HIV-1-infected donors can exhibit a clear cytolytic potential, in contrast to the fact that healthy individuals display few of these cells (3, 4). These studies indicate the real existence of CD4+ CTLs in HIV-1 infections.The roles of CD4+ CTLs in the control of an HIV-1 infection have not been widely explored. It is known that Gag-specific CD4+ CTLs can suppress HIV-1 replication in a human T-cell leukemia virus type 1-immortalized CD4+ T-cell line (31). However, the functions of CD4+ T cells specific for other HIV-1 antigens remain unclear. On the other hand, the abilities of CD4+ CTLs to suppress HIV-1 replication in infected macrophages and CD4+ T cells may be different, as in the case of CD8+ CTLs for HIV-1-infected macrophages (17). In this study, we identified Nef-specific CD4+ T cells and investigated their ability to kill HIV-1 R5 virus-infected macrophages and HIV-1 X4 virus-infected CD4+ T cells and to suppress HIV-1 replication in the infected macrophages and CD4+ T cells. The results obtained in the present study show for the first time the ability of HIV-1-specific CD4+ CTLs to suppress HIV-1 replication in natural host cells, i.e., macrophages and CD4+ T cells.  相似文献   

10.
11.
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   

12.
13.
The immune correlates of human/simian immunodeficiency virus control remain elusive. While CD8+ T lymphocytes likely play a major role in reducing peak viremia and maintaining viral control in the chronic phase, the relative antiviral efficacy of individual virus-specific effector populations is unknown. Conventional assays measure cytokine secretion of virus-specific CD8+ T cells after cognate peptide recognition. Cytokine secretion, however, does not always directly translate into antiviral efficacy. Recently developed suppression assays assess the efficiency of virus-specific CD8+ T cells to control viral replication, but these assays often use cell lines or clones. We therefore designed a novel virus production assay to test the ability of freshly ex vivo-sorted simian immunodeficiency virus (SIV)-specific CD8+ T cells to suppress viral replication from SIVmac239-infected CD4+ T cells. Using this assay, we established an antiviral hierarchy when we compared CD8+ T cells specific for 12 different epitopes. Antiviral efficacy was unrelated to the disease status of each animal, the protein from which the tested epitopes were derived, or the major histocompatibility complex (MHC) class I restriction of the tested epitopes. Additionally, there was no correlation with the ability to suppress viral replication and epitope avidity, epitope affinity, CD8+ T-cell cytokine multifunctionality, the percentage of central and effector memory cell populations, or the expression of PD-1. The ability of virus-specific CD8+ T cells to suppress viral replication therefore cannot be determined using conventional assays. Our results suggest that a single definitive correlate of immune control may not exist; rather, a successful CD8+ T-cell response may be comprised of several factors.CD8+ T cells may play a critical role in blunting peak viremia and controlling human immunodeficiency virus (HIV) and simian immunodeficiency virus (SIV) replication. The transient depletion of CD8+ cells in SIV-infected macaques results in increased viral replication (26, 31, 51, 70). The emergence of virus-specific CD8+ T cells coincides with the reduction of peak viremia (12, 39, 42, 63), and CD8+ T-cell pressure selects for escape mutants (6, 9, 13, 28, 29, 38, 60, 61, 85). Furthermore, particular major histocompatibility complex (MHC) class I alleles are overrepresented in SIV- and HIV-infected elite controllers (15, 29, 33, 34, 46, 56, 88).Because it has been difficult to induce broadly neutralizing antibodies (Abs), the AIDS vaccine field is currently focused on developing a vaccine designed to elicit HIV-specific CD8+ T cells (8, 52, 53, 82). Investigators have tried to define the immune correlates of HIV control. Neither the magnitude nor the breadth of epitopes recognized by virus-specific CD8+ T-cell responses correlates with the control of viral replication (1). The quality of the immune response may, however, contribute to the antiviral efficacy of the effector cells. It has been suggested that the number of cytokines that virus-specific CD8+ T cells secrete may correlate with viral control, since HIV-infected nonprogressors appear to maintain CD8+ T cells that secrete several cytokines, compared to HIV-infected progressors (11, 27). An increased amount of perforin secretion may also be related to the proliferation of HIV-specific CD8+ T cells in HIV-infected nonprogressors (55). While those studies offer insight into the different immune systems of progressors and nonprogressors, they did not address the mechanism of viral control. Previously, we found no association between the ability of SIV-specific CD8+ T-cell clones to suppress viral replication in vitro and their ability to secrete gamma interferon (IFN-γ), tumor necrosis factor alpha (TNF-α), or interleukin-2 (IL-2) (18).Evidence suggests that some HIV/SIV proteins may be better vaccine targets than others. CD8+ T cells recognize epitopes derived from Gag as early as 2 h postinfection, whereas CD8+ T cells specific for epitopes in Env recognize infected cells only at 18 h postinfection (68). Additionally, a previously reported study of HIV-infected individuals showed that an increased breadth of Gag-specific responses was associated with lower viral loads (35, 59, 65, 66). CD8+ T-cell responses specific for Env, Rev, Tat, Vif, Vpr, Vpu, and Nef were associated with higher viral loads, with increased breadth of Env in particular being significantly associated with a higher chronic-phase viral set point.None of the many sophisticated methods employed for analyzing the characteristics of HIV- or SIV-specific immune responses clearly demarcate the critical qualities of an effective antiviral response. In an attempt to address these questions, we developed a new assay to measure the antiviral efficacy of individual SIV-specific CD8+ T-cell responses sorted directly from fresh peripheral blood mononuclear cells (PBMC). Using MHC class I tetramers specific for the epitope of interest, we sorted freshly isolated virus-specific CD8+ T cells and determined their ability to suppress virus production from SIV-infected CD4+ T cells. We then looked for a common characteristic of efficacious epitope-specific CD8+ T cells using traditional methods.  相似文献   

14.
Developing an immunotherapy to keep human immunodeficiency virus type 1 (HIV-1) replication suppressed while discontinuing highly active antiretroviral therapy (HAART) is an important challenge. In the present work, we evaluated in vitro whether dendritic cells (DC) electroporated with gag mRNA can induce HIV-specific responses in T cells from chronically infected subjects. Monocyte-derived DC, from therapy-naïve and HAART-treated HIV-1-seropositive subjects, that were electroporated with consensus codon-optimized HxB2 gag mRNA efficiently expanded T cells, secreting gamma interferon (IFN-γ) and interleukin 2 (IL-2), as well as other cytokines and perforin, upon restimulation with a pool of overlapping Gag peptides. The functional expansion levels after 1 week of stimulation were comparable in T cells from HAART-treated and treatment-naïve patients and involved both CD4+ and CD8+ T cells, with evidence of bifunctionality in T cells. Epitope mapping of p24 showed that stimulated T cells had a broadened response toward previously nondescribed epitopes. DC, from HAART-treated subjects, that were electroporated with autologous proviral gag mRNA equally efficiently expanded HIV-specific T cells. Regulatory T cells did not prevent the induction of effector T cells in this system, whereas the blocking of PD-L1 slightly increased the induction of T-cell responses. This paper shows that DC, loaded with consensus or autologous gag mRNA, expand HIV-specific T-cell responses in vitro.Studies of immune responses generated in human immunodeficiency virus type 1 (HIV-1)-infected individuals suggest that CD8+ T cells play an important role in the defense against the virus. In acute HIV infection, the appearance of HIV-specific CD8+ T cells is associated with a decline in viremia (11, 32). More-direct evidence for the role of CD8+ T cells in viral control is deduced from studies of simian immune deficiency virus (SIV)-infected rhesus macaques in which the depletion of the CD8+ T cells results in an increase of the viral load and rapid disease progression (41, 55), although this is not always the case (35). Among HIV-infected humans, long-term nonprogressors (LTNP) with an undetectable viral load have higher levels of multifunctional HIV-specific CD8+ T cells in comparison to patients with rapidly progressive disease (53). Conversely, the HIV-specific CD8+ T cells from rapid progressors release low levels of interleukin 2 (IL-2) and high levels of gamma interferon (IFN-γ), they have a reduced proliferative capacity, and their perforin expression is impaired or exhausted (42, 69). Moreover, during primary and chronic infection, viral escape mutations are often observed as a consequence of immunological pressure mediated by SIV- and HIV-specific CD8+ T cells (3, 12, 20, 23, 50). During this process of viral adaptation, all the previous variants are stored as proviral DNA (46).Although current highly active antiretroviral therapy (HAART) may suppress viral replication and protect against disease progression, it is unable to eliminate the proviral latent reservoir. Moreover, as a consequence of low or absent HIV antigenic stimulation, HIV-1-specific cytotoxic T lymphocyte (CTL) responses tend to wane during HAART (16, 39). Therapy interruption invariably results in a viral rebound to pretreatment levels, indicating that no protective immunity has been built up during therapy (38). On the other hand, the partial immune reconstitution, induced by HAART, opens a window of opportunity to boost T-cell immunity by therapeutic vaccination. Clearly, it is not sufficient to enhance the response against the circulating virus. To minimize the risk of escape, it is equally important that immune responses against the entire latent reservoir are activated (49).Dendritic cells (DC) are the most powerful antigen-presenting cells (APC) that can stimulate effective immune responses both in vitro and in vivo (5, 9, 62). In the context of DC-based immunotherapy, many groups have used DC expressing HIV antigens (e.g., pulsed with peptides, transduced with different vectors, or loaded with apoptotic infected cells) to stimulate memory (19, 34, 59, 69) or even primary (13, 14, 33, 63, 66, 67) CD8+ T cells in vitro. In vivo, SIV-specific CD8+ and CD4+ T-cell responses were induced in macaques using DC expressing SIV antigen (63). Finally, Lu and Andrieu and Lu et al. (36, 37) showed that DC pulsed with chemically inactivated autologous virus specifically stimulated HIV-specific immune responses in vitro and in vivo in cells of HIV-1-seropositive individuals.Recently, we (47, 48, 61) and others (9, 15, 22, 28, 40, 54, 57) have shown that transfection with mRNA is more effective than mRNA lipofection, peptide pulsing, or viral transduction to generate primary (65) and memory (57) responses. Furthermore, we demonstrated that DC from treatment-naïve HIV-1-seropositive subjects can efficiently be transfected with HIV gag and env mRNA, derived either from consensus subtype B or autologous viral or proviral HIV, and that these DC readily trigger autologous CD4+ and CD8+ T cells to release IFN-γ and IL-2 in a short-term ex vivo enzyme-linked immunospot (ELISPOT) assay (60).Our previous study (60) considered only the direct ex vivo immune responses of untreated HIV-1-seropositive persons, who have, by definition, a rather damaged immune system (42). Therefore, with the ultimate aim to develop an immunotherapy based on DC, we decided to evaluate the responses of treatment-naïve and HAART-treated HIV-1-seropositive persons after 1 week of stimulation with electroporated DC. Besides IFN-γ production, other parameters were also evaluated, such as a series of other cytokines, measured in various ways (by ELISPOT, microbead assay, and intracellular cytometry), and the potential influence of regulatory T cells (Treg) on the response. Finally, because HIV escapes very easily from the immune system, we also investigated if it is possible to use autologous proviral gag mRNA and to broaden the immune response.  相似文献   

15.
A broad Gag-specific CD8+ T-cell response is associated with effective control of adult human immunodeficiency virus (HIV) infection. The association of certain HLA class I molecules, such as HLA-B*57, -B*5801, and -B*8101, with immune control is linked to mutations within Gag epitopes presented by these alleles that allow HIV to evade the immune response but that also reduce viral replicative capacity. Transmission of such viruses containing mutations within Gag epitopes results in lower viral loads in adult recipients. In this study of pediatric infection, we tested the hypothesis that children may tend to progress relatively slowly if either they themselves possess one of the protective HLA-B alleles or the mother possesses one of these alleles, thereby transmitting a low-fitness virus to the child. We analyzed HLA type, CD8+ T-cell responses, and viral sequence changes for 61 mother-child pairs from Durban, South Africa, who were monitored from birth. Slow progression was significantly associated with the mother or child possessing one of the protective HLA-B alleles, and more significantly so when the protective allele was not shared by mother and child (P = 0.007). Slow progressors tended to make CD8+ T-cell responses to Gag epitopes presented by the protective HLA-B alleles, in contrast to progressors expressing the same alleles (P = 0.07; Fisher''s exact test). Mothers expressing the protective alleles were significantly more likely to transmit escape variants within the Gag epitopes presented by those alleles than mothers not expressing those alleles (75% versus 21%; P = 0.001). Reversion of transmitted escape mutations was observed in all slow-progressing children whose mothers possessed protective HLA-B alleles. These data show that HLA class I alleles influence disease progression in pediatric as well as adult infection, both as a result of the CD8+ T-cell responses generated in the child and through the transmission of low-fitness viruses by the mother.Human immunodeficiency virus (HIV)-specific CD8+ T cells play a central role in controlling viral replication (12). It is the specificity of the CD8+ T-cell response, particularly the response to Gag, that is associated with low viral loads in HIV infection (7, 17, 34). Although immune control is undermined by the selection of viral mutations that prevent recognition by the CD8+ T cells, evasion of Gag-specific responses mediated by protective class I HLA-B alleles typically brings a reduction in viral replicative capacity, facilitating subsequent immune control of HIV (2, 20, 21). The same principle has been demonstrated in studies of simian immunodeficiency virus infection (18, 22).Recent studies showed that the class I HLA-B alleles that protect against disease progression present more Gag-specific CD8+ T-cell epitopes and drive the selection of more Gag-specific escape mutations than those alleles that are associated with high viral loads (23). These protective HLA-B alleles not only are beneficial to infected individuals expressing those alleles but also benefit a recipient following transmission, since the transmitted virus carrying multiple Gag escape mutations may have substantially reduced fitness (3, 4, 8). However, there is no benefit to the recipient if he or she shares the same protective allele as the donor because the transmitted virus carries escape mutations in the Gag epitopes that would otherwise be expected to mediate successful immune control in the recipient (8, 11).The sharing of HLA alleles between donor and recipient occurs frequently in mother-to-child transmission (MTCT). The risk of MTCT is related to viral load in the mother, and a high viral load is associated with nonprotective alleles, such as HLA-B*18 and -B*5802. This may contribute in two distinct ways to the more rapid progression observed in pediatric HIV infection (24, 26, 27). First, because infected children share 50% or more of their HLA alleles with the transmitting mother, they are less likely than adults to carry protective HLA alleles (16). Thus, infected children as a group carry fewer protective HLA alleles and more nonprotective HLA alleles. Second, even when the child has a protective allele, such as HLA-B*27, this allele does not offer protection if the maternally transmitted virus carries escape mutations within the key Gag epitopes that are presented by the protective allele (11, 19).However, it is clear that infected children who possess protective alleles, such as HLA-B*27 or HLA-B*57, can achieve durable immune control of HIV infection if the virus transmitted from the mother is not preadapted to those alleles (6, 10). HIV-specific CD8+ T-cell responses are detectable from birth in infected infants (32). Furthermore, as in adult infection (3, 8), HIV-infected children have the potential to benefit from transmission of low-fitness viruses in the situation where the mother possesses protective HLA alleles and the child does not share those protective alleles. MTCT of low-fitness viruses carrying CD8+ T-cell escape mutations was recently documented (28; J. Prado et al., unpublished data).In this study, undertaken in Durban, South Africa, we set out to test the hypothesis that HIV-infected children are less likely to progress rapidly to disease if either the infected child or the transmitting mother possesses a protective HLA allele that is not shared. The HLA alleles most strongly associated with low viral loads and high CD4 counts in a cohort of >1,200 HIV-infected adults in Durban are HLA-B*57 (-B*5702 and -B*5703), HLA-B*5801, and HLA-B*8101 (16; A. Leslie et al., unpublished data). These four alleles all present Gag-specific CD8+ T-cell epitopes, and in each case the escape mutations selected in these epitopes reduce viral replicative capacity (2-4, 8, 21, 23).Analyzing a previously described cohort of 61 HIV-infected children in Durban (24, 26, 32), South Africa, who were all monitored from birth, we first addressed the question of whether possession of any of these four alleles by either mother or child is associated with slower disease progression in the child and then determined whether sharing of protective alleles by mother and child affects the ability of the child to make the Gag-specific CD8+ T-cell responses restricted by the shared allele.  相似文献   

16.
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   

17.
18.
This study examined bacteria-immune interactions in a mouse model possessing microbiota-dependent immune regulatory features similar to those occurring in human atopy, colitis, and immune regulation. Associations between the abundance of several bacterial phylotypes and immunoregulatory target cell types were identified, suggesting that they may play a role in these phenotypes.Bacteria are involved in critical aspects of immune system development and regulation (5, 23, 26, 29). Mice raised under germfree conditions exhibit a variety of abnormalities, including hypoplastic Peyer''s patches, reduced numbers of IgA-producing cells, relatively unstructured spleen and lymph nodes, and hypogammaglobulinemic serum (23). Remarkably, after several weeks of exposure to standard intestinal microbiota, normal immune structure and function are restored. Mechanistic details underlying microbe-immune interactions have been recently elucidated for two common intestinal bacteria. Bacteroides thetaiotaomicron was shown to induce the angiogenin Ang4, a component of innate immunity possessing microbicidal activity against a wide range of intestinal microbes, including both bacterial and fungal pathogens (16). In addition, studies of the Bacteroides fragilis zwitterionic capsular polysaccharide A have established it as a cognate antigen of certain CD4+ T cells, which programs immune effector polarization (24) and protection of mice from infection by Helicobacter hepaticus through several immune-mediated mechanisms (25). Resident microbiota also modify the interaction of dendritic cells with regulatory T-cell populations, with resultant susceptibility to chronic inflammatory disease, like colitis (15, 28).Recent work by Braun and colleagues has characterized a mouse model with unique immunologic features linking resident microbiota with levels of regulatory CD8+ T cells (13, 17, 39). This model is comprised of two physically isolated colonies of isogenic mice harboring distinct microbial communities: conventional floras (CF) and restricted floras (RF). CF refers to C57BL/6 mice housed in a standard specific-pathogen-free facility, while RF refers to C57BL/6 mice containing a different complement of intestinal microorganisms (13, 30), originally created by transferring several nonpathogenic anaerobic bacteria into antibiotic-treated mice (13). RF mice differ from CF mice in several immunologic phenotypes, including selective reduction of marginal zone (MZ) B cells (39), plasmacytoid dendritic cells (pDC) (13), and invariant natural killer (iNK) T cells (38a), as well as naïve CD4+ and CD8+ T cells (17). In addition, RF mice were shown to be resistant to colitis under genetic or adoptive transfer conditions that permit disease activity in CF mice (2). RF mice also cleared experimental infections by Campylobacter jejuni more slowly than did their CF counterparts (6). The resulting concept is that certain resident microbiota, which may be more abundant in RF mice than in CF mice, induce invariant Qa-1 T cells, with resultant changes in host immunoregulation and microbial surveillance (2).An important issue raised by the foregoing observations is the identity of resident microbiota responsible for this host immunoregulatory response. The objective of this study was to develop a methodology, based on bacteria-immune interactions in the RF/CF mouse model, to identify candidate microbiota. In this study, we employed a series of experiments examining associations between the population densities of bacterial rRNA genes and several immunologic features that differ between CF and RF mice.  相似文献   

19.
20.
设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号