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 19 毫秒
1.
Development and regulation of chlamydia-responsive murine B lymphocytes   总被引:2,自引:0,他引:2  
We have examined characteristics of chlamydia-stimulated mouse B cells as well as cells that regulate polyclonal responses in vitro. B lymphocyte proliferation stimulated by chlamydia arises at a similar time as Escherichia coli lipopolysaccharide (LPS)-induced proliferative responses during ontogeny. In contrast, development of immunoglobulin (Ig)-secreting cells after chlamydia stimulation is delayed by several weeks relative to ontogeny of LPS-inducible plaque-forming cells (PFC). The lack of Ig secretion by immature B cells is not due to a deficiency of Lyb5+ B lymphocytes, since X-linked immunodeficient (xid) NBF1 mice that lack this B lymphocyte population respond well to chlamydia stimulation. Adherent cells are important for chlamydia-stimulated B lymphocyte differentiation, but are not as necessary for their proliferation. Neither adult adherent cells nor T cells can correct the inability of immature spleen cells to develop into Ig-secreting cells; spleen cells from 2-wk-old mice (i.e., immature B cells) will not suppress adult B lymphocyte responses to chlamydia. When B lymphocytes are separated according to their buoyant densities, chlamydia stimulates low density (activated) B cells to proliferate and differentiate better than high density (resting) cells. Proliferative responses to chlamydia arise earlier during ontogeny, do not require adherent cells, and can proceed to a relatively greater extent in resting B cell population (compared with activated B cells) than induction of Ig-secreting cells.  相似文献   

2.
The modulation of membrane Ia on human B lymphocytes   总被引:2,自引:0,他引:2  
Using flow cytometry techniques, changes in surface Ia (DR and DS) expression on human B lymphocytes were correlated with changes in the cell cycle following stimulation with anti-mu. The effect of interleukin (IL)-1, IL-2, B-cell growth factor (BCGF), and interferons on Ia expression on resting B cells was also examined. A population of resting B lymphocytes was cultured in vitro with 100 micrograms/ml of anti-mu and immunofluorescently stained for DR and DS at various times following stimulation. Detectable increases in DR and DS expression were found within 8 hr, and the major increases (twofold and fourfold) in DR and DS expression occurred over the next 48 hr. Using cell cycle inhibitors and propidium iodide staining, it was demonstrated that the enhanced DR and DS expression following anti-mu stimulation began during G0 to G1 transition and increased as the cells progressed through G1 phase. During S and G2/M phases, there were minimal further increases in surface Ia. Although prolonged exposure of B cells to anti-mu was required for cellular activation, cell size enlargement, and progression into S phase, a brief exposure to anti-mu, insufficient for cellular activation, markedly enhanced Ia expression. Thus anti-mu-stimulated resting human B lymphocytes rapidly increase their surface Ia expression. This increase occurs predominantly prior to entrance into S phase and can occur in the absence of significant cellular activation. Interferons have been reported to modulate surface Ia expression on a human lymphoid cell line and on monocytes and supernatants with BCGF activity to enhance surface Ia expression on murine B cells; however, neither alpha-interferon, gamma-interferon, IL-1, IL-2, nor BCGF modified surface DR expression on normal resting human B cells.  相似文献   

3.
Functional studies of both polyclonal and antigen-specific responses have suggested that murine B cells differ in the expression of an antigen recognized by a rat anti-mouse monoclonal antibody, called J11d. Using both positive and negative selection, we now demonstrate that the J11d marker is differentially displayed on B lymphocytes responding to LPS vs anti-mu, as well as on unprimed vs specific antigen-primed B cells. Thus, cytotoxic elimination of cells expressing high levels of J11d (J11d-hi) reduced LPS-driven B cell proliferation by 60 to 80% but had no effect on anti-mu stimulated B cell growth. Interestingly, equal numbers of positively selected J11d-hi B cells responded similarly to LPS and anti-mu plus B cell growth factors, a result that suggests that the response to anti-mu of the J11d-lo B cells is normally masked by the majority J11-d-hi cells. In further studies, the primary PFC response of normal murine spleen cells to fluorescein (FL)-coupled TI antigens or to LPS in vitro was reduced dramatically by cytotoxic J11d antibody treatment. In contrast, the anti-FL PFC response of spleen cells from mice primed 1 wk previously with FL-Ficoll was not affected by J11d antibody treatment, whereas the response of these FL-primed B cells to TNP (to which the mice were not primed) was greatly reduced by J11d + complement treatment. Our data indicate that antigen-experienced (activated) B cells are primarily found in the J11d-lo B cell subset and that unprimed (resting) B cells are found in the J11d-hi population, although both populations of murine B cells can respond to anti-mu. These studies also provide further evidence for B cell heterogeneity.  相似文献   

4.
The effects of a preparation containing partially purified, EL4-derived B cell growth factor(s) (BCGF) on B cell growth and proliferation have been examined by using B lymphocyte subpopulations separated on the basis of size. BCGF was found to maintain and enhance proliferation of a significant proportion of large activated B cells. In contrast, small resting B cells required the presence of BCGF and a second stimulus such as anti-IgM antibody (anti-mu) to be induced to proliferate. This disparity was not due to a lack of an effect of BCGF on small resting B cells. A factor contained within the partially purified EL4 supernatant produced time-dependent increases in cell size and RNA content in all subpopulations. These effects were independent of possible effects due to contaminating lymphokines such as interleukin 2 (IL 2), concanavalin A (Con A), and phorbol myristate acetate (PMA). Nonmitogenic doses of lipopolysaccharide (LPS) failed to show similar effects. Our data suggest that B cells at all levels of in vivo activation are responsive to stimulation by a growth factor present in EL4 supernatant, as manifested by cell growth and RNA synthesis. This activity has not previously been described for BCGF preparations. However, because the partially purified, EL4-derived supernatant used as BCGF in these studies has not been purified to homogeneity, we cannot conclude whether the factors that induce resting B cells to increase in size are the same as the growth factors that synergize with anti-mu to induce B cell proliferation or that maintain the proliferation of activated B cells.  相似文献   

5.
Previous studies have shown that lymph node (LN) T cells from mice given repeated injections of anti-mu antisera from birth (mu sm) fail to mount secondary T proliferative responses to antigen in vitro after s.c. priming in vivo. This finding raised the possibility that priming of T cells in LN depends on the presence of B cells, Ig+ B lymphocytes being absent in mu sm. In support of this idea, the present paper shows that the priming defect in LN of mu sm can be largely overcome by injecting B cell populations s.c. 1 day before s.c. priming with antigen. Restoration of LN priming was observed with s.c. injection of highly purified populations of small B cells but not with heat-killed or lightly irradiated B cells. Homing studies indicated that approximately 10% of s.c.-injected B cells reached the draining LN. In other studies, irradiated mice injected i.v. with purified T cells manifested poor priming in LN after s.c. injection of antigen. It was reasoned that the LN priming defect in this situation reflected the lack of B cells in irradiated mice, B cells being highly radiosensitive. In support of this notion, it was found that s.c. injection of B cells into irradiated recipients of T cells led to high priming of T cells in LN after s.c. injection of antigen. Although T cells exposed to antigen in B-depleted LN of mu sm and irradiated mice gave negligible T proliferative responses in vitro, low but significant levels of primed T helper function were detected in a sensitive T helper assay in vivo. In light of this finding, our working hypothesis is that the initial induction of T cells to antigen in LN is controlled by resident dendritic cells (or other non-B antigen-presenting cells), the main role of B cells being to control the clonal expansion of activated T cells.  相似文献   

6.
Spleen cells of two rat strains, Lewis and Brown Norway (BN), have been activated by lectins and by antibodies specific for immunoglobulin isotypes embedded in their cell membranes. Optimal concentrations of antibodies specific for mu, gamma, or delta-chains of rat augments in vitro incorporation of 3H-TdR 5 to 18-fold in Lewis B lymphocytes and 1.5 to 4-fold in BN B lymphocytes. In addition, F(ab')2 fragments of anti-Ig reagents induced Lewis splenic B cells but not BN B cells to incorporate 3H-TdR. Responses to LPS and dextran sulfate, B lymphocyte mitogens, measured by radioactive uptake, were five to 10 times greater in Lewis B cell populations than in BN B cell populations. Density of surface Ig isotypes and capping kinetics were similar in the two rat strains, although the percentage of T cells, T cell subsets, B cells, and Ia+ B cells differed in the spleens of these strains of rats. Both T lymphocytes and macrophages were needed in culture to effect an optimal response. IL-2 restored the response in B cell cultures depleted of T cells and macrophages, and enhanced 3H-TdR uptake in whole spleen cells of Lewis but not BN rats. The strain-dependent responsiveness of B cells to specific anti-Ig reagents or B cell mitogens appears to be associated with inherent (genetic) defects in T cells and B cells or defects in T cell to B cell cooperation in BN rats.  相似文献   

7.
Mice bearing a transplantable CE mammary carcinoma have been shown to have greatly augmented rates of neutrophil production coupled with a marked diminution of bone marrow lymphocytes. The objective of the present study was to test whether the loss of lymphocytes, and especially of B cells, from the bone marrow and spleen of tumor-bearing animals was due to a reduced rate of cell production and if so, at what level this response was regulated. A modified 3H-TdR pulse and chase analysis was used to assess the rates of production of small lymphocytes and B cells (stained for c mu and s mu) at weekly intervals after CE tumor transplantation. 3H-TdR was infused continuously for 24 hr, and radioautographs were prepared of bone marrow and spleen cells 0, 24, and 48 hr after termination of the infusion. Pre-B cells (c mu+s mu-) essentially disappeared from the femoral bone marrow by the end of 1 wk of tumor growth, followed by a great reduction in the number of c mu+s mu+ cells in the marrow and s mu + cells in the spleen. Although pre-B cells appeared in the peripheral marrow (caudal vertebrae, metatarsal bones) and spleen of tumor-bearing mice, these cells could not compensate for the continued decrease in the numbers of more mature B cells. In normal mice, during the 48-hr chase period, newly formed, 3H-TdR-labeled, small lymphocytes and s mu+ cells continued to emerge from the prelabeled precursor compartment at a steady rate, but after 1 wk of tumor growth, the number of small lymphocytes and s mu+ cells emerging from the precursor compartment fell steadily during the 48-hr chase period. During the second and third weeks of tumor growth, a steady state appears to have been reached in B cell production, which was at a level approximately 10 times below that of normal. Because pre-B cells are normally maintained by a less mature precursor population (2), the initial disappearance of c mu+s mu- cells suggests that the CE mammary carcinoma exerts its modulatory influence on primary B cell production by inhibiting or eliminating the cells that eventually feed into the pre-B compartment. The nature of the regulatory factors apparently secreted by the tumor and the more precise identity of the target cells are under investigation.  相似文献   

8.
N,N'-Diacetylputrescine (tetramethylenebisacetamide [TMBA]) and its six carbon analog, hexamethylenebisacetamide (HMBA), inhibited the proliferative response of human B lymphocytes to anti-mu and formalinized Cowan I strain Staphylococcal aureus (SAC) stimulation. In contrast, B cell growth factor-stimulated proliferation of human B cells was minimally inhibited by TMBA or HMBA. The antiproliferative effect of these diamine derivatives was specific for anti-mu (or SAC) activation of normal B cells, because the proliferation of PHA-stimulated human T cells and transformed human B cells was not affected by the presence of TMBA or HMBA. The inhibitory effect of diacetyl diamines on anti-mu (or SAC)-induced B cell activation was dose dependent and persisted after removal of the diamine derivatives from the culture media. These studies show that diacetylated derivatives of polyamines modulate human B cell activation in vitro by specific abrogation of anti-mu or SAC activation.  相似文献   

9.
Caspases are a group of cysteine-related proteases that control the process of apoptosis and may also be involved in the control of lymphocyte activation. We show here that the broad-spectrum caspase inhibitor benzyloxycarbonyl (Cbz)-Val-Ala-Asp (Ome)-fluoromethylketone (zVAD-fmk) prevents the proliferation of resting human B tonsilar lymphocytes mediated by the B cell mitogen SAC or the combination of anti-mu Ab and IL-2. zVAD-fmk inhibits IL-2-induced phosphorylation of the retinoblastoma protein, and cyclin D2 expression. However, neither the IL-2-mediated proliferation of cycling activated B cells nor that of lymphoma cells were inhibited by zVAD-fmk, suggesting that only the early steps of SAC- or IL-2-mediated B cell activation were sensitive to the inhibitory properties of zVAD-fmk. Our data also demonstrated that the inhibitory effect of zVAD-fmk was not observed when B cells were activated with IL-4 in the presence of either anti-mu Ab or anti-CD40 Ab. Thus, our results suggest that caspase activation is required for the IL-2-mediated entry of primary B lymphocytes into the cell cycle and show that caspase activation plays different roles in IL-2- and IL-4-mediated B cell proliferation.  相似文献   

10.
Resting murine splenic B lymphocytes (B cells) can be stimulated to proliferate by exposure to a variety of polyclonal activators. To investigate changes in glycoprotein synthesis that occur during the activation process, N-glycosylation activity was assessed by following the incorporation of [2-3H]mannose into dolichol-linked oligosaccharide intermediates and glycoprotein after B cells were exposed to anti-immunoglobulin M (anti-mu). Stimulation of B cells by anti-mu resulted in a dramatic induction of N-glycosylation activity. The incorporation of radiolabeled mannose into oligosaccharide-lipid increased 9-fold while the rate of labeling of glycoprotein increased 27-fold between 18 and 38 h after exposure to anti-mu. Maximal stimulation of N-glycosylation activity was observed at an anti-mu concentration of 20-50 micrograms/ml. Similar results were obtained when B cells were activated by bacterial lipopolysaccharide (LPS), another polyclonal activating agent. The major dolichol-bound oligosaccharide labeled during the induction period was determined to be Glc3Man9GlcNAc2 by HPLC analysis. Nearly full induction of oligosaccharide-lipid synthesis and protein N-glycosylation was also seen when DNA synthesis was suppressed by activating B cells with anti-mu in a serum-free medium, or by activating with anti-mu or LPS in the presence of hydroxyurea. The results suggest that the N-glycosylation pathway is induced during the G0 to G1 transition or during the G1 period, and that entry into S phase is not required. These studies describe a striking developmental increase in N-glycosylation activity and extend the information on biochemical changes occurring during the activation of B cells.  相似文献   

11.
The effects of interleukin 1 on human B cell activation and proliferation   总被引:19,自引:0,他引:19  
The precise role of B cell surface immunoglobulin (slg) in the activation of B cells is unclear at present. In particular, it is uncertain whether ligands interacting with the B cell slg suffice to induce proliferation, or simply induce a state of activation in which the B cell becomes responsive to growth factors made by accessory cells. We have examined the effects of two ligands, Staphylococcus aureus Cowan strain I (SAC) and antihuman mu chain (anti-mu), which interact with B cell slg on highly purified human peripheral blood and tonsillar B cells cultured at low cell concentrations. The effects on B cell proliferation of these ligands alone or in combination with highly purified interleukin 1 (IL 1) or a supernatant of a human T-T hybridoma containing a B cell growth factor (BCGF) were studied. SAC with its high cell wall content of protein A triggered maximal B cell proliferation which was not increased further by IL 1 or BCGF. High concentrations of soluble F(ab')2 fragments of goat anti-mu chain also induced significant B cell proliferation. Lower concentrations of anti-mu resulted in little or no B cell proliferation but activated the B cell to a state of responsiveness to both IL 1 and BCGF. IL 1 by itself had no effect on the proliferation of unstimulated B cells or on the proliferation of in vivo-activated B cells which responded to BCGF in vitro, but demonstrated clear synergy with low concentrations of anti-mu antibody. BCGF alone augmented the proliferation of unstimulated B cells, presumably by acting on B cells which had undergone some degree of activation in vivo. In addition, it showed marked synergy with anti-mu antibody, which resulted in proliferation similar in magnitude to that induced by SAC. This synergy was far greater than that seen between anti-mu antibody and IL 1, and the resulting proliferative response was only slightly increased by the presence of IL 1. We conclude that the importance of accessory cell factors for the initial rounds of B cell proliferation depends on the strength of the initial slg-mediated activation signal. When this is strong, the response is maximal and independent of accessory cells or accessory cell factors. When it is suboptimal, a moderate synergy is seen with IL 1 and a dramatic synergy with BCGF.  相似文献   

12.
13.
A radioautographic immunolabeling technique has been developed to detect pre-B cells bearing cytoplasmic mu chains among populations of bone marrow lymphoid cells identified by conventional hematologic stains. 125I-Anti-mu antibody was applied either to fixed marrow smears, labeling total mu chains both in the cytoplasm (c mu) and at the cell surface (s mu), or to cell suspensions, labeling s mu alone. In stained radioautographs the incidence of c mu+ s mu- pre-B cells was derived both indirectly by subtracting values for s mu+ cells from those for total mu+ cells of various sizes in normal mice and directly by the total mu chain labeling in mice depleted of s mu+ cells by anti-IgM treatment in vivo. Binding specificity was demonstrated by the displacement of labeling by nonradioactive anti-mu antibody. The c mu+ s mu- cells showed a bimodal size distribution. They accounted for 40% of the large lymphoid cells and 30% of the small lymphocytes in the marrow. A further 50% of the small lymphocytes were B lymphocytes (s mu+) and 8% were T lymphocytes (Thy 1.2+). Thus, the technique demonstrates the presence of c mu+ s mu- pre-B cells among both proliferating large lymphoid cells and nondividing small lymphocytes, as classically defined in marrow smears. In addition, the results reveal a broad size distribution of mu- lymphoid cells, including a subset of small lymphocytes which lack c mu, s mu, and Thy 1.2 and thus cannot be assigned to either B or T lineage by these criteria. The findings suggest that in addition to B cells the marrow may produce other types of lymphoid cells, yet to be defined.  相似文献   

14.
To test the hypothesis that resting and previously activated B lymphocytes differ in their proliferative and differentiative responses to various Th cell-derived stimuli, we have examined the interactions of purified small (resting) and large (activated) murine B cells with rabbit Ig-specific Th1 and Th2 clones in the presence of the Ag analogue, rabbit anti-mouse Ig antibody. Small numbers of Th2 cells induce strong Ag-dependent proliferation of and Ig secretion by both resting and activated B lymphocytes. In contrast, Th1 clones stimulate lower responses of activated B cells and fail to stimulate small resting B cells. An interaction with Th1 clones does make small B cells responsive to the Th2-derived cytokine, IL-4, indicating that Th1 clones are capable of delivering some but not all the stimuli necessary for the induction of humoral immunity. Finally, in order to compare the responses of small and large B cells to cognate interactions and secreted cytokines, we used an autoreactive I-Ak-specific Th2 line. This line induces proliferation of and Ig secretion by I-Ak expressing but not H-2d resting and activated B cells as a result of cognate interactions. However, when the H-2d B cells are bystanders in the presence of cytokine secretion by this Th2 line, or are directly exposed to Th2-derived cytokines, both small and large B cells are induced to proliferate but only the large B cells secrete antibody. These results indicate that the magnitude and nature of antibody responses depend on three principal factors: the cytokines produced by Th cells, the state of activation of the responding B lymphocytes, and whether the B cells are recipients of cognate help or are bystanders at the site of T cell stimulation. Our findings also confirm the view that cognate T-B interactions are most efficient for initiating B cell responses and may allow B cells to subsequently respond to a variety of T cell-derived cytokines.  相似文献   

15.
B cell hybridomas with Ia and IgM molecules on the cell membrane were treated with either purified goat anti-mouse mu antibody (anti-mu) or monoclonal rat anti-mouse IgM antibody (anti-IgM). The spontaneous uptake of [3H] thymidine by these cells was markedly inhibited by both reagents. These hybrid cells could be induced to differentiate into IgM-secreting cells in the presence of these reagents at high frequency. Furthermore, the induction of IgM secretion by B cell hybridomas treated with these antibodies was completely T cell independent, and cell division was not required for the differentiative response to anti-mu. In addition, F(ab')2 fragments of anti-mu showed more effects on proliferation and differentiation of these cells than intact anti-mu. Interestingly, TH2.54, a subline of B cell hybridomas, could generate IgG2a production as well as IgM when incubated with anti-mu. These findings suggest very strongly that the interaction of either goat anti-mu or monoclonal rat anti-IgM with surface IgM molecules on the cell membrane of the B cell hybridomas inhibits in vitro spontaneous proliferation, and results in providing signals for differentiation into Ig-secreting cells without T cell factors.  相似文献   

16.
Anti-mu, anti-gamma, and anti-delta antibodies induce proliferation of splenic B lymphocytes from young Lewis rats, measured by 3H-TdR uptake. In contrast, splenic B cells of aged Lewis rats respond poorly or not at all to these reagents. T lymphocytes or interleukin 2 (IL-2) of young or aged rats augment the uptake of 3H-TdR in cultures of "young" B cells responding to anti-Ig reagents or LPS and DxS, but have no significant effect on the responses of "old" B cells. Analysis of spleen cells of young and aged rats in a fluorescence-activated cell sorter indicates the density of mu, gamma, and delta isotypes is reduced in "old" B cells, and that B cells of aged rats are significantly larger than those of young rats. These results delineate anatomic and structural changes in B lymphocytes of aged rats.  相似文献   

17.
Human peripheral blood, spleen, and lymph node B cells were stimulated with Cowan I Staphylococcus aureus (SA) or F(ab')2 fragments of anti-mu antibody (anti-mu) and various lymphokines and were analyzed for proliferation and generation of Ig-secreting cells (ISC). SA alone but not anti-mu stimulated minimal proliferation of each population. Recombinant IL 2 (r-IL 2) effectively promoted proliferation of SA-stimulated blood and spleen B cells, but supported less vigorous responses of lymph node B cells. By contrast, r-IL 2 enhanced DNA synthesis of all anti-mu-stimulated B cells early in culture, but did not promote sustained proliferation of anti-mu-stimulated lymph node B cells and only promoted ongoing DNA synthesis of some anti-mu-activated blood (eight out of 17) and spleen (five out of 14) B cell preparations. Recombinant interferon-gamma (r-IFN-gamma) and a commercial preparation of B cell growth factor (BCGF) also augmented DNA synthesis of all three B cell populations stimulated with SA or anti-mu early in culture, but neither alone was able to sustain maximal proliferation. Markedly enhanced sustained proliferation of all three anti-mu- and SA-stimulated B cell populations was noted when cultures were supported by the combination of r-IL 2 and BCGF, or to a lesser extent by r-IL 2 and r-IFN-gamma. The generation of ISC from SA-stimulated blood or spleen but not lymph node B cells was effectively supported by r-IL 2 alone. Differentiation of lymph node B cells required the combination of r-IL 2 and BCGF. These studies emphasize the importance of both the activation stimulus and the origin of the B cells in determining the lymphokine requirements of human B cell responsiveness.  相似文献   

18.
We have examined phospholipid metabolism in murine B lymphocytes stimulated with anti-Ig bound to Sepharose. T cell-depleted splenic B lymphocytes cultured with Sepharose-coupled, affinity-purified goat anti-mouse Ig (GAMIg) increased the incorporation of 32PO4 into phosphatidic acid and phosphatidylinositol within 3 hr and increased [3H]-thymidine uptake at 48 hr. No increase in labeling was observed in phosphatidylethanolamine, phosphatidylcholine, or phosphatidylserine. Based on both negative and positive selection procedures, it was demonstrated that these responses occurred in B lymphocytes. In contrast to the thymidine uptake response of the GAMIg-stimulated B lymphocytes, the phospholipid response did not require the presence of accessory cells or exogenous cytokines. The same selective changes in phospholipid metabolism were observed in neoplastic B lymphocytes (BCL1) after treatment with Sepharose anti-mu, but not with Sepharose anti-Ia or Sepharose normal Ig. The dose-response relationships of 32PO4 incorporation into phosphatidic acid and phosphatidylinositol and [3H] thymidine uptake were nearly identical in BCL1 cells. The results of these experiments indicate that interaction of B lymphocytes with insolubilized anti-Ig results in prompt and selective changes in phospholipid metabolism that appear to be correlated with B lymphocyte proliferation.  相似文献   

19.
We have studied the activation of human resting B cells by a carbohydrate antigen, mannan, with a polymannose branched repetitive structure. Mannan has been extracted from the cell wall of the Candida albicans yeast. For this purpose, dense G0 B lymphocytes were purified from tonsils. Mannan antigen was shown to trigger B cell activation, since an increase of cell volume and RNA synthesis occurred. B cell proliferation was observed following addition of recombinant interleukin 2, but not following addition of recombinant interleukin 4 or low-molecular-weight BCGF. The B cell activation appears to be mannan-specific since B cells obtained from mannan-sensitized subjects but not from unsensitized subjects were responsive. The observation that mannan antigen can directly activate specific dense B lymphocytes can be related to the previous observation that the in vitro anti-mannan antibody production does not require a cognate T-B cell interaction.  相似文献   

20.
The response of T cells to minor lymphocyte-stimulating locus (Mls) determinants remains poorly understood with respect to the antigenic determinants responsible for T cell stimulation and the types of APC capable of stimulating the response. In this report, we demonstrate that highly purified dendritic cells (DC) as well as B cells have the capacity to stimulate Mls-specific responses. Unseparated spleen cells, purified DC, resting B cells, and activated B cells were compared for their capacity to stimulate several Mls-reactive T cell hybridomas. Whereas the entire panel of Mls-reactive T cell hybridomas was stimulated strongly by unseparated spleen cells and activated B cells, the hybridomas responded only weakly to purified DC or resting B cells. Activation of resting B cells with either B cell stimulatory factor-1 (1 day pre-treatment) or LPS/dextran (2 or 3 day pre-treatment) greatly augmented their Mls-stimulatory capacity. In contrast, the Mls-stimulatory capacity of DC was not augmented by a 1-day pre-treatment with either B cell stimulatory factor-1 or supernatant from the DC-induced primary anti-Mls-MLR. In the primary anti-Mls-MLR, both purified DC and LPS/dextran-stimulated B blasts were found to elicit vigorous T cell proliferative responses. Much weaker responses were elicited by unseparated spleen cells. The stimulation of the primary anti-Mls-MLR by purified DC was further confirmed by producing Mls-specific T cell clones which were preferentially stimulated by DC. Autologous (Mlsb) DC were found to markedly enhance the primary anti-Mls-MLR response to small numbers of Mlsa B blasts. Thus, DC possess other "accessory cell" properties that augment the primary anti-Mls-MLR despite the predicted low level of Mls determinant expression on DC based on the results obtained with Mls-reactive hybridomas. Possible accessory cell properties of DC relevant to this phenomenon are discussed.  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