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.
CD80 and CD86 seem to play an important role in the allergen-induced secretion of interleukin (IL)-5 and IL-13. Up to now, the expressions of CD80 (B7.1) and CD86 (B7.2) on monocytes and the kinetics of the expression of these molecules on lipopolysaccharide-stimulated monocytes in nonatopic asthma have not been defined. Using monoclonal antibodies, we have compared the expressions of CD80 (B7.1) and CD86 (B7.2) on the monocytes of healthy persons and nonatopic asthmatic patients. We have also assessed the effect of CD80 and CD86 inactivation on IL-4 and interferon (IFN)-gamma production in nonatopic asthmatics and healthy subjects. We found that a low expression of CD80 (1.64 +/- 0.65 vs. 3.53 +/- 1.43%) and a moderate expression of CD86 (41.25 +/- 13.4 vs. 49.46 +/- 11.49%) on the studied monocytes were characteristic for asthma. In nonatopic asthma patients inactivation of CD80 or CD86 blockade significantly reduced IFN-gamma production by T lymphocytes (p < 0.02; p < 0.03). In both the studied groups, anti-CD80 antibodies did not diminish T lymphocyte production of IL-4. However, anti-CD86 antibodies significantly (p < 0.04) reduced the IL-4 concentration in culture supernatants. Our results confirm that both the CD80 and CD86 molecules play an important role in the maintenance and amplification of the inflammatory process. It suggests that in the inflammatory process that occurs in nonatopic bronchial asthma, Th1 as well as Th2 lymphocytes are equally important.  相似文献   

2.
3.
The ability of innate immune cells to differentially respond to various bacterial components provides a mechanism by which the acquired immune response may be tailored to specific pathogens. The response of innate immune cells to bacterial components provides regulatory signals to cognate immune cells. These signals include secreted cytokines and costimulatory molecules, and to a large extent they determine the quantitative and qualitative nature of the immune response. In order to determine if innate immune cells can differentially respond to bacterial components, we compared the responses of macrophages to two bacterially derived molecules, cholera toxin (CT) and lipopolysaccharide (LPS). We found that CT and LPS differentially regulated the expression of interleukin-12 (IL-12) and CD80-CD86 but not that of IL-1beta. LPS and CT each induced IL-1beta expression in macrophages, while only LPS induced IL-12 and only CT induced CD80-CD86. These differences were markedly potentiated in gamma interferon (IFN-gamma)-treated macrophages, in which LPS potently induced IL-12 and CD80-CD86 expression. In contrast, IFN-gamma treatment had no effect on the expression of IL-1beta. These results define a molecular basis for the differential pathogenicities of bacterial toxins and are relevant to the design of vaccine adjuvants able to selectively induce desired types of immunity.  相似文献   

4.
In addition to T cell receptor triggering, activation of T cells requires co-stimulatory signals that have been shown to be mainly initiated through CD28. We analyzed the expression and function of the two ligands for CD28, B7-1 (CD80) and B7-2 (CD86), on human Langerhans cells (LC), the antigen-presenting cells from epidermis. Human LC freshly isolated from epidermis (fLC) expressed significant level of B7-2, which was increased upon a short culture in vitro. In contrast, B7-1 was undetectable on fLC but appeared at the cell surface after a 3-day culture in vitro. Pre-incubation of 18-h cultured LC with anti-B7-2 monoclonal antibodies (mAb) was sufficient to abrogate the binding of CTLA4-Ig fusion protein, while a combination of both mAb against B7-1 and B7-2 was necessary to obtain a complete inhibition of CTLA4-Ig binding on 3-day cultured LC, showing the absence of a third CTLA4 ligand. The function of B7-1 and B7-2 on human LC has been analyzed by adding mAb at the beginning of mixed epidermal cell lymphocyte reactions. Anti-B7-2 mAb and CTLA4-Ig, but not anti-B7-1 mAb, strongly inhibited allogeneic, as well as recall antigen-induced T cell proliferation supported by fLC or 3-day cultured LC. Collectively, these results demonstrate that B7-2 is the major ligand for CD28/CTLA4 at the LC surface and that it plays a crucial role in human LC co-stimulatory function with little, if any, dependence on B7-1 expression.  相似文献   

5.
6.
In contrast to the inhibitory pathway of NK cell regulation, much less is known about stimulatory or activation signals in NK cells. Both CD80 and CD86 function as costimulatory molecules in T-cell cytotoxicity. Several previous reports, most of them in the murine system, have indirectly or directly indicated the possible role of B7 molecules (CD80 and CD86) triggering NK cell-mediated cytotoxicity in vitro. Nevertheless, only little is known about the role of these molecules on human target cells. Therefore, anti-CD80 and anti-CD86 mAbs were used in blocking experiments and both were shown to inhibit lysis by human NK cells. The degree of inhibition observed was variable. 64% of these NK clones were strongly inhibited by both anti-CD80 and anti-CD86 (Type 1). A small number (19%) were only moderately inhibited by both of these antibodies (Type 2), and 17% of these NK clones were inhibited strongly by anti-CD86 but weakly or not at all by anti-CD80 (Type 3). To further examine the importance of these proteins, B7.1 (CD80) and B7.2 (CD86) genes were transfected into the mouse mastocytoma P815 cell line that could not be killed by the human NK cells. These transfectant cell lines were then tested in cytotoxicity assays using a number of human NK lines. Expression of the CD80 and CD86 molecules resulted in enhanced lysis of P815 by most of the NK lines tested. Thus, both CD80 and CD86 molecules are involved in triggering of human NK cells.  相似文献   

7.
Phytohemagglutinin (PHA)-derived T lymphoblasts or T cell clones from patients genetically deficient in IL-12R beta 1 (IL-12R beta 1(-/-)) or IFN-gamma R1 (IFN-gamma R1(-/-)) produced two- to threefold reduced IFN-gamma levels compared to the corresponding cells from healthy individuals after anti-CD3 and PMA stimulation. Moderate IFN-gamma production was observed in PHA-derived T lymphoblasts or T cell clones derived from healthy subjects in the presence of anti-IFN-gamma R1 or anti-IL-12 mAb, whereas it was negligible in the presence of both mAb. However, when anti-IFN-gamma R1 and/or anti-IL-12 mAb were added during restimulation, the cells produced normal levels of IFN-gamma, indicating that both IFN-gamma and IL-12 had an effect on the priming phase. Moderate production of IFN-gamma was partially enhanced only in IFN-gamma R1(-/-) T cell clones generated in the presence of IL-12, but was almost completely abolished when IL-12R beta 1(-/-) and IFN-gamma R1(-/-) T cell clones were generated in the presence of anti-IFN-gamma R1 or anti-IL-12 mAb, respectively. IL-4 production was enhanced in T cell clones from IL-12R beta 1(-/-),but not from IFN-gamma R1(-/-) patients, whereas IL-10 and IL-2 production did not differ significantly in polyclonal T cells or clones from healthy and deficient individuals. These results indicate that IL-12R beta 1- and IFN-gamma R1-dependent signals co-ordinately regulate IFN-gamma, but not IL-2 and IL-10 production, whereas only IL-12 negatively controls IL-4 production by in vitro-generated T cell clones. Thus, although IL-12 and IFN-gamma signals are each sufficient for moderate production of IFN-gamma by human T cells, both are needed for optimal IFN-gamma production, and in the absence of both IFN-gamma production is completely abrogated.  相似文献   

8.
IL-12 is a potent inducer of IFN-gamma production and Th1 responses. Co-stimulation mediated by B7 has been shown to synergize with IL-12 for optimal IFN-gamma production and proliferation in vitro. In this study, we examined the requirement of CD28/B7 interactions for optimal induction of IL-12 receptor(R) beta1 and beta2 expression and IFN-gamma. IL-12-induced IFN-gamma production and STAT4 nuclear translocation were markedly reduced in CD28(-/-) splenocytes compared to that of wild-type (WT) splenocytes. Analysis of IL-12R expression revealed that IL-12 induced similar levels of IL-12R beta2 mRNA expression in WT and CD28(-/-) cells. In contrast, IL-12R beta1 expression was impaired in CD28(-/-) cells, indicating that expression of IL-12R beta1 and beta2 is differentially regulated by CD28. CD28(-/-) CD4(+) but not CD8(+) cells exhibited a defect in IL-12Rbeta1 expression that was associated with a marked decrease in IL-12 binding as well as IL-12-induced IFN-gamma production. IL-2 could restore IL-12R expression to CD28(-/-) CD4(+) cells, however, this occurred independently of IL-2-induced proliferation. Thus, these findings identify distinct requirements for CD28 in the capacity of CD4(+) and CD8(+) cells to respond maximally to IL-12.  相似文献   

9.
CD80 and CD86 have the capacity to either stimulate or inhibit T cell responses through their receptors CD28 and cytotoxic T lymphocyte-associated antigen 4 (CTLA-4). Blockade of CD80 and CD86 in autoimmune disease settings has revealed distinct outcomes, yet the differential functions of CD80 and CD86 are still unclear. We have studied the ability of individual ligands to stimulate primary responses in human CD4(+) T cells. Our data reveal both quantitative and qualitative differences between the ligands. Both CD80 and CD86 demonstrated the capacity to costimulate T cell proliferation. However, CD80 committed a greater number of T cells to divide with faster kinetics, consistent with it being a superior ligand for CD28. Once cell division had been initiated, all T cells undergoing cell division expressed CTLA-4, irrespective of whether CD80 or CD86 costimulation was used. However, only in the presence of CD80 was evidence of CTLA-4 engagement and inhibitory function observed. Finally, differences between CD80 and CD86 costimulation extended to the T cell phenotype, in particular the levels of CD40 ligand expression.  相似文献   

10.
In order to examine the immune response at the site of pathology in tuberculosis, we analysed cytokines present in lung granulomas, their associations with each other and with caseous necrosis as well as the phenotype of the cellular infiltrate. Paraffin-embedded tissue from the lungs of seven patients with pulmonary tuberculosis was analysed by immunohistochemistry and in situ hybridization to detect interferon-gamma (IFN-gamma), tumour necrosis factor-alpha (TNF-alpha) and interleukin-4 (IL-4) proteins and IL-12p40 mRNA. All seven patients had granulomas staining positive for IFN-gamma, TNF-alpha and IL-12p40, but only four stained positive for IL-4. Cells with the morphology of lymphocytes, macrophages and giant cells expressed TNF-alpha, IFN-gamma and IL-4 protein. Furthermore, CD68-positive myeloid cells expressed IL-12p40 mRNA, as expected, but a subset of CD3-positive lymphocytes also expressed this mRNA. These lymphocytes producing IL-12p40 also stained positive for CD8 but not CD4. A total of 141 granulomas were scored for the presence or absence of cytokine or necrosis and two major associations were identified. The first association was between IFN-gamma and IL-12, with 76% of granulomas staining positive for both cytokines. Unexpectedly, those granulomas positive for IL-4 were always positive for IFN-gamma. The second association was between TNF-alpha and caseous necrosis, where all necrotic granulomas were TNF-alpha positive. This association was modulated by IL-4. Therefore, heterogeneity of cellular infiltrate and cytokine expression is observed between adjacent granulomas in the same patient.  相似文献   

11.
IL-18 is a proinflammatory cytokine that plays an important role in NK cell activation and T(h)1 response. IL-18 has a structural homology to IL-1, particularly IL-1beta. IL-18R, composed of IL-1R-related protein (IL-18Ralpha) and IL-1R accessory protein-like (IL-18Rbeta), belongs to the IL-1R family. Furthermore, IL-18R at least partly shares the signal transducing system with IL-1R. Thus, the IL-18-IL-18R system has a striking similarity to the IL-1-IL-1R system. For this reason, we regarded it important to investigate whether, like IL-18, IL-1beta synergizes with IL-12 in inducing IFN-gamma production from human T cells and plays an important role in the T(h)1 response. Here we show that IL-12 and IL-1beta synergistically induce T cells to proliferate and produce IFN-gamma without their TCR engagement. IL-12 stimulation induced an increase in the proportion of T cells positive for IL-18R. Then, IL-12-stimulated T cells responded to IL-18 or IL-1beta by their proliferation and IFN-gamma production, although levels of IL-1beta-induced responses were lower. CD4(+)CD45RA(+) T cells, although they constitutively expressed IL-18Rbeta mRNA, did not express IL-18Ralpha mRNA. Phytohemagglutinin (PHA) stimulation alone induced IL-18Ralpha mRNA without affecting the expression of IL-18Rbeta mRNA. T(h)1-inducing conditions (PHA, IL-12 and anti-IL-4) further increased this expression. We also show that T(h)1 cells but not T(h)2 cells have increased expression of IL-18R and IL-1R, and produce IFN-gamma in response to IL-18 and/or IL-1beta.  相似文献   

12.
Trogocytosis is a process which involves the transfer of membrane fragments and cell surface proteins between cells. Various types of T cells have been shown to be able to acquire membrane-bound proteins from antigen-presenting cells and their functions can be modulated following trogocytosis. However, it is not known whether induced regulatory T cells (iTregs) can undergo trogocytosis, and if so, what the functional consequences of this process might entail. In this study, we show that iTregs can be generated from CD80−/−CD86−/− double knockout (DKO) mice. Using flow cytometry and confocal fluorescence microscopy, we demonstrate that iTregs generated from DKO mice are able to acquire both CD80 and CD86 from mature dendritic cells (mDCs) and that the acquisition of CD86 occurs to a higher extent than that of CD80. Furthermore, we found that after co-incubation with iTregs, dendritic cells (DCs) downregulate their surface expression of CD80 and CD86. The trogocytosis of both CD80 and CD86 occurs in a cytotoxic T lymphocyte-associated antigen-4 (CTLA-4), CD28 and programmed death ligand-1 (PDL1)-independent manner. Importantly, we showed that iTregs that acquired CD86 from mDCs expressed higher activation markers and their ability to suppress naive CD4+ T-cell proliferation was enhanced, compared to iTregs that did not acquire CD86. These data demonstrate, for the first time, that iTregs can acquire CD80 and CD86 from mDCs, and the acquisition of CD86 may enhance their suppressive function. These findings provide novel understanding of the interaction between iTregs and DCs, suggesting that trogocytosis may play a significant role in iTreg-mediated immune suppression.  相似文献   

13.
CD28 induces immunostimulatory signals in dendritic cells via CD80 and CD86   总被引:11,自引:0,他引:11  
Bidirectional signaling along the B7-CTLA-4 coreceptor pathway enables reciprocal conditioning of T cells and dendritic cells. Although T cells can instruct dendritic cells to manifest tolerogenic properties after CTLA-4 engagement of B7, such a B7-mediated signaling is not known to occur in response to CD28. Here we show that mouse dendritic cells were induced by soluble CD28 to express interleukin 6 and interferon-gamma. Production of interleukin 6 required B7-1 (CD80), B7-2 (CD86) and p38 mitogen-activated protein kinase and prevented interferon-gamma-driven expression of immunosuppressive tryptophan catabolism. In vivo, an adjuvant activity of soluble CD28 was demonstrated as enhanced T cell-mediated immunity to tumor and self peptides and protection against microbial and tumor challenge. Thus, different ligands of B7 can signal dendritic cells to express functionally distinct effector responses.  相似文献   

14.
Dendritic cells (DC) are potent antigen-presenting cells derived from CD34 bone marrow stem cells. They undergo a series of maturational steps that allow them to stimulate primary T cell responses. Several cytokines are known to contribute to this process. In this study murine DC maturing from bone marrow progenitors under the influence of granulocyte macrophage colony stimulating factor and tumour necrosis factor-alpha were found to produce IL-12 as measured by ELISA and by flow cytometry to detect intracellular cytokine. Administration of additional IL-12 from day 3 to 7 of culture altered the function and phenotype of DC; enhanced stimulation of T cell proliferation by DC in allogeneic mixed leukocyte reactions was associated with an increase in the surface expression of CD80 on DC. These effects were dose dependent, and were consistently seen with IL-12 at 25 ng/ml and were less marked with IL-12 at 50 ng/ml. These results show that IL-12 is both produced by DC and can increase their stimulatory capacity. The findings suggest that there may be an autocrine effect of IL-12 on DC maturation and function.   相似文献   

15.
The role of CD8+ T cells in adaptive immunity is well documented and involves numerous effector mechanisms including direct cytolysis of targets and secretion of cytokines. The role of CD8+ T cells in innate immunity has not been previously appreciated. Using J774 macrophages infected in vitro with the intracellular bacterium, Listeria monocytogenes (LM), we show that CD8+ T cells isolated from na?ve C57BL/6 (B6) mice respond rapidly by secreting IFN-gamma. CD8+ T cells secreting IFN-gamma can also be found in na?ve B6 mice 16 h after infection with LM. This rapid IFN-gamma response is TCR-independent and mediated through the actions of IL-12 and IL-18. Cell surface staining and cell sorting experiments indicate that these novel CD8+ T cells express memory markers. In vitro CFSE-labeling experiments show that IFN-gamma-secreting CD8+ T cells proliferate rapidly after 2 days in culture and after 4 days constitute the majority of the CD8+ T cell population. Together, these data suggest an important role for IFN-gamma-secreting CD8+ T cells in the innate response to bacterial pathogens.  相似文献   

16.
17.
Although CD28 triggering provides an important co-stimulatory signal to T cells, blocking the CD80/CD86 - CD28 interaction with CTLA-4lg fusion protein is not sufficient for tolerance induction in vivo or in vitro. According to more recent data, interruption of the CD40 - CD154 interaction might complement the effect of CTLA-4lg and induce graft acceptance. We studied the effects of a blocking anti-CD40 monoclonal antibody (mAb) and/or blocking anti-CD80/anti-CD86 mAb in cultures of human peripheral blood mononuclear cells (PBMC) stimulated with allogeneic PBMC. T cells activated by alloantigens in the presence of anti-CD80, anti-CD86 and anti-CD40 entered a state of alloantigen-specific non-responsiveness as evidenced upon restimulation by lack of proliferation, cytotoxic activity, and IL-2, IL-5 and IL-13 production. IFN-gamma production during restimulation was less than in the control cultures, while the production of IL-10 was enhanced. Addition of recombinant IL-2 during the restimulation rescued alloantigen-specific activity. We conclude that the simultaneous blocking of the CD40 - CD154 and CD80/CD86 - CD28 interaction during allogeneic T cell activation induces T cell anergy. Since anergic cells induced by this treatment still produce high levels of IL-10, the latter could contribute to modulation of antigen-presenting cell activity and to bystander suppression of residual reactive T cells.  相似文献   

18.
CD30 is expressed on activated T cells that, as has been suggested, preferentially produce IFN-gamma. Interleukin 12 increases antigen-induced CD30 expression on T cells and IFN-gamma production. Synthesis of IFN-gamma can be augmented further by IL-18. The aim of our study was to investigate whether IL-18 affects the IL-12 induced CD30 expression and cytokine production by allergen or PPD specific T cells. Mononuclear cells of healthy or atopic volunteers were stimulated with PPD or allergen, respectively, to obtain specific T cell lines. T cells were restimulated with appropriate antigen and antigen-presenting cells in the presence of IL-12, IL-18 or a combination of these cytokines. After 3 days, expression of CD30 was investigated on CD4 and CD8 T cells and IFN-gamma and IL-4 cytokine production was estimated in the culture supernatants. Flow cytometric analyses showed no effect of IL-18 on CD30 expression during IL-12 co-stimulation. At the same time after the optimal stimulation for CD30 expression, the levels of IFN-gamma were high in PPD-stimulated cell lines but have not been up-regulated by IL-18. IFN-gamma levels were much lower in allergen-stimulated T cells and although they were up-regulated by IL-12 there was no additional or synergistic effect from IL-18. IL-18, however, increased production of IL-4 in allergen-stimulated cell lines. Our studies provide new information about IL-18 activity on human cells and question its exclusive role in Th1 mediated responses.  相似文献   

19.
We investigated the efficacy of a combination of anti-CD80 and CD86 (CD80 + 86) monoclonal antibodies (mAb), anti-interleukin (IL)-12 mAb, or both, for prophylaxis in a mouse acute graft-versus-host-disease (GVHD) model. The treatment with a combination of anti-CD80 + 86 mAb efficiently reduced the lethality of GVHD, whereas mAb against either CD80 or CD86 alone had an effect. A delay in lymphocyte reconstitution and GVHD-associated histological changes in organs was observed at 30 days post-bone marrow transplantation (BMT) even in the anti-CD80 + 86 mAb-treated mice, although these manifestations were resolved by 100 days. In vitro, host alloantigen-specific T cell proliferative responses and generation of CTL were significantly reduced by anti-CD80 + 86 treatment. Furthermore, anti-CD80 + 86 mAb preferentially inhibited the production of interferon (IFN)-γ, but not IL-4 and IL-10, when cultures were assayed at 21 days. Although the anti-IL-12 mAb treatment alone inhibited the generation of cytotoxic T lymphocytes and IFN-γ production in vitro, administration of anti-IL-12 mAb in vivo reversed the beneficial effects of anti-CD80 + 86 treatment on host survival post-BMT. The adverse effect of anti-IL-12 treatment seems to result from impairment of natural immunity and hematopoiesis, rather than as a consequence of an incomplete blockade of T helper (Th)1 responses. Our results suggest that the prevention of GVHD-induced death results from the efficient blockade of Th1 cell activation by the anti-CD80 + 86 treatment. However, further treatment is required for a complete prevention of GVHD, which seems to be partly mediated by Th2 cells.  相似文献   

20.
Langerhans cells (LCs) represent an immature population of myeloid dendritic cells (DCs). As a result of their unique Birbeck granules (BGs), langerin expression, and heterogeneous maturation process, they differ from other immature DCs. Monocyte-derived LCs (MoLCs) mimic epidermal LCs. MoLCs with characteristic BGs are generated by culturing blood-derived monocytes with granulocyte macrophage-colony stimulating factor, interleukin (IL)-4, and transforming growth factor-beta1. Here, we compare maturation-induced antigen expression and cytokine release of LCs with MoLCs. To achieve comparable cell populations, LCs and MoLCs were isolated by CD1c cell sorting, resulting in high purity. In unstimulated cells, CD40 was expressed at equal levels. After stimulation with CD40 ligand (CD40L), LCs and MoLCs acquired CD83 and increased CD86. High CD80 expression was exclusively detected in CD1c-sorted MoLCs. Human leukocyte antigen-DR and CD54 expression was found in all cell populations, however, at different intensities. CD40 triggering increased the potency of LCs and MoLCs to stimulate CD4+ T cell proliferation. Activated MoLCs released IL-12p70 and simultaneously, anti-inflammatory IL-10. The application of the Toll-like receptor ligands peptidoglycan, flagellin, and in particular, lipopolysaccharide (LPS) increased the corelease of these cytokines. LCs secreted IL-10 at a comparable level with MoLCs but failed to produce high amounts of IL-12p70 after application of danger signals. These data indicate that MoLCs as well as LCs display no maturation arrest concerning CD83 and CD86 expression. In difference to MoLCs, LCs resisted activation by CD40L and LPS in terms of IL-12 production. This shows that natural and generated LCs share similar features but differ in relevant functions.  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