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.
Whether CD8 T cell memory exists outside secondary lymphoid organs is unclear. Using an adoptive transfer system that enables tracking of OVA-specific CD8 T cells, we explored the antigenic requirements for inducing CD8 T cell memory and identified intestinal mucosa memory cells. Although systemic immunization with soluble OVA induced clonal expansion, memory CD8 cells were not produced. In contrast, infection with virus-encoding OVA induced memory CD8 cells in the periphery and the lamina propria and intraepithelial compartments of the intestinal mucosa. Mucosal memory cells expressed a distinct array of adhesion molecules as compared with secondary lymphoid memory cells, suggesting that there may be separate mucosal and systemic memory pools. Mucosal CD8 memory cells rapidly produced IFN-gamma after Ag stimulation. Reactivation of memory cells by Ag feeding resulted in increased cell size and up-regulation of CD28 and CD11c. CD8 mucosal memory cells exhibited ex vivo lytic activity that was up-regulated dramatically following Ag reencounter in vivo. Interestingly, reactivation of memory cells did not require CD28-mediated costimulation. The ability of the intestinal mucosa to maintain CD8 memory cells provides a potential mechanism for effective mucosal vaccination.  相似文献   

2.
In a state of lymphopenia, naive and memory CD4 T cells compete with each other for expansion at the expense of naive T cells. This competition prevents the proliferation as well as the phenotypic and functional conversion of naive T cells to "memory-like" T cells and may consequently prevent immune pathology frequently associated with lymphopenia-induced proliferation of naive cells. However, in T cell replete mice, memory T cells do not compete with naive T cells, indicating independent homeostatic control of naive and memory CD4 T cells in conditions that do not involve profound lymphopenia. Moreover, within the memory compartment, subsequent generation of new memory T cells precludes the survival of memory-like T cells. Thus, memory T cells have a major role in the control of lymphopenia-induced proliferation of naive cells because they inhibit both the generation of memory-like T cells and their persistence within the memory compartment.  相似文献   

3.
When memory cells are transferred to syngeneic irradiated recipients and then challenged at various times after transfer, a precipitous decline in the ability of these cells to mount a secondary response is seen. Using this model we have investigated some of the influences which antigen can exert on the memory cell population. The results indicate that antigen may: 1) either stimulate the memory cells to proliferate and form new memory cells or stimulate memory cells to become antibody forming cells and 2) selectively trigger the memory cells for low or high affinity antibody production. This selective antigen triggering appeared to depend upon its concentration: high dose antigen challenge led to the production of large amounts of lower affinity antibody but stimulated less memory cell proliferation while low dose challenge showed just the opposite. Control experiments indicated that recruitment of new memory cells from a virgin precursor population was not responsible for these observations. Our results thus suggest that an asymmetrical division of memory cells is occurring in which antigen can exert selective influences in much the same way as seen with virgin precursor cells.  相似文献   

4.
CD8 memory T cells are enriched in the bone marrow, a site where these cells are thought to receive homeostatic signals. However, the primary site where CD8 memory T cells receive their cytokine-induced homeostatic signals has recently come under debate. In this study, we demonstrate that the bone marrow contains a fraction of CD8 memory phenotype T cells with elevated expression of glucocorticoid-induced TNFR-related protein (GITR). In contrast, splenic and lymph node memory phenotype T cells have GITR levels similar to those on naive T cells. The bone marrow GITR(hi) memory T cells have a phenotype indicative of cytokine activation, with higher CD122 and lower CD127 than do the GITR(basal) memory T cells. Remarkably, these bone marrow-specific GITR(hi) cells are almost completely ablated in the absence of IL-15, whereas TNFR2 and 4-1BB expression on the CD8 memory T cells are IL-15 independent. Furthermore, adoptively transferred splenic CD8 memory phenotype T cells show IL-15-dependent GITR upregulation upon entry into the bone marrow. This result implies that the selective appearance of GITR(hi) memory phenotype T cells in the bone marrow reflects the local microenvironment rather than a different subset of memory T cells. GITR(-/-) mice have a lower frequency of CD8 memory phenotype cells in the bone marrow, yet the GITR(-/-) cells hyperproliferate compared with those in wild-type mice. Taken together, these data suggest that GITR plays a role in the survival of CD8 memory phenotype T cells and that GITR upregulation represents a precise marker of cells that have responded to IL-15.  相似文献   

5.
Following acute infection in some mouse models, CD4(+) memory T cells steadily decline over time. Conversely, in humans, CD4(+) memory T cells can be maintained for many years at levels similar to CD8(+) T cells. Because we previously observed that the longevity of Th1 memory cell survival corresponded to their functional avidity, we hypothesized that secondary challenge, which enriches for high functional avidity Th1 responders, would result in more stable Th1 memory populations. We found that following a heterologous secondary challenge, Th1 memory cells were maintained at stable levels compared with primary Th1 memory cells, showing little to no decline after day 75 postinfection. The improved stability of secondary Th1 memory T cells corresponded to enhanced homeostatic turnover; enhanced trafficking of effector memory Th1 cells to tissue sites of infection, such as the liver; and acquisition or maintenance of high functional avidity following secondary challenge. Conversely, a weaker homologous rechallenge failed to induce a stable secondary Th1 memory population. Additionally, homologous secondary challenge resulted in a transient loss of functional avidity by Th1 memory cells recruited into the secondary response. Our findings suggest that the longevity of Th1 memory T cells is dependent, at least in part, on the combined effects of primary and secondary Ag-driven differentiation. Furthermore, they demonstrate that the quality of the secondary challenge can have profound effects on the longevity and function of the ensuing secondary Th1 memory population.  相似文献   

6.
Defining the cellular composition of the memory T cell pool has been complicated by an inability to distinguish effector and memory T cells. We present here an activation profile assay, using anti-CD3 and antigenic stimuli, that clearly distinguishes effector and memory CD4 T cells and defines subsets of long-lived memory CD4 T cells based on CD62 ligand (CD62L) expression. The CD62L(low) memory subset functionally resembles effector cells, exhibiting hyper-responsiveness to antigenic and anti-CD3 mediated stimuli, high proliferative capacity, and rapid activation kinetics. The CD62L(high) memory subset functionally resembles resting memory cells, exhibiting hyporesponsiveness to anti-CD3 stimuli, lower proliferative capacity, and slower activation kinetics. Our results indicate that the memory CD4 T cell pool is heterogeneous, consisting of persisting effectors and resting memory T cells.  相似文献   

7.
Protection against malaria often decays in the absence of infection, suggesting that protective immunological memory depends on stimulation. Here we have used CD4(+) T cells from a transgenic mouse carrying a T cell receptor specific for a malaria protein, Merozoite Surface Protein-1, to investigate memory in a Plasmodium chabaudi infection. CD4(+) memory T cells (CD44(hi)IL-7Rα(+)) developed during the chronic infection, and were readily distinguishable from effector (CD62L(lo)IL-7Rα(-)) cells in acute infection. On the basis of cell surface phenotype, we classified memory CD4(+) T cells into three subsets: central memory, and early and late effector memory cells, and found that early effector memory cells (CD62L(lo)CD27(+)) dominated the chronic infection. We demonstrate a linear pathway of differentiation from central memory to early and then late effector memory cells. In adoptive transfer, CD44(hi) memory cells from chronically infected mice were more effective at delaying and reducing parasitemia and pathology than memory cells from drug-treated mice without chronic infection, and contained a greater proportion of effector cells producing IFN-γ and TNFα, which may have contributed to the enhanced protection. These findings may explain the observation that in humans with chronic malaria, activated effector memory cells are best maintained in conditions of repeated exposure.  相似文献   

8.
Humoral immune responses elicited after secondary exposure to immunizing Ag are characterized by robust and elevated reactivity of memory B cells that exceed those of naive B cells during the primary response. The mechanism underlying this difference in responsiveness of naive vs memory B cells remains unclear. We have quantitated the response of naive and memory human B cells after in vitro stimulation with T cell-derived stimuli. In response to stimulation with CD40 ligand alone or with IL-10, both IgM-expressing and Ig isotype-switched memory B cells entered their first division 20-30 h earlier than did naive B cells. In contrast, the time spent traversing subsequent divisions was similar. Consistent with previous studies, only memory cells differentiated to CD38(+) blasts in a manner that increased with consecutive division number. These differentiated CD38(+) B cells divided faster than did CD38(-) memory B cell blasts. Proliferation of CD40 ligand-stimulated naive B cells as well as both CD38(+) and CD38(-) cells present in cultures of memory B cells was increased by IL-10. In contrast, IL-2 enhanced proliferation of CD38(-) and CD38(+) memory B cell blasts, but not naive cells. Thus, memory B cells possess an intrinsic advantage over naive B cells in both the time to initiate a response and in the division-based rate of effector cell development. These differences help explain the accelerated Ab response exhibited by memory B cells after secondary challenge by an invading pathogen, a hallmark of immunological memory.  相似文献   

9.
Memory B cells can persist for a lifetime and be reactivated to yield high affinity, isotype switched plasma cells. The generation of memory B cells by Ag immunization requires adjuvants that generally contain TLR agonists. However, requirements for memory B cell activation and the role of TLRs in this activation are not well understood. In this study, we analyzed the response of memory B cells from immunized mice to TLR9 and 4 agonists CpG oligodeoxynucleotides (ODN) and LPS. Mouse memory B cells express both TLR9 and 4, and respond to both CpG ODN and LPS in vitro by differentiating into high affinity IgG secreting plasma cells. In contrast, neither CpG ODN nor LPS alone is sufficient to activate memory B cells in vivo. Ag is required for the clonal expansion of Ag-specific memory B cells, the differentiation of memory B cells to high affinity IgG secreting plasma cells, and the recall of high affinity Ab responses. The Ag-specific B cells that have not yet undergone isotype switching showed a relatively higher expression of TLR4 than memory B cells, which was reflected in a heightened response to LPS, but in both cases yielded mostly low affinity IgM secreting plasma cells. Thus, although memory B cells are sensitive to TLR agonists in vitro, TLR agonists alone appear to have little affect on B cell memory in vivo.  相似文献   

10.
Human memory B cells comprise isotype-switched and nonswitched cells with both subsets displaying somatic hypermutation. In addition to somatic hypermutation, CD27 expression has also been considered a universal memory B cell marker. We describe a new population of memory B cells containing isotype-switched (IgG and IgA) and IgM-only cells and lacking expression of CD27 and IgD. These cells are present in peripheral blood and tonsils of healthy subjects and display a degree of hypermutation comparable to CD27+ nonswitched memory cells. As conventional memory cells, they proliferate in response to CpG DNA and fail to extrude rhodamine. In contrast to other recently described CD27-negative (CD27neg) memory B cells, they lack expression of FcRH4 and recirculate in the peripheral blood. Although CD27neg memory cells are relatively scarce in healthy subjects, they are substantially increased in systemic lupus erythematosus (SLE) patients in whom they frequently represent a large fraction of all memory B cells. Yet, their frequency is normal in patients with rheumatoid arthritis or chronic hepatitis C. In SLE, an increased frequency of CD27neg memory cells is significantly associated with higher disease activity index, a history of nephritis, and disease-specific autoantibodies (anti-dsDNA, anti-Smith (Sm), anti-ribonucleoprotein (RNP), and 9G4). These findings enhance our understanding of the B cell diversification pathways and provide mechanistic insight into the immunopathogenesis of SLE.  相似文献   

11.
Molecular and functional profiling of memory CD8 T cell differentiation   总被引:40,自引:0,他引:40  
Kaech SM  Hemby S  Kersh E  Ahmed R 《Cell》2002,111(6):837-851
How and when memory T cells form during an immune response are long-standing questions. To better understand memory CD8 T cell development, a time course of gene expression and functional changes in antigen-specific T cells during viral infection was evaluated. The expression of many genes continued to change after viral clearance in accordance with changes in CD8 T cell functional properties. Even though memory cell precursors were present at the peak of the immune response, these cells did not display hallmark functional traits of memory T cells. However, these cells gradually acquired the memory cell qualities of self-renewal and rapid recall to antigen suggesting the model that antigen-specific CD8 T cells progressively differentiate into memory cells following viral infection.  相似文献   

12.
Apoptosis is a critical regulator of homeostasis in the immune system. In this study we demonstrate that memory CD8(+) T cells are more resistant to apoptosis than naive cells. After whole body irradiation of mice, both naive and memory CD8(+) T cells decreased in number, but the reduction in the number of naive cells was 8-fold greater than that in memory CD8(+) T cells. In addition to examining radiation-induced apoptosis, we analyzed the expansion and contraction of naive and memory CD8(+) T cells in vivo following exposure to Ag. We found that memory CD8(+) T cells not only responded more quickly than naive cells after viral infection, but that secondary effector cells generated from memory cells underwent much less contraction compared with primary effectors generated from naive cells (3- to 5-fold vs 10- to 20-fold decrease). Increased numbers of secondary memory cells were observed in both lymphoid and non-lymphoid tissues. When naive and memory cells were transferred into the same animal, secondary effectors underwent less contraction than primary effector cells. These experiments analyzing apoptosis of primary and secondary effectors in the same animal show unequivocally that decreased downsizing of the secondary response reflects an intrinsic property of the memory T cells and is not simply due to environmental effects. These findings have implications for designing prime/boost vaccine strategies and also for optimizing immunotherapeutic regimens for treatment of chronic infections.  相似文献   

13.
Proliferative renewal of memory CD8 T cells is essential for maintaining long-term immunity. In this study, we examined the contributions that various tissue microenvironments make toward the homeostatic proliferation of Ag-specific memory CD8 T cells. We found that dividing memory T cells were present in both lymphoid and nonlymphoid tissues. However, the bone marrow was the preferred site for proliferation and contained a major pool of the most actively dividing memory CD8 T cells. Adoptive transfer studies indicated that memory cells migrated through the bone marrow and divided there preferentially. These results show that the bone marrow is not only the source of stem cells for generating naive T cells but also provides the necessary signals for the self-renewal of memory T cells.  相似文献   

14.
CD4+ T lymphocytes of individuals infected with human immunodeficiency virus type 1 (HIV-1) exhibit a qualitative defect in their ability to mount memory responses to previously encountered antigens although their responses to mitogens remain normal. T cells responsible for memory responses can be distinguished from naive T cells based on differential expression of isoforms of the tyrosine phosphatase CD45. It has been suggested that memory CD4+ T cells from infected individuals have a greater virus burden than naive CD4+ T cells and that this accounts for the loss of recall responses in infected individuals. However, it has been unclear whether naive and memory T cells are equally susceptible to infection and to the cytopathic effects of the virus. We therefore infected highly purified resting naive and memory CD4+ T cells from HIV-1-seronegative individuals with HIV-1(LAI). Infected cells were then stimulated with phytohemagglutinin to render them permissive for viral replication. Cell viability and growth rate were monitored for 8 to 10 days as indicators of cytopathic effects induced by HIV-1(LAI). Our results indicated that naive and memory CD4+ T cells display marked differences in susceptibility to the cytopathic effects induced by HIV-1(LAI), infection. The cytopathic effects induced by HIV-1(LAI) were much more severe in memory CD4+ T cells than in naive CD4+ T cells. Differential cytopathic effects in naive and memory T cells were not due to differences in virus entry into and replication in these cell populations. Rather, memory cells were more susceptible to cytopathic effects. Pronounced cytopathic effects in memory cells were clearly detectable at 7 day postinfection. Cell death occurred at the single-cell level and was not accompanied by syncytium formation. The growth rate of infected memory CD4+ T cells was also severely compromised compared to that of naive CD4+ T cells, whereas the growth rates of both uninfected naive and memory CD4+ T cells were approximately the same. At least a portion of the dying cells exhibited biochemical changes characteristic of apoptosis. These results suggest that the selective functional defects present in the memory CD4+ T-cell subset of HIV-1-infected individuals may in part be the result of the greater susceptibility of memory T cells to cytopathic effects induced by HIV-1.  相似文献   

15.
Virtual memory T cells are foreign antigen‐inexperienced T cells that have acquired memory‐like phenotype and constitute 10–20% of all peripheral CD8+ T cells in mice. Their origin, biological roles, and relationship to naïve and foreign antigen‐experienced memory T cells are incompletely understood. By analyzing T‐cell receptor repertoires and using retrogenic monoclonal T‐cell populations, we demonstrate that the virtual memory T‐cell formation is a so far unappreciated cell fate decision checkpoint. We describe two molecular mechanisms driving the formation of virtual memory T cells. First, virtual memory T cells originate exclusively from strongly self‐reactive T cells. Second, the stoichiometry of the CD8 interaction with Lck regulates the size of the virtual memory T‐cell compartment via modulating the self‐reactivity of individual T cells. Although virtual memory T cells descend from the highly self‐reactive clones and acquire a partial memory program, they are not more potent in inducing experimental autoimmune diabetes than naïve T cells. These data underline the importance of the variable level of self‐reactivity in polyclonal T cells for the generation of functional T‐cell diversity.  相似文献   

16.
Ag-experienced or memory T cells have increased reactivity to recall Ag, and can be distinguished from naive T cells by altered expression of surface markers such as CD44. Memory T cells have a high turnover rate, and CD8(+) memory T cells proliferate upon viral infection, in the presence of IFN-alphabeta and/or IL-15. In this study, we extend these findings by showing that activated NKT cells and superantigen-activated T cells induce extensive bystander proliferation of both CD8(+) and CD4(+) memory T cells. Moreover, proliferation of memory T cells can be induced by an IFN-alphabeta-independent, but IFN-gamma- or IL-12-dependent pathway. In these conditions of bystander activation, proliferating memory (CD44(high)) T cells do not derive from activation of naive (CD44(low)) T cells, but rather from bona fide memory CD44(high) T cells. Together, these data demonstrate that distinct pathways can induce bystander proliferation of memory T cells.  相似文献   

17.
Naive T cell circulation is restricted to secondary lymphoid organs. Effector and memory T cells, in contrast, acquire the ability to migrate to nonlymphoid tissues. In this study we examined whether nonlymphoid tissues contribute to the differentiation of effector T cells to memory cells and the long-term maintenance of memory T cells. We found that CD4, but not CD8, effector T cell differentiation to memory cells is impaired in adoptive hosts that lack secondary lymphoid organs. In contrast, established CD4 and CD8 memory T cells underwent basal homeostatic proliferation in the liver, lungs, and bone marrow, were maintained long-term, and functioned in the absence of secondary lymphoid organs. CD8 memory T cells found in nonlymphoid tissues expressed both central and effector memory phenotypes, whereas CD4 memory T cells displayed predominantly an effector memory phenotype. These findings indicate that secondary lymphoid organs are not necessary for the maintenance and function of memory T cell populations, whereas the optimal differentiation of CD4 effectors to memory T cells is dependent on these organs. The ability of memory T cells to persist and respond to foreign Ag independently of secondary lymphoid tissues supports the existence of nonlymphoid memory T cell pools that provide essential immune surveillance in the periphery.  相似文献   

18.
Following activation within secondary lymphoid tissue, CD8 T cells must migrate to targets, such as infected self tissue, allografts, and tumors, to mediate contact-dependent effector functions. To test whether the pattern of migration of activated CD8 T cells was dependent on the site of Ag encounter, we examined the distribution of mouse Ag-specific CD8 T cells following local challenges. Our findings indicated that activated CD8 T cells migrated pervasively to all nonlymphoid organs irrespective of the site of initial Ag engagement. Using an adoptive transfer system, migration of nonlymphoid memory cells was also examined. Although some limited preference for the tissue of origin was noted, transferred CD8 memory T cells from various nonlymphoid tissues migrated promiscuously, except to the intestinal mucosa, supporting the concept that distinct memory pools may exist. However, regardless of the tissue of origin, reactivation of transferred memory cells resulted in widespread dissemination of new effector cells. These data indicated that recently activated primary or memory CD8 T cells were transiently endowed with the ability to traffic to all nonlymphoid organs, while memory cell trafficking was more restricted. These observations will help refine our understanding of effector and memory CD8 T cell migration patterns.  相似文献   

19.
Mechanisms of activation of alloimmune memory cells by immunologically nonspecific signals were investigated utilizing the mitogenic oxidizing agents, neuraminidase and galactose oxidase (NAGO) and sodium periodate (IO4?). Direct activation of primary long-term human mixed-lymphocyte culture (MLC) cells (memory cells) with either NAGO or IO4? resulted in increased specific secondary cytolytic activity. Kinetics of the proliferative and cytotoxic responses resulting from such treatment of memory cells paralleled those resulting from treatment of memory cells with irradiated cells that were the stimulators in the primary MLC. Indirect activation of memory cells by NAGO or IO4?-treated and irradiated syngeneic cells also resulted in increased specific secondary cytolytic activity. In contrast, peripheral blood mononuclear cells (PBM) activated by the mitogenic oxidizing agents did not develop cytolytic activity toward syngeneic or multiple allogeneic target cells, despite extensive proliferative responses. Cytotoxicity generated by either direct or indirect activation of memory cells by IO4? was prevented by treatment of the oxidized cells with the reducing agent, sodium borohydride. Incubation of memory cells in supernatants from 24-hr cultures of PBM activated with either NAGO or IO4? resulted in proliferation and in an increase in cytolytic activity in memory cells, but not in PBM. These findings indicate that alloimmune memory cells can be activated by immunologically nonspecific lymphocyte-derived signals that do not depend on alloantigen or lectin.  相似文献   

20.
The distribution of functionally active hapten-specific B memory cells was investigated. Using antigen-pretreated lethally irradiated recipients, a marked accumulation of adoptively transferred B memory cells was demonstrated in lymph nodes containing specific antigen, but not in lymph nodes containing non-cross-reacting hapten conjugates. This difference in responsiveness between lymph nodes containing specific versus those containing nonspecific antigen developed over a period 3–5 days after memory cell transfer. The localization of antigen specific cells was T-cell independent; both carrier-primed T helper cells and specific antigenic challenge, however, were required to trigger the localized B memory cells into antibody production. Specific B memory cell accumulation did not result from an expansion of the antigen-specific cell population due to local proliferation induced by antigen depots in the lymph nodes to challenge. Rather, the results indicated that recirculating B memory cells had progressively accumulated through retention by antigen in the lymph node. These findings suggest that, in the absence of T-cell help and specific antigenic challenge, B memory cells accumulate in lymphoid tissue (follicles) without responding and provide persistent local memory for the humoral immune response.  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