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 15 毫秒
1.
Endothelial cells play a critical role in monocyte differentiation. Platelets also affect terminal maturation of monocytes in vitro. P-selectin is an important adhesion molecule expressed on both endothelial cells and activated platelets. We investigated its effects on human peripheral blood monocyte differentiation under the influence of different cytokines. Generation of dendritic-like cells (DLCs) from peripheral blood monocytes was promoted by immobilized P-selectin in the presence of M-CSF and IL-4 as judged by dendritic cell (DC) morphology; increased expression of CD1a, a DC marker; low phagocytic activity; and high alloreactivity to naive T cells. In contrast to typical DCs, DLCs expressed CD14 and FcgammaRIII (CD16). These features link the possible identity of DLCs to that of an uncommon CD14(+)CD16(+)CD64(-) monocyte subset found to be expanded in a variety of pathological conditions. Functionally, DLCs generated by P-selectin in combination with M-CSF plus IL-4 primed naive allogeneic CD4(+) T cells to produce significantly less IFN-gamma than cells generated by BSA in the presence of M-CSF and IL-4. P-selectin effects on enhancing CD14(+)CD16(+) DLC generation were completely abrogated by pretreatment of cells with the protein kinase C delta inhibitor rottlerin, but not by classical protein kinase C inhibitor G?6976. Immobilized P-selectin also inhibited macrophage differentiation in response to M-CSF alone as demonstrated by morphology, phenotype, and phagocytosis analysis. The effects of P-selectin on macrophage differentiation were neutralized by pretreatment of monocytes with Ab against P-selectin glycoprotein ligand 1. These results suggest a novel role for P-selectin in regulating monocyte fate determination.  相似文献   

2.
In human blood two monocyte populations can be distinguished, i.e., the CD14(++)CD16(-)DR(+) classical monocytes and the CD14(+)CD16(+)DR(++) proinflammatory monocytes that account for only 10% of all monocytes. We have studied TNF production in these two types of cells using three-color immunofluorescence and flow cytometry on whole peripheral blood samples stimulated with either LPS or with the bacterial lipopeptide S-(2,3-bis(palmitoyloxy)-(2-RS)-propyl)-N-palmitoyl-(R)-Cys-(S)-Ser-(S)-Lys(4)-OH,trihydrochloride (Pam3Cys). After stimulation with LPS the median fluorescence intensity for TNF protein was 3-fold higher in the proinflammatory monocytes when compared with the classical monocytes. After stimulation with Pam3Cys they almost exclusively responded showing 10-fold-higher levels of median fluorescence intensity for TNF protein. The median fluorescence intensity for Toll-like receptor 2 cell surface protein was found 2-fold higher on CD14(+)CD16(+)DR(++) monocytes, which may explain, in part, the higher Pam3Cys-induced TNF production by these cells. When analyzing secretion of TNF protein into the supernatant in PBMCs after depletion of CD16(+) monocytes we found a reduction of LPS-induced TNF by 28% but Pam3Cys-induced TNF was reduced by 64%. This indicates that the minor population of CD14(+)CD16(+) monocytes are major producers of TNF in human blood.  相似文献   

3.
CD8(+) T cells become exhausted, inducing cell surface protein programmed cell death-1 (PD-1) as chronic virus diseases or tumors progress, but underlying mechanisms of this are unclear. We previously showed that M-CSF is important for developing tolerogenic dendritic cells (DCs) from human CD14(+) monocytes. In this article, we identify M-CSF-derived DCs (M-DCs) after stimulation with IL-10 as myeloid-derived suppressor cells with additional tolerogenic activities to CD8(+) T cells. IL-10 increased PD-1 ligand expression on M-DC, and IL-10-stimulated M-DCs (M-DC/IL-10) induced expression of PD-1 on, and apoptosis of, CD8(+) T cells and phagocytosed CD8(+) T cells. Enhanced phagocytic activity of M-DC/IL-10 required IFN-γ, which further increased PD-1 ligand and PD-2 ligand expression on M-DC/IL-10. IFN-γ-stimulated M-DC/IL-10 cells were phenotypically macrophage-like cells with little or no expression of CD86, a costimulatory molecule, but with high expression levels of CD14, CD200R, and CD80. No phagocytic activity was detected with GM-CSF-derived DCs. We propose that phagocytosis by IFN-γ-stimulated M-DC/IL-10 cells, which may be DCs or, alternatively, a unique subset of macrophages, may be a mechanism by which IFN-γ-producing CD8(+) T cells are tolerized after type 1 immune responses to chronic virus or tumor, and that IFN-γ links effector CD8(+) T cells to their phagocytic clearance.  相似文献   

4.
The inducible costimulatory (ICOS) molecule is expressed by activated T cells and has homology to CD28 and CD152. ICOS binds B7h, a molecule expressed by APC with homology to CD80 and CD86. To investigate regulation of ICOS expression and its role in Th responses we developed anti-mouse ICOS mAbs and ICOS-Ig fusion protein. Little ICOS is expressed by freshly isolated mouse T cells, but ICOS is rapidly up-regulated on most CD4(+) and CD8(+) T cells following stimulation of the TCR. Strikingly, ICOS up-regulation is significantly reduced in the absence of CD80 and CD86 and can be restored by CD28 stimulation, suggesting that CD28-CD80/CD86 interactions may optimize ICOS expression. Interestingly, TCR-transgenic T cells differentiated into Th2 expressed significantly more ICOS than cells differentiated into Th1. We used two methods to investigate the role of ICOS in activation of CD4(+) T cells. First, CD4(+) cells were stimulated with beads coated with anti-CD3 and either B7h-Ig fusion protein or control Ig fusion protein. ICOS stimulation enhanced proliferation of CD4(+) cells and production of IFN-gamma, IL-4, and IL-10, but not IL-2. Second, TCR-transgenic CD4(+) T cells were stimulated with peptide and APC in the presence of ICOS-Ig or control Ig. When the ICOS:B7h interaction was blocked by ICOS-Ig, CD4(+) T cells produced more IFN-gamma and less IL-4 and IL-10 than CD4(+) cells differentiated with control Ig. These results demonstrate that ICOS stimulation is important in T cell activation and that ICOS may have a particularly important role in development of Th2 cells.  相似文献   

5.
In elderly subjects and in patients with chronic inflammatory diseases, there is an increased subset of monocytes with a CD14(+)CD16(+) phenotype, whose origin and functional relevance has not been well characterized. In this study, we determined whether prolonged survival of human CD14(++)CD16(-) monocytes promotes the emergence of senescent cells, and we analyzed their molecular phenotypic and functional characteristics. We used an in vitro model to prolong the life span of healthy monocytes. We determined cell senescence, intracellular cytokine expression, ability to interact with endothelial cells, and APC activity. CD14(+)CD16(+) monocytes were senescent cells with shortened telomeres (215 ± 37 relative telomere length) versus CD14(++)CD16(-) cells (339 ± 44 relative telomere length; p < 0.05) and increased expression of β-galactosidase (86.4 ± 16.4% versus 10.3 ± 7.5%, respectively; p = 0.002). CD14(+)CD16(+) monocytes exhibited features of activated cells that included expression of CD209, release of cytokines in response to low-intensity stimulus, and increased capacity to sustain lymphocyte proliferation. Finally, compared with CD14(++)CD16(-) cells, CD14(+)CD16(+) monocytes showed elevated expression of chemokine receptors and increased adhesion to endothelial cells (19.6 ± 8.1% versus 5.3 ± 4.1%; p = 0.033). In summary, our data indicated that the senescent CD14(+)CD16(+) monocytes are activated cells, with increased inflammatory activity and ability to interact with endothelial cells. Therefore, accumulation of senescent monocytes may explain, in part, the development of chronic inflammation and atherosclerosis in elderly subjects and in patients with chronic inflammatory diseases.  相似文献   

6.
The glycoprotein CD86 is an important costimulatory molecule that has been shown to be predominantly expressed on APCs, such as dendritic cells, macrophages, and B cells. More recently, CD86 was also detected on T cells in specific pathological conditions. The mechanisms of how CD86 might be induced and its functional role in T cells are not well understood. In the present study, we showed that treatment with IL-2 markedly upregulated CD86, but not CD80, in human CD4(+) and CD8(+) T cells. This upregulation occurred in the absence of bystander cells, and isolated naive CD4(+) or CD8(+) T cells exhibited different time-dependent CD86-expression patterns in response to IL-2. Upregulation of CD86 on activated T cells was reduced by Abs that block IL-2 and IL-2Rα (CD25), indicating a receptor-mediated mechanism. IL-2-dependent CD86 upregulation was blocked by pharmacological inhibitors of the NFAT and mammalian target of rapamycin pathways and was largely reduced by simultaneous exposure to IFN-α. Importantly, a marked increase in CD86 on T cells was also observed in vivo in IL-2-treated patients. In conclusion, IL-2 upregulates CD86 expression on human CD4(+) and CD8(+) T cells via a receptor-dependent mechanism that involves the NFAT and mammalian target of rapamycin pathways.  相似文献   

7.
This study looks at immunoincompetent CD4(+) T cells in adult peripheral blood (APB) using cytokine production in response to a superantigen as a measure of function. We compared the function of APB CD38(+)CD4(+) and CD38(-/low)CD4(+) T cells to that of cord blood (CB) CD4(+) T cells. APB CD4(+) T cell blasts produce substantial amounts of IL-2 in response to TSST-1 restimulation, while CB CD4(+) T cell blasts produce less. APB CD38(+)CD4(+) T cells produce low levels of IL-4 and IFN-gamma in response to TSST-1, even after activation, while APB CD38(-/low)CD4(+) T cells retain their ability to produce high levels of these cytokines despite high CD38 expression. These results suggest that the developmental stage of APB CD38(+)CD4(+) T cells lies between that of CB CD4(+) T cells and APB CD38(-/low)CD4(+) T cells and that APB CD38(+)CD45RO(-)CD4(+) T cells gradually cease to express CD38 as they acquire full function. We reconsider CD4(+) cell maturation and response to TSST-1 and discuss the implications of T cell maturity on infectious diseases.  相似文献   

8.
The mechanisms underlying the immunomodulatory functions of mesenchymal stem cells (MSC) on dendritic cells (DC) have been shown to involve soluble factors, such as IL-6 or TGF-beta, or cell-cell contact, or both depending on the report referenced. In this study, we intend to clarify these mechanisms by examining the immunosuppressive effect of human adult MSC on adult DC differentiated from CD34(+) hemopoietic progenitor cells (HPC). MSC have been shown to inhibit interstitial DC differentiation from monocytes and umbilical CD34(+) HPC. In this study, we confirm that MSC not only halt interstitial DC but also Langerhans cell differentiation from adult CD34(+) HPC, as assessed by the decreased expression of CD1a, CD14, CD86, CD80, and CD83 Ags on their cell surface. Accordingly, the functional capacity of CD34(+) HPC-derived DC (CD34-DC) to stimulate alloreactive T cells was impaired. Furthermore, we showed that 1) MSC inhibited commitment of CD34(+) HPC into immature DC, but not maturation of CD34-DC, 2) this inhibitory effect was reversible, and 3) DC generated in coculture with MSC (MSC-DC) induced the generation of alloantigen-specific regulatory T cells following secondary allostimulation. Conditioned medium from MSC cultures showed some inhibitory effect independent of IL-6, M-CSF, and TGF-beta. In comparison, direct coculture of MSC with CD34(+) HPC resulted in much stronger immunosuppressive effect and led to an activation of the Notch pathway as assessed by the overexpression of Hes1 in MSC-DC. Finally, DAPT, a gamma-secretase inhibitor that inhibits Notch signaling, was able to overcome MSC-DC defects. In conclusion, our data suggest that MSC license adult CD34(+) HPC to differentiate into regulatory DC through activation of the Notch pathway.  相似文献   

9.
Despite steady progress in elimination of measles virus globally, measles infection still causes 500,000 annual deaths, mostly in developing countries where endemic measles strains still circulate. Many adults are infected every year in China, with symptoms more severe than those observed in children. In this study, we have used blood samples from adult measles patients in Shanghai and age-matched healthy controls to gain an understanding of the immune status of adult measles patients. IFN-alpha mRNA was reduced in patient PBMC compared with healthy controls. In contrast, gene expression and plasma production of IL-2, IL-10, and IFN-gamma were elevated in patient blood. A similar cytokine profile was observed at early times when cultured PBMC were infected with a clinical isolate of measles virus. In contrast to previous studies in pediatric patients, we did not find a reduction in total CD4(+) and CD8(+) T cells in patient PBMC. Interestingly, we found that CD4(+)CD25(+)CD127(low) regulatory T cells were significantly increased in patient PBMC compared with controls. Using intracellular cytokine staining we also show that the measles virus induces IL-10-producing CD14(+) and CD4(+)CD25(+) cells in PBMC. Our results show that adult measles patients in the acute phase of the disease have a mixed Th1/Th2 type response, accompanied with severe immunosuppression of both innate and adaptive responses including suppression of type I IFN. Both regulatory T cells and plasma IL-10 may contribute to the immunosuppression.  相似文献   

10.
Costimulatory ligands CD80 and CD86 have different binding preferences and affinities to their receptors, CD28 and CTLA-4. Earlier, we demonstrated that CD80 binds to CTLA-4 with higher affinity and has a role in suppressing T cell response. The current study demonstrates that not only did blockade of CD86 upon Ag presentation by bone marrow-derived dendritic cells (DC) to OVA-specific T cells result in induction of hyporesponsive T cells but also that these T cells could suppress the proliferative response of effector T cells. These T cells showed TGF-beta1 on their surface and secreted TGF-beta1 and IL-10 upon restimulation. Although blockade of CTLA-4 and neutralization of IL-10 profoundly inhibited the induction of these TGF-beta1(+) T cells, their ability to suppress the effector T cell proliferation was abrogated by neutralization of TGF-beta1 alone. Induction of TGF-beta1(+) and IL-10(+) T cells was found to be independent of natural CD4(+)CD25(+) regulatory T cells, demonstrating that preferential ligation of CTLA-4 by CD80 induced IL-10 production by effector T cells, which in turn promoted the secretion of TGF-beta1. Treatment of prediabetic NOD mice with islet beta cell Ag-pulsed CD86(-/-) DCs, but not CD80(-/-) DCs, resulted in the induction of TGF-beta1- and IL-10-producing cells, significant suppression of insulitis, and delay of the onset of hyperglycemia. These observations demonstrate not only that CD80 preferentially binds to CTLA-4 but also that interaction during Ag presentation can result in IL-10-dependent TGF-beta1(+) regulatory T cell induction, reinstating the potential of approaches to preferentially engage CTLA-4 through CD80 during self-Ag presentation in suppressing autoimmunity.  相似文献   

11.
We show that there are differences in the soluble factors in cord blood (CB) and adult serum and that these differences play a role in T cell function. Thus, the mitogen and alloantigen-specific proliferative response of adult T cells was enhanced with increasing concentrations of adult serum and CB serum, but to a lesser extent with CB serum. In addition, proliferation of T cells induced by stimulation through the T cell receptor alone (via CD3 stimulation), could be enhanced with adult but not CB serum. However, CB serum enhanced the IL-2-specific proliferative response of pure T cells whereas adult serum did not. To determine whether there was an anti-inflammatory cytokine within CB serum which could induce these results, we assayed our serum samples for anti-inflammatory cytokines. IL-13 could not be detected in any serum sample, whereas IL-10 could be detected in adult but not CB serum (P < 0.002). However, there was a significant difference in the levels of macrophage colony stimulating factor (M-CSF) detected in adult and CB serum samples (P < 0.01). M-CSF was detected in 6/7 CB serum samples (mean +/- SD was 3.8 +/- 2.3 ng/ml) and 0/5 adult serum samples. Furthermore, anti-M-CSF antibody restored the reduced allo-response of T cells incubated in CB serum. Thus, M-CSF may act as a suppressor factor in CB serum. Whether this is sufficient to explain the lack of an allo-response by the foetus to the mother, or the reduced graft-versus-host disease when CB is used instead of bone marrow in stem cell transplantation, is yet to be determined.  相似文献   

12.
Previously, we have shown that priming of therapeutic CD8(+) T cells in tumor vaccine-draining lymph nodes of mice vaccinated with GM-CSF secreting B16BL6 melanoma cells occurs independent of CD4 T cell help. In this study, we examined the contribution of the major costimulatory molecules, CD40 ligand (CD40L), CD80, and CD86, in the priming of CD8(+) T cells. Priming of therapeutic CD8(+) T cells by a GM-CSF-transduced tumor vaccine did not require CD40 and CD40L interactions, as therapeutic T cells could be generated from mice injected with anti-CD40L Ab and from CD40L knockout mice. However, costimulation via either CD80 or CD86 was required, as therapeutic T cells could be generated from mice injected with either anti-CD80 or anti-CD86 Ab alone, but administration of both Abs completely inhibited the priming of therapeutic T cells. Blocking experiments also identified that priming of therapeutic T cells in MHC class II-deficient mice required TNFR and IL-12 signaling, but signaling through CD40, lymphotoxin-betaR, or receptor activator of NF-kappaB was not essential. Thus, cross-priming of therapeutic CD8(+) T cells by a tumor vaccine transduced with GM-CSF requires TNFR, IL-12, and CD28 signaling.  相似文献   

13.
Dendritic cells (DCs) are professional antigen-presenting cells that are required for the initiation of the immune response. DCs have been shown to be generated from CD34+pluripotent hematopoietic progenitor cells in the bone marrow and cord blood (CB), but relatively little is known about the effect of cryopreservation on functional maturation of DCs from hematopoietic stem cells. In this work we report the generation of DCs from cryopreserved CB CD34+cells. CB CD34+cells were cryopreserved at −80°C for 2 days. Cryopreserved CB CD34+cells as well as freshly isolated CB CD34+cells cultured with granulocyte—macrophage colony-stimulating factor (GM-CSF)/stem cell factor (SCF)/tumor necrosis factor-α (TNF-α) for 14 days gave rise to CD1a+/CD4+/CD11c+/CD14/CD40+/CD80+/CD83+/CD86+/HLA-DR+cells with dendritic morphology. DCs derived from cryopreserved CB CD34+cells showed a similar endocytic capacity for fluorescein isothiocyanate-labeled dextran and lucifer yellow when compared with DCs derived from freshly isolated CB CD34+cells. Flow cytometric analysis revealed that two CC chemokine receptors (CCRs), CCR-1 and CCR-3, were expressed on the cell surface of DCs derived from both cryopreserved and freshly isolated CB CD34+cells, and these DCs exhibited similar chemotactic migratory capacities in response to regulated on activation normal T-cell expressed and secreted. DCs derived from cryopreserved as well as freshly isolated CB CD34+cells were more efficient than peripheral blood mononuclear cells in the primary allogeneic T-cell response. These results indicate that frozen CB CD34+cells cultured with GM-CSF/TNF-α/SCF gave rise to dendritic cells which were morphologically, phenotypically and functionally similar to DCs derived from fresh CB CD34+cells.  相似文献   

14.
BACKGROUND: Cynomolgus monkeys (Macaca fascicularis) are widely used animal models in biomedical research. However, the phenotypic characteristics of cynomolgus monkey (CM) B cells in peripheral blood (PB) and lymphoid organs are poorly understood. METHODS: FACS analyses of PB-, spleen-, lymph node (LN)-, and bone marrow (BM)-derived B cells were performed. RESULTS: CM peripheral blood B cells have a smaller fraction of CD27(-) (naive) cells ( approximately 40%), as compared to human blood samples ( approximately 70%). Similar to humans, an early activation marker, CD23, is expressed more on CD27(-) CM naive B cells, as compared to CD27(+) B cells. The mean fraction of B cells exhibiting a memory phenotype is similar to that seen in human blood. Unlike humans, CM blood contains a subset of CD20(++)CD80(+)CD21(-)IgM(+/-)CD27(+)CD19(+)FSC(++)BAFF-R(low) B cells that are likely of germinal center origin. Thus, CM blood contains (i) a higher percentage of B cells that express the co-stimulatory molecule CD80, and (ii) a lower fraction of B cells that are CD21(+), as compared to human blood. Due to the relative paucity of information on B-cell subsets in organs of healthy humans, a direct comparison between human and CM lymphoid organ data is limited. The fraction of CD27(+) and CD23(+) B cells appears to be similar, while the fraction of CD80(+) B cells appears to be higher than that seen in human lymphoid organs. CM spleens and to some extent lymph nodes have a distinct subset of CD21(++) cells that are also CD80(+/-)CD23(low)IgM(++)CD27(+/-)FSC(++). This subset is phenotypically similar to the marginal zone B cells present in human spleen and LN samples. We also provide detailed analyses on the fraction of lymphoid organ B cells that express CD21, CD23, CD32, and/or CD80 B-cell markers. CONCLUSIONS: In general, cynomolgus monkey B-cell subsets are similar to those seen in humans, as well as to those seen in other nonhuman primates. However, there are some clear differences between human and cynomolgus monkey B-cell subsets. These findings have direct implications for a variety of in vivo studies in cynomolgus monkeys, ranging from basic research on primate B-cell differentiation to models of infectious diseases and trials of new B-cell targeting therapeutic agents.  相似文献   

15.
Dipeptidylpeptidase IV (CD26) is a multifunctional ectoenzyme involved in T cell activation that has been implicated in autoimmune pathophysiology. Because IL-17-producing CD4(+) T cells (Th17 cells) are important mediators of autoimmune disease, we analyzed the expression of CD26 and its enzymatic function on human Th17 cells. Analysis of CD26 expression on different CD4(+) T helper subsets showed that CD26 expression is highest on CD4(+) T cells producing type 17 cytokines (e.g., IL-22, IL-17, GM-CSF, or TNF) compared with Th1, Th2, and regulatory T cells. Phenotypic analysis revealed that CD26(++)CD4(+) T cells express the type 17 differentiation molecules CD161, CCR6, lL-23R, and retinoic acid-related orphan receptor-γt. Furthermore, sorted CD26(++)CD4(+) T cells contain >90-98% of Th17 cells, indicating that CD26(++) T cells harbor the Th17 lineage. A comparison with CD161 and CCR6 indicated that analysis of CD26 coexpression may improve the phenotypic characterization of Th17 cells. Of note, CD26(++) Th17 cells are enriched in the inflamed tissue of patients with hepatitis and inflammatory bowel disease. Functional analysis in migration assays revealed that CD26 expressed on Th17 cells is enzymatically active. Indeed, CD26 negatively regulates the chemotactic CD4(+) T cell response to the inflammatory chemokines CXCL9-12 that can be restored by pharmacological blockade of the enzymatic center of CD26. In summary, these results strongly suggest that CD26 may contribute to the orchestration of the immune response by Th17 cells in human inflammatory diseases. They also suggest that the phenotypic analysis of Th17 cells may be facilitated by determination of CD26 expression.  相似文献   

16.
Several studies have provided evidence that activation of antigen-specific T cells requires interactions between CD28 on T cells and its ligands, CD80 and CD86, on antigen-presenting cells (APCs). However, the effects of CD80 and CD86 on cytokine production in patients with Hymenoptera venom allergy who receive venom immunotherapy remain unclear. We examined the effects of CD80 and CD86 on Th1- and Th2-cytokine production before and after venom immunotherapy in patients with wasp-venom allergy. Peripheral blood mononuclear cells (PBMCs) were isolated from patients with wasp-venom allergy before and after three months of venom immunotherapy. CD4+ T cells and monocytes were isolated as APCs from PBMCs and were cocultured with wasp venom in the presence of anti-CD80 or -CD86 blocking antibodies. Interleukin (IL)-4, IL-10, and interferon (IFN)-gamma were measured by enzyme-linked immunosorbent assay. The expression of CD80 and CD86 on CD14+ PBMCs was detected by fluorescence-activated cell-sorter analysis. The expression of CD86, but not that of CD80, on CD14+ PBMCs cocultured with venom increased after three months of venom immunotherapy, but not before venom immunotherapy. Blockade of CD86 reduced IL-10 production after three months of venom immunotherapy. IL-10 production promoted by CD86 costimulation may be involved in the mechanism of venom immunotherapy in patients with venom allergy.  相似文献   

17.
The CD28 ligands CD80 and CD86 are expressed on APC, and both provide costimulatory function. However, the reason for the expression of two separate CD28 ligands remains unclear. We have previously shown that blockade of CD80 costimulation by Y100F-Ig, a CTL-associated Ag-4 (CTLA4)-Ig mutant that does not bind CD86, inhibits the development of lung inflammatory immune responses, but does not affect blood eosinophilia or Ab production. Each of those responses was inhibited by treatment with CTLA4-Ig, which binds both CD80 and CD86. To clarify the mechanism underlying these observations we have developed a model of lung inflammation using adoptively transferred CD4(+) T cells expressing a Valpha11(+)Vbeta3(+) transgenic TCR specific for I-E(k) and moth cytochrome c. Treatment with Y100F-Ig inhibited the induction of lung eosinophilia in adoptively transferred mice. However, Y100F-Ig did not detectably affect the accumulation of Ag-specific T cells at the site of peptide deposit or in the draining lymphoid tissues. Acquisition of an activated phenotype and expression of adhesion molecules required for migration into the lung were modestly affected. Importantly, treatment with Y100F-Ig diminished the ability of T cells to produce the cytokines IL-4 and IL-5 following intranasal challenge with Ag. All the responses examined were severely inhibited by treatment with CTLA4-Ig. We conclude that T cells require CD80 costimulation for the optimal production of IL-5 following intranasal administration of Ag. Decreased IL-5 production is the most likely explanation for the diminished airway eosinophilia observed.  相似文献   

18.
19.
Mamoni RL  Blotta MH 《Cytokine》2006,35(3-4):207-216
Human infection with Paracoccidioides brasiliensis may result in three major outcomes: paracoccidioidomycosis-infection (PI), which is observed in healthy carriers living in endemic areas and the adult form (AF) and juvenile form (JF) of the disease. In this study we proposed to examine the intracellular expression of IFN-gamma, TNF-alpha, IL-2, IL-10, IL-12, CXCL8, CXCL9 and CXCL10 by peripheral blood mononuclear cells (PBMC) of patients with the JF and AF of the disease, as well as of PI individuals stimulated with PMA plus ionomycin, LPS or anti-CD3 plus anti-CD28, by flow cytometry. The results showed that PI individuals present a higher percentage of cells producing IFN-gamma, TNF-alpha, IL-2, CXCL9 and CXCL10 when compared to AF and JF patients. IFN-gamma was predominantly detected in CD3(+)CD8(+) T cells, whereas IL-2 and TNF-alpha were mainly expressed in CD3(+)CD4(+) cells. Monocytes of PI individuals also presented higher expression of CD80 and lower expression of CD86 when compared to JF and AF patients, and higher expression of HLA-DR, only when compared to JF patients. These results indicate that the differential production of cytokines and chemokines, as well as the expression of co-stimulatory molecules involved in antigen presentation, may influence the outcome of PCM infection.  相似文献   

20.
Massive infection of memory CD4 T cells is a hallmark of early simian immunodeficiency virus (SIV) infection, with viral infection peaking at day 10 postinfection (p.i.), when a majority of memory CD4 T cells in mucosal and peripheral tissues are infected. It is not clear if mononuclear cells from the monocyte and macrophage lineages are similarly infected during this early phase of explosive HIV and SIV infections. Here we show that, at day 10 p.i., Lin(-) HLA-DR(+) CD11c/123(-) CD13(+) CD14(-) macrophages in the jejunal mucosa were infected, albeit at lower levels than CD4 memory T cells. Interestingly, Lin(-) HLA-DR(+) CD11c/123(-) CD13(+) CD14(-) macrophages in peripheral blood, like their mucosal counterparts, were preferentially infected compared to Lin(-) HLA-DR(+) CD11c/123(-) CD13(+) CD14(+) monocytes, suggesting that differentiated macrophages were selectively infected by SIV. CD13(+) CD14(-) macrophages expressed low levels of CD4 compared to CD4 T cells but expressed similar levels of CCR5 as lymphocytes. Interestingly, CD13(+) CD14(-) macrophages expressed Apobec3G at lower levels than CD13(+) CD14(+) monocytes, suggesting that intracellular restriction may contribute to the differential infection of mononuclear subsets. Taken together, our results suggest that CD13(+) CD14(-) macrophages in mucosal and peripheral tissues are preferentially infected very early during the course of SIV infection.  相似文献   

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