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.
Rheumatoid arthritis (RA) is characterized by premature immune aging with accumulation of degenerate T cells deficient for CD28. Gene expression profiling of CD4(+)CD28(-) and CD4(+)CD28(+) T cells to discover disease-promoting activities of CD28(-) T cells identified expression of CD70 as a most striking difference. Hence, CD70 was significantly more expressed in CD4 T cells from RA patients compared with age-matched controls (p < 0.006). The underlying mechanism was a failure to repress CD70 expression after activation-dependent induction. This defect in RA was not related to differential promoter demethylation. CD70 on bystander CD4(+)CD28(-) T cells functioned by lowering the threshold for T cell activation; admixture of CD4(+)CD28(-) T cells augmented TCR-induced responses of autologous naive CD4(+)CD28(+) T cells, particularly of low-avidity T cells. The data support a model in which CD70 expressed on T cells causes degeneracy in T cell responses and undermines tolerance mechanisms that normally control T cell autoreactivity.  相似文献   

2.
IL-2 and IL-15 are lymphocyte growth factors produced by different cell types with overlapping functions in immune responses. Both cytokines costimulate lymphocyte proliferation and activation, while IL-15 additionally promotes the development and survival of NK cells, NKT cells, and intraepithelial lymphocytes. We have investigated the effects of IL-2 and IL-15 on proliferation, cytotoxicity, and cytokine secretion by human PBMC subpopulations in vitro. Both cytokines selectively induced the proliferation of NK cells and CD56(+) T cells, but not CD56(-) lymphocytes. All NK and CD56(+) T cell subpopulations tested (CD4(+), CD8(+), CD4(-)CD8(-), alphabetaTCR(+), gammadeltaTCR(+), CD16(+), CD161(+), CD158a(+), CD158b(+), KIR3DL1(+), and CD94(+)) expanded in response to both cytokines, whereas all CD56(-) cell subpopulations did not. Therefore, previously reported IL-15-induced gammadelta and CD8(+) T cell expansions reflect proliferations of NK and CD56(+) T cells that most frequently express these phenotypes. IL-15 also expanded CD8alpha(+)beta(-) and Valpha24Vbeta11 TCR(+) T cells. Both cytokines stimulated cytotoxicity by NK and CD56(+) T cells against K562 targets, but not the production of IFN-gamma, TNF-alpha, IL-2, or IL-4. However, they augmented cytokine production in response to phorbol ester stimulation or CD3 cross-linking by inducing the proliferation of NK cells and CD56(+) T cells that produce these cytokines at greater frequencies than other T cells. These results indicate that IL-2 and IL-15 act at different stages of the immune response by expanding and partially activating NK receptor-positive lymphocytes, but, on their own, do not influence the Th1/Th2 balance of adaptive immune responses.  相似文献   

3.
4.
T cells specific for persistent pathogens accumulate with age and express markers of immune senescence. In contrast, much less is known about the state of T cell memory for acutely infecting pathogens. Here we examined T cell responses to influenza in human peripheral blood mononuclear cells from older (>64) and younger (<40) donors using whole virus restimulation with influenza A (A/PR8/34) ex vivo. Although most donors had pre-existing influenza reactive T cells as measured by IFNγ production, older donors had smaller populations of influenza-responsive T cells than young controls and had lost a significant proportion of their CD45RA-negative functional memory population. Despite this apparent dysfunction in a proportion of the older T cells, both old and young donors' T cells from 2008 could respond to A/California/07/2009 ex vivo. For HLA-A2+ donors, MHC tetramer staining showed that a higher proportion of influenza-specific memory CD8 T cells from the 65+ group co-express the markers killer cell lectin-like receptor G1 (KLRG1) and CD57 compared to their younger counterparts. These markers have previously been associated with a late differentiation state or immune senescence. Thus, memory CD8 T cells to an acutely infecting pathogen show signs of advanced differentiation and functional deterioration with age. There was a significant negative correlation between the frequency of KLRG1(+)CD57(+) influenza M1-specific CD8 T cells pre-vaccination and the ability to make antibodies in response to vaccination with seasonal trivalent inactivated vaccine, whereas no such trend was observed when the total CD8(+)KLRG1(+)CD57(+) population was analyzed. These results suggest that the state of the influenza-specific memory CD8 T cells may be a predictive indicator of a vaccine responsive healthy immune system in old age.  相似文献   

5.
Recent evidence suggests that suppression of the cellular immune response is often attributable to populations of functionally distinct T cells that act to down-regulate Ag-specific effector T cells. Using flow cytometry, we evaluated tumor-infiltrating lymphocytes (TIL) from patients undergoing neurosurgical resection of glioblastoma multiforme (GBM), metastatic lung carcinoma, and meningioma for markers known to be expressed on immunoregulatory T cells. Ex vivo phenotypic characteristics, cellular proliferation, and cytokine expression patterns were compared between T cell subsets found in the PBMC and within TIL from fresh tumor samples. Interestingly, nearly half of all T cells infiltrating GBM specimens were CD56(+) T cells, while much smaller percentages of similar cells were identified within metastatic lung tumors and meningiomas. CD56(+) T cells identified within GBM were not canonical, or "invariant," NKT cells, as they demonstrated diverse TCR expression, a primarily CD4 single-positive phenotype, and lack of CD1d reactivity. The percentage of CD56(+) T cells exhibiting evidence of proliferation within GBM was 3- to 4-fold higher than the proportion of proliferating CD56(-) T cells from these lesions. In addition, direct ex vivo analysis of cytokine expression by TIL from GBM demonstrated significant numbers of IL-4/IL-13 positive cells, cytokines that are integral in the cell-mediated repression of tumor immunity in experimental models. We propose that GBM has a unique capacity to recruit and activate CD4(+)CD56(+) T cells, a population that has not been previously described within human tumors.  相似文献   

6.
HIV infection is characterized by a gradual deterioration of immune function, mainly in the CD4 compartment. To better understand the dynamics of HIV-specific T cells, we analyzed the kinetics and polyfunctional profiles of Gag-specific CD4(+) and CD8(+) T cell responses in 12 subtype C-infected individuals with different disease-progression profiles, ranging from acute to chronic HIV infection. The frequencies of Gag-responsive CD4(+) and CD8(+) T cells showed distinct temporal kinetics. The peak frequency of Gag-responsive IFN-γ(+)CD4(+) T cells was observed at a median of 28 d (interquartile range: 21-81 d) post-Fiebig I/II staging, whereas Gag-specific IFN-γ(+)CD8(+) T cell responses peaked at a median of 253 d (interquartile range: 136-401 d) and showed a significant biphasic expansion. The proportion of TNF-α-expressing cells within the IFN-γ(+)CD4(+) T cell population increased (p = 0.001) over time, whereas TNF-α-expressing cells within IFN-γ(+)CD8(+) T cells declined (p = 0.005). Both Gag-responsive CD4(+) and CD8(+) T cells showed decreased Ki67 expression within the first 120 d post-Fiebig I/II staging. Prior to the disappearance of Gag-responsive Ki67(+)CD4(+) T cells, these cells positively correlated (p = 0.00038) with viremia, indicating that early Gag-responsive CD4 events are shaped by viral burden. No such associations were observed in the Gag-specific CD8(+) T cell compartment. Overall, these observations indicated that circulating Gag-responsive CD4(+) and CD8(+) T cell frequencies and functions are not synchronous, and properties change rapidly at different tempos during early HIV infection.  相似文献   

7.
8.
Although CD8(+) T cells play a central role as immune effectors, CD4(+) T cells act to control the activation and persistence of the CD8(+) T cell response in autoimmune disease, antiviral immunity, and experimental systems with immunogenic model tumor Ag. However, little information is available on the effects of CD4(+) T cells on the function of endogenous CD8(+) T lymphocytes recognizing authentic tumor rejection Ag with limited immunogenicity. We report here that the prophylactic or postchallenge administration of T helper Th1-type and Th2-type CD4(+) clones specific for an unmutated rejection Ag (murine P815AB, resembling tumor-specific shared Ag in humans) leads to the induction of P815AB-specific reactivity in vivo and concomitant tumor destruction, with quantitative rather than qualitative differences characterizing the antitumor activity of Th1 vs Th2 cells. Because the transferred CD4(+) cells lacked direct antitumor activity in vitro and required the de novo generation of P815AB-specific CD8(+) T cells in vivo, these findings suggest that CD4(+) lymphocytes can enhance the ability of host APC to initiate an endogenous CD8(+) T cell response to authentic, poorly immunogenic tumor rejection Ag.  相似文献   

9.
Mice (I-A(b-/-)) that lack CD4(+) T cells remain healthy for at least three months after respiratory exposure to the murine gamma-herpesvirus 68 (gammaHV68), then succumb with symptoms of chronic wasting disease. Postexposure challenge of gammaHV68-infected I-A(b+/+) and I-A(b-/-) mice with a recombinant vaccinia virus (Vacc-p56) expressing an antigenic gammaHV68 peptide caused a massive increase in the numbers of D(b)p56-specific CD8(+) T cells. Previous experiments showed that, despite the large numbers of potential CTL effectors, there was little effect on the long-term survival of the CD4-deficient group and no diminution in the level of persistent virus shedding and latency. Comparison of the expanded CD8(+)D(b)p56(+) sets in the I-A(b+/+) and I-A(b-/-) mice indicated that these two T cell populations were not identical. More CD69(high)CD8(+) D(b)p56(+) T cells were found in the CD4-deficient mice, an effect that might be thought to reflect higher Ag load. By contrast, the mean fluorescence intensity of staining for the CD44 glycoprotein was diminished on CD8(+)D(b)p56(+) T cells from the I-A(b-/-) group, the level of CTL activity was lower on a per cell basis, and the relative prevalence of IFN-gamma(+)TNF-alpha(+) T cells detected after in vitro stimulation with the p56 peptide was decreased. Given that this experimental system provides an accessible model for evaluating postexposure vaccination protocols that might be used in diseases like HIV/AIDS, the further need is to clarify the underlying molecular mechanisms and the relative significance of lack of CD4(+) T help vs higher Ag load for these expanded CD8(+) effector populations.  相似文献   

10.
Deleterious effect of HIV-1 plasma viremia on B cell costimulatory function   总被引:9,自引:0,他引:9  
HIV infection leads to numerous immunologic defects, including impaired B cell function. An effective humoral response requires bidirectional interactions between B cells and CD4(+) T cells, critical of which are interactions between CD80/CD86 expressed on activated B cells and CD28 expressed on responder CD4(+) T cells. In the present study, we examined the effect of active HIV replication on B cell costimulatory function. Induction of CD80/CD86 on B cells following B cell receptor and CD40 triggering and responsiveness of CD4(+) T cells to activated B cells were investigated in a system where B cells of HIV-infected patients were compared concurrently to B cells of HIV-negative donors. In contrast to HIV-aviremic patients, B cells of HIV-viremic patients were ineffective at stimulating CD4(+) T cells, as measured by the induction of activation markers and proliferation. The importance of interactions of CD80/CD86 and CD28 in activating CD4(+) T cells was clear; the ablation of a normal response following the addition of neutralizing anti-CD86/CD80 Abs mirrored the response of CD4(+) T cells to B cells of HIV-viremic patients, while the addition of exogenous CD28 ligands partially restored the poor CD4(+) T cell response to the B cells of HIV-viremic patients. Ineffective B cell costimulatory function in HIV-viremic patients was associated with low induction of CD80/CD86 expression on B cells. Our findings further delineate the scope of defects associated with cognate B cell-CD4(+) T cell interactions in HIV infection and suggest that therapeutic interventions designed to enhance CD28-dependent costimulatory pathways may help restore immune functions.  相似文献   

11.
Lipid rafts are critical to the assembly of the T-cell receptor (TCR) signaling machinery. It is not known whether lipid raft properties differ in CD4+ and CD8+ T cells and whether there are age-related differences that may account in part for immune senescence. Data presented here showed that time-dependent interleukin-2 (IL-2) production was different between CD4+ and CD8+ T cells. The defect in IL-2 production by CD4+ T cells was not due to lower levels of expression of the TCR or CD28. There was a direct correlation between the activation of p56(Lck) and LAT and their association/recruitment with the lipid raft fractions of CD4+ and CD8+ T cells. p56Lck, LAT and Akt/PKB were weakly phosphorylated in lipid rafts of stimulated CD4+ T cells of elderly as compared to young donors. Lipid rafts undergo changes in their lipid composition (ganglioside M1, cholesterol) in CD4+ and CD8+ T cells of elderly individuals. This study emphasizes the differential role of lipid rafts in CD4+ and CD8+ T-cell activation in aging and suggests that the differential localization of CD28 may explain disparities in response to stimulation in human aging.  相似文献   

12.
The expression of the collagen receptor alpha(1)beta(1) integrin (VLA-1) on CD4(+) T cells is largely restricted to CCR7(-)CD45RO(+) cells that localize to inflamed tissues. Moreover, neutralizing alpha(1) integrin, in vivo, has been shown to compromise cell-mediated immunity. Our current study shows that the expression of VLA-1 on human CD4(+) T cells is restricted to conventional effectors. In contrast, Foxp3(+) T regulatory cells (Tregs) do not express this receptor. Moreover, Foxp3 or VLA-1 expression remained a mutually exclusive event in CD4(+) T cells even upon polyclonal anti-CD3-induced activation. Because TNFalpha blockade ameliorates certain T cell-dependent autoimmune disorders in humans, we investigated, in vitro, whether neutralizing TNFalpha affected the balance between the proinflammatory VLA-1(+) effectors and the counteracting Tregs. We found that anti-CD3 stimulation of freshly isolated PBL from healthy individuals, coupled with continuous TNFalpha blockade, inhibited the typical activation-dependent generation of CD4(+)VLA-1(+) Th1 cells. In contrast, it augmented the outgrowth of VLA-1(neg/dim)CD25(high) and Foxp3(+)CD4(+) T cells. Indeed, repeated anti-CD3 stimulation coupled with TNFalpha blockade generated CD4(+) T cell lines enriched for VLA-1(-)Foxp3(+) Tregs. Importantly, these CD4(+) T cells displayed potent suppressive functions toward autologous CD4(+) PBL, including the suppression of the activation-dependent induction of VLA-1(+) effectors. Thus, we propose a novel mechanism by which anti-TNFalpha therapy may restore self-tolerance, by shifting the balance between VLA-1(+) effectors and Foxp3(+) Tregs, during immune activation, in favor of the latter suppressor cell population.  相似文献   

13.
We developed a transgenic (Tg) mouse that expresses TGF-beta under control of the IL-2 promoter to investigate Th3 cell differentiation both in vitro and in vivo. We previously found that repetitive in vitro Ag stimulation results in constant expression of Foxp3 in TGF-beta-Tg Th3 cells that acquire regulatory function independent of surface expression of CD25. To examine the differentiation and function of Th3 cells in vivo and to compare them with thymic-derived CD4(+)CD25(+) regulatory T cells (Treg), we introduced the TGF-beta transgene into T cells of IL-2-deficient (IL-2(-/-)) mice. We found that the induction, differentiation, and function of TGF-beta-derived Foxp3(+) Th3 cells were independent of IL-2, which differs from thymic Tregs. In an environment that lacks functional CD25(+) thymic-derived Tregs, expression of the TGF-beta transgene in IL-2(-/-) mice led to the induction of distinct CD25(-) regulatory cells in the periphery. These cells expressed Foxp3 and efficiently controlled hyperproliferation of T cells and rescued the IL-2(-/-) mouse from lethal autoimmunity. Unlike IL-2(-/-) animals, TGF-beta/IL-2(-/-) mice had normal numbers of T cells, B cells, macrophages, and dendritic cells and did not have splenomegaly, lymphadenopathy, or inflammation in multiple organs. Accumulation of Foxp3(+) cells over time, however, was dependent on IL-2. Our results suggest that TGF-beta-derived Foxp3(+)CD25(+/-) Th3 regulatory cells represent a different cell lineage from thymic-derived CD25(+) Tregs in the periphery but may play an important role in maintaining thymic Tregs in the peripheral immune compartment by secretion of TGF-beta.  相似文献   

14.
A large proportion of the CD8(+) T cell pool in persons chronically infected with HIV consists of cells that show features of replicative senescence, an end stage characterized by irreversible cell cycle arrest, multiple genetic and functional changes, and shortened telomeres. The objective of our research was to determine whether constitutive expression of the gene for the human telomerase (hTERT) can prevent senescence-induced impairments in human virus-specific CD8(+) T cells, particularly in the context of HIV-1 disease. Our results indicate that hTERT-expressing HIV-specific CD8(+) lymphocytes show both an enhanced and sustained capacity to inhibit HIV-1 replication in in vitro coculture experiments, as well as prolonged ability to produce IFN-gamma and TNF-alpha in response to stimulation with HIV-1-derived peptides, as compared with vector-transduced controls. Loss of CD28 expression, the signature change of replicative senescence in cell culture, was retarded in those CD8(+) T cell cultures that had high levels of CD28 at the time of hTERT transduction. These findings suggest that telomere shortening may be the primary driving force behind several aspects of CD8(+) T cell dysfunction associated with replicative senescence. We also demonstrate reduced accumulation of the p16(INK4a) and p21(WAF1) cell cycle inhibitors in hTERT-transduced lymphocytes, providing a possible mechanism by which stable hTERT expression is able to circumvent the senescence barrier in CD8(+) T cells. Given the key role of CD8(+) T cell function in controlling a variety of acute and latent viral infections, approaches to retard the functional decrements associated with replicative senescence may lead to novel types of immunotherapy.  相似文献   

15.
We have identified a large population of CD3(-)7(+) cells in human fetal gut. Three- and four-color flow cytometry revealed a distinct surface Ag profile on this population; the majority were negative for CD4 and CD8, whereas most of the remainder expressed the CD8alphaalpha homodimer. In contrast about half of CD3(+) cells expressed CD4 and half expressed CD8alpha. A large proportion of CD3(-)7(+) cells expressed CD56, CD94, and CD161, and whereas CD3(+) T cells also expressed CD161, they only rarely expressed CD56 or CD94. Further studies were conducted to determine whether the CD3(-)7(+) cells have the potential to differentiate into CD3(+) cells. About half of CD3(-)7(+) cells contain intracellular CD3epsilon. Rearranged TCR gamma-chains were detected in highly purified CD3(-)7(+) cells as an early molecular sign of T cell commitment, and the pattern of rearrangement with V regions spliced to the most 5' Jgamma segment is reminiscent of early thymocyte differentiation. In reaggregate thymic organ cultures, CD3(-)7(+) cells also gave rise to CD3(+) T cells. Thus, we demonstrate that the CD3(-)7(+) cells present in the human fetal gut display a distinct phenotype and are able to develop into CD3(+) T cells.  相似文献   

16.
Aging in humans is associated with increased infections and the reduced proliferative capacity of T cells, part of the more global phenomenon termed immune senescence. The etiology of immune senescence is unknown but the accumulation of virus-specific memory T cells may be a contributory factor. We have examined CD8 T cell responses to two persistent herpesvirus infections, CMV and EBV, and to a recurrent virus infection, influenza, in different age cohorts of healthy donors using HLA-peptide tetramers and intracellular cytokine detection. Of these, CMV appears to be the most immunogenic, with the CD8 T cell response representing over 10% of the CD8 pool in many elderly donors. Interestingly, the effect of age upon EBV-specific responses depends upon donor CMV sero-status. In CMV seropositive donors, the magnitude of the EBV-specific immune response is stable with age, but in CMV seronegative donors, the response to EBV increases significantly with age. By contrast, the influenza-specific CD8 T cell immune response decreases with age, independent of CMV status. The functional activity of the herpesvirus-specific immune response decreases in elderly donors, although the characteristic phenotypes of CMV- and EBV-specific memory populations are retained. This demonstrates that aging is associated with a marked accumulation of CMV-specific CD8 T cells together with a decrease in immediate effector function. Moreover, infection with CMV can reduce prevailing levels of immunity to EBV, another persistent virus. These results suggest that carriage of CMV may be detrimental to the immunocompetent host by suppressing heterologous virus-specific immunity during aging.  相似文献   

17.
HIV-1 Ag-specific CD4(+) T cell proliferative responses in human subjects with advanced, untreated HIV-1 disease are often weak or undetectable. Conversely, HIV-1-specific CD4(+) T cell proliferation is occasionally detected following suppression of HIV-1 replication with highly active antiretroviral therapy (HAART). These observations suggest that unchecked HIV-1 replication may lead to depletion or dysfunction of HIV-1-specific CD4(+) T cells, and that these defects may be partially corrected by viral suppression and subsequent immune reconstitution. However, the impact of this immune reconstitution on the repertoire of HIV-1-specific CD4(+) T cells has not been thoroughly evaluated. To examine the HIV-1-specific CD4(+) T cell repertoire in this clinical setting, we established HIV-1 p24-specific CD4(+) T cell clones from a successfully HAART-treated subject whose pretreatment peripheral CD4 count was 0 cells/ micro l. Eleven different p24-specific CD4(+) T cell clonotypes were distinguished among 13 clones obtained. Most clones produced both IFN-gamma and IL-4 upon Ag stimulation. Clones targeted eight distinct epitopes that varied in their conservancy among HIV-1 strains, and responses were restricted by one of three MHC II molecules. Clones showed a range of functional avidities for both protein and peptide Ags. Additional studies confirmed that multiple HIV-1 p24-derived epitopes were targeted by IFN-gamma-producing CD4(+) cells from subjects first treated with HAART during advanced HIV-1 disease (median, 4.5 peptides/subject; range, 3-6). These results suggest that in HAART-treated subjects whose peripheral CD4(+) T cell pools were once severely depleted, the HIV-1-specific CD4(+) T cell repertoire may include a diverse array of clonotypes targeting multiple HIV-1 epitopes.  相似文献   

18.
p56lck, a lymphocyte-specific tyrosine protein kinase, binds to the cytoplasmic tails of the T-cell surface molecules CD4 and CD8. Cross-linking of CD4 expressed on the surface of murine thymocytes, splenocytes, and CD4+ T-cell lines induced tyrosine phosphorylation of p56lck dramatically. Cross-linking of CD8 stimulated tyrosine phosphorylation of p56lck strongly in murine L3 and GA4 cells, slightly in splenocytes, but not detectably in thymocytes. Differing effects of cross-linking on in vitro tyrosine kinase activity of p56lck were observed. An increase in the in vitro kinase activity of p56lck, when assayed with [Val5]-angiotensin II as an exogenous substrate, was found to accompany cross-linking of CD4 in three cell lines. No stimulation of the in vitro kinase activity, however, was observed after cross-linking of CD8 in L3 cells. The phosphorylation of p56lck at Tyr-394, the autophosphorylation site, was stimulated by cross-linking in all cell lines examined. Tyr-394 was the predominant site of increased tyrosine phosphorylation in two leukemic cell lines. In the other two cell lines, the phosphorylation of both Tyr-394 and an inhibitory site, Tyr-505, was found to increase. In contrast to cross-linking with antibodies, no striking increase in the tyrosine phosphorylation of p56lck was stimulated by antigenic stimulation. Therefore, the effect of antibody-induced aggregation of CD4 and CD8 on the tyrosine phosphorylation of p56lck differs, at least quantitatively, from what occurs during antigen-induced T-cell activation.  相似文献   

19.
Vasoactive intestinal peptide (VIP) has potent antiproliferative and anti-inflammatory functions in the immune system. Two structurally distinct G-protein-associated receptors, VIP receptor type 1 (VPAC1) and VIP receptor type 2 (VPAC2), mediate the biological effects of VIP. The regulation of VIP receptor gene expression and the distribution of these receptors in different compartments of the human immune systems are unknown. This study reports, for the first time, a quantitative analysis of VPAC1 and VPAC2 mRNA expression in resting and activated T cells as well as in resting monocytes. Purified human peripheral blood CD4(+) T cells and CD8(+) T cells were stimulated via the TCR/CD3 receptor complex. Using the novel fluorometric-based kinetic (real-time) RT-PCR, we determined that VPAC1 is constitutively expressed in resting T cells and monocytes; the levels of expression were significantly higher in monocytes and CD4(+) T cells than in CD8(+) T cells. VPAC1 mRNA expression is significantly higher relative to VPAC2 in resting CD4(+) T cells and CD8(+) T cells. VPAC2 is expressed at very low levels in resting T cells but is not detectable in resting monocytes. In vitro stimulation of Th cells with soluble anti-CD3 plus PMA induced a T cell activation-dependent down-regulation of VPAC1. VPAC1 is down-regulated under conditions of optimal T cell stimulation. Our results suggest that selective VIP effects on T cell function may be mediated via selective expression of VPAC1 and VPAC2 on T cells and monocytes. Furthermore, down-regulation of VPAC1 in CD4(+) T cell subpopulations is highly correlated with T cell activation.  相似文献   

20.
Within the ovarian cancer microenvironment, there are several mechanisms that suppress the actions of antitumor immune effectors. Delineating the complex immune microenvironment is an important goal toward developing effective immune-based therapies. A dominant pathway of immune suppression in ovarian cancer involves tumor-associated and dendritic cell (DC)-associated B7-H1. The interaction of B7-H1 with PD-1 on tumor-infiltrating T cells is a widely cited theory of immune suppression involving B7-H1 in ovarian cancer. Recent studies suggest that the B7-H1 ligand, programmed death receptor-1 (PD-1), is also expressed on myeloid cells, complicating interpretations of how B7-H1 regulates DC function in the tumor. In this study, we found that ovarian cancer-infiltrating DCs progressively expressed increased levels of PD-1 over time in addition to B7-H1. These dual-positive PD-1(+) B7-H1(+) DCs have a classical DC phenotype (i.e., CD11c(+)CD11b(+)CD8(-)), but are immature, suppressive, and respond poorly to danger signals. Accumulation of PD-1(+)B7-H1(+) DCs in the tumor was associated with suppression of T cell activity and decreased infiltrating T cells in advancing tumors. T cell suppressor function of these DCs appeared to be mediated by T cell-associated PD-1. In contrast, ligation of PD-1 expressed on the tumor-associated DCs suppressed NF-κB activation, release of immune regulatory cytokines, and upregulation of costimulatory molecules. PD-1 blockade in mice bearing ovarian cancer substantially reduced tumor burden and increased effector Ag-specific T cell responses. Our results reveal a novel role of tumor infiltrating PD-1(+)B7-H1(+) DCs in mediating immune suppression in ovarian cancer.  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