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.
The experiments in this paper demonstrate that monoclonal anti-Lyb2.1 antibody enhances the proliferative response of anti-immunoglobulin (anti-Ig)-stimulated but not of dextran sulfate-stimulated B cells. The magnitude of this enhanced B-cell proliferation is comparable to that induced by BSF-1 on anti-Ig-stimulated cells. The ability of this antibody to enhance B-cell proliferation does not result from its ability to neutralize the suppressive effects on B-cell activation that is mediated by the Fc fragment of anti-Ig antibody as it is equally as effective in enhancing B-cell proliferative responses stimulated by F(ab')2 fragments of anti-Ig. BSF-1 and Anti-Lyb2.1 appear to stimulate nonoverlapping pathways leading to B-cell activation since the enhanced responses induced by the combination of BSF-1 and anti-Lyb2.1 on anti-Ig-stimulated cells are additive even when maximum quantities of these activators are employed. There is also a marked difference in their activity on T cells; while BSF-1 can enhance T-cell proliferation in synergy with phorbol ester, anti-Lyb2.1 is ineffective in this regard. These data, while consistent with the suggestion that the Lyb2 surface determinant on B cells may be involved in B-cell activation, indicate that it is distinct from the receptors for BSF-1 or BCGF-II.  相似文献   

2.
Previous reports of the response of B lymphocytes to soluble anti-immunoglobulin (anti-Ig) antibodies have yielded conflicting data. While most studies show activation of B cells, others have shown inhibitory effects. In the assay reported in this report, we were able to obtain widely diverse responses of human B-cell populations to anti-Ig antibody. An explanation of this variability was established by resort to an animal (murine) model. Mice maintained in a pathogen-free environment failed to respond or responded only weakly to anti-Ig antibody. Mice which had previously received heavy antigenic stimulation, but at the time of the experiment were not undergoing any known challenge, showed a marked positive response. Mice deliberately challenged with lipopolysaccharide (LPS) 24 hr prior to anti-Ig antibody exposure showed a high background mitogenesis in control cultures, which was inhibited by anti-Ig antibody. This preliminary study suggests that response to anti-Ig antibody differs in each phase of B-cell differentiation. In future studies it is hoped that this variability in response can be used to characterize different subsets of B-cell differentiation separated by physical or phenotypic parameters.  相似文献   

3.
In order to compare and contrast the requirements of virgin and memory B cells for B-cell differentiation factors, a model system was developed in which low-density rat B cells isolated from 4-week primed antigen-draining lymph nodes were cultured in vitro. This large low-density cell population contained B cells which were 90% surface IgM positive and 60% IgD positive and showed moderately elevated Ia staining. When the cell population was stimulated with antigen plus lymphokines or lymphokines alone, antigen-specific IgG antibody was secreted; this was used as a measure of memory cell differentiation. When the cell population was stimulated with mitogen (lipopolysaccharide plus dextran sulfate) plus lymphokines, polyclonal IgG and IgM secretion was seen and was used as a measure of virgin B-cell differentiation. Using this system, we found that lymphokines contained in a Con A-induced rat spleen cell supernatant (CSN) were sufficient to drive both memory and virgin B-cell differentiation. In contrast, lymphokines contained in the supernatant from the murine T-cell hybridoma B151K12 (B151CFS) were able to induce large amounts of polyclonal IgM and IgG secretion but did not support memory B-cell differentiation. When recombinant human IL-2 was added to these cultures, it acted synergistically to augment virgin B-cell differentiation, but this combination of lymphokines was still not able to support memory B-cell differentiation. Furthermore, recombinant rat interferon-gamma and a commercial source of human BCGF, with or without IL-2, were unable to promote significant virgin or memory B-cell differentiation. These data support the hypothesis that memory B cells and virgin B cells differ in their lymphokine requirements for differentiation into antibody-secreting cells.  相似文献   

4.
We have utilized several B-cell lymphomas that are growth inhibited by anti-Ig reagents as models for tolerance induction. In a previous communication, we demonstrated that the growth inhibition by anti-Ig can be partially prevented by the recombinant lymphokine, IL-4. In this paper, we report that complete protection of B lymphomas from anti-Ig was provided by a type 2 helper cell clone, D10.G4, when these T cells were activated by monoclonal anti-CD3. Conditioned medium from anti-CD3-stimulated D10.G4 cells also provided protection from anti-Ig. In contrast, little protection was observed with activated cells from a type 1 T-cell clone, A.E7. Furthermore, we show that combinations of IL-4 and tumor necrosis factors (both TNF alpha and TNF beta), as well as IL-4, effected partial protection by themselves and enhanced the activity of the other lymphokine if used in a pretreatment protocol. However, anti-cytokine antibodies were ineffective at reversing the T-cell-mediated protection. The possibility that direct T:B-cell contact mediates part of the protective signal is discussed.  相似文献   

5.
Bone marrow-derived lymphocytes (B cells) with or without receptors for a third component of complement (CR) were studied in their responsiveness to the F(ab′)2 fragment of antiimmunoglobulins (anti-Ig). Spleen cells from C57BL/6J mice were fractionated by the centrifugation over Ficoll-Hypaque density gradient after they were rosetted with erythrocyte-antibody complement complexes. The cells in the interface fraction responded poorly to anti-Ig, while the cells in the pellet fraction responded well. The low responsiveness of CR(?) B cells was confirmed by assaying the responsiveness of cells passed through a Sephadex G-10-complement column. Reduced response of CR(?) B cells could not be explained by the depletion of helper or accessory cells. The relationship between CR, B-cell differentiation and proliferative capacity of B cells is discussed.  相似文献   

6.
Investigation of the activation of splenic B cells by anti-immunoglobulin (Ig) antibody has enabled us to characterize the anti-Ig-responsive B cell and to analyze the phenotypic changes which accompany proliferation and differentiation. The anti-Ig antibody-responsive B cell population was characterized by the expression of high levels of the B2 antigen and represented approximately 40% of splenic B cells. Brisk mitogenesis which peaked at 3 to 4 days was induced by anti-Ig antibody. The proliferative phase was characterized phenotypically by a dramatic decline in B2 antigen expression, with most cells showing no detectable B2 by 4 days post-activation. The other hallmark of this phase was de novo expression of a group of "activation antigens." These included the B cell-restricted antigens B-LAST 1, BB1, and B5, and the T cell-associated interleukin 2 receptor and T12 antigens. Concomitantly, B1, B4, and Ia expression increased, the increase being roughly proportional to the increase in cell size. After day 4, the mitogenic response progressively diminished, while Ig synthesis increased. During this differentiation phase, cell surface antigens again displayed a distinct sequence of changes. The five activation antigens and the B1, B4, and Ia antigens began to decrease. However, two markers, T10 and PCA-1, which are found on plasmacytomas, appeared and their level of expression steadily increased. These changes and the appearance of morphologically identifiable plasma cells required the presence of T cells in this system. T cell supernatants alone induced Ig secretion but did not induce expression of PCA-1 or the appearance of cells with plasma cell morphology. The culture system developed in this study has allowed us to analyze the antigenic changes following activation by anti-Ig antibody. This sequence of changes has not only permitted the identification of antigens which, by their appearance at distinct stages may have an important role in proliferation and differentiation of B cells, but also provides us with the means of studying the function of each antigen.  相似文献   

7.
Regulatory role of CD19 molecules in B-cell activation and differentiation   总被引:5,自引:0,他引:5  
Cluster of differentiation ([CD]) 19 antigens are B-cell-specific molecules expressed on virtually all human cells of the B-lymphocyte lineage except plasma cells. We produced a new anti-CD19 monoclonal antibody (McAb), CLB-CD19, that was used to study the role of CD19 molecules in B-cell activation. Anti-CD19 McAb induced mobilization of free intracellular calcium ([Ca2+]i) in Daudi cells, but not in normal spleen or tonsillar B cells, for which crosslinking with a second anti-mouse Ig antibody was not required. Anti-CD19 McAb inhibited B-cell proliferation induced by anti-IgM coupled to Sepharose beads. This inhibitory effect was overcome by the addition of nonmitogenic concentrations of phorbol myristate acetate. Anti-CD19 McAb did not interfere with Staphylococcus aureus- or B-cell growth factor-induced B-cell proliferation. Anti-CD19 McAb inhibited T-cell-dependent polyclonal B-cell differentiation in pokeweed mitogen-, interleukin 2-, or anti-CD3-driven culture systems. Delayed addition studies showed that once differentiation of B cells was induced, CD19 molecules lost their regulating function. Taken together, our results indicate that CD19 molecules play a regulatory role in B-cell proliferation and differentiation.  相似文献   

8.
Mouse spleen cells were depleted of immunoglobulin (Ig)-bearing B cells by rosetting with glutaraldehyde-fixed, tannic acid-treated RBC coupled with antibody to mouse Ig (anti-Ig) and removing the rosetted cells by density gradient centrifugation. The method was routinely greater than 90% effective in removing B cells as assayed by the failure of anti-Ig rosette-depleted primed spleen cells to generate antibody-producing cells in vitro in response to specific antigen or of anti-Ig rosette-depleted nonprimed spleen cells to generate a polyclonal antibody response. T cells were not removed by the rosetting procedure as measured by helper T-cell activity. The greater effectiveness of the rosetting procedure in removing potential IgG-secreting, non-IgM-bearing B cells is shown relative to other commonly used B-cell depletion procedures. Because the RBC in the rosetting reagent are fixed with glutaraldehyde, the rosetting reagent is stable for many months. Such stability makes constantly available a convenient means for B-cell removal, as well as reducing consumption of antisera.  相似文献   

9.
It is shown that the anti-trinitrophenyl (TNP) response of human B cells to trinitrophenyl polyacrylamide beads (TNP-PAA) is monocyte dependent. This response is abolished by extensive adherent cell depletion and restored by the addition of monocytes. The optimal response is obtained with 3% monocytes, higher numbers being suppressive. Supernatants from muramyl dipeptide (MDP)-activated monocytes can restore the response of monocyte-depleted preparations even when cells are cultured at suboptimal concentration. A partially purified preparation of interleukin (IL-1) has a comparable restorative ability. The following arguments suggest that monocytes do not function as antigen-presenting cells for this particulate antigen: (i) anti-genpulsed monocytes induce neither an anti-TNP response nor a specific T-cell proliferative response; (ii) allogeneic monocytes function as well as autologous monocytes to restore the response of nonadherent cells; (iii) HLA-DR-negative cells from the human leukemia cell line K562 can replace monocytes for this response. Monocyte supernatants do not replace T cells for the response of B-enriched lymphocytes, showing that T cells are directly involved in B-cell activation.  相似文献   

10.
The mitogenic responses of separated rabbit lymphocyte populations functionally analogous to mouse T and B cells have been tested in vitro. Purified T cells were prepared by passage over nylon wool (NW) and purified B cells prepared by treatment with antithymocyte serum and complement (ATS + C). ATS + C kills 70% of peripheral blood lymphocytes (PBL's) and 50% of the spleen cells while passage over NW yields 40% of the applied PBL's and 5–23% of the applied spleen cells. NW-purified T cells from the spleen or PBL's respond fully to concanavalin A (Con A) but have a reduced response to phytohemaglutinin (PHA) and little or no response to goat anti-rabbit immunoglobulin (anti-Ig). PBL's that survive ATS + C (B cells) are stimulated by anti-Ig but not by Con A or PHA. B cells purified from spleen do not respond to Con A or PHA but will respond to anti-Ig under appropriate conditions. A full spleen B-cell response to anti-Ig required removal of Ig produced by the cultures that blocked anti-Ig stimulation. It is concluded that, for rabbit lymphocytes, Con A and PHA are primarily T-cell mitogens and that anti-Ig is primarily a B-cell mitogen. However, the mitogen response of unfractionated PBL or spleen cell populations indicates an overlap in reactivity. This could be due to cells sharing T and B properties, alteration of cell populations by the fractionation procedures used, or recruitment of one population in the presence of a mitogenic response of the other population.  相似文献   

11.
Overnight exposure of adult splenic B cells to anti-Ig, a surrogate for antigen/tolerogen, can result in a hyporesponsive state in terms of antibody synthesis. Since B cells treated with either intact of F(ab')2 fragments of anti-Ig will exit the G0 phase of the cell cycle and enter G1 or S, respectively, we examined which steps in B-cell activation were required for this form of hyporesponsiveness. We found that B-cell hyporesponsiveness could be induced under conditions leading to either abortive or productive B-cell cycle progression, depending on the immunogenic challenge employed. Thus, PMA + ionomycin, concanavalin A, PMA alone, or ionomycin alone induced hyporesponsiveness. Each of these reagents is able to drive B-cell exit from G0 into G1 and cause class II hyperexpression. We next examined the effect of cyclosporin A (CSA), a reagent that blocks anti-Ig but not by PMA-induced class II hyperexpression. Interestingly, CSA only interfered with the induction of B-cell hyporesponsiveness with anti-Ig. These results suggest that upregulation of MHC class II may be coincident with a CSA-sensitive tolerance pathway in B cells stimulated by anti-Ig. Finally, IL-4 pretreatment was found to ablate hyporesponsiveness induced by either intact anti-Ig or PMA. These results parallel the Fc-dependent induction of hyporesponsiveness reported earlier (G. Warner and D. W. Scott, J. Immunol. 146, 2185, 1991). We propose that crosslinking of surface Ig, leading to cell cycle progression out of G0 as well as class II hyperexpression, in the absence of a cognate T cell signal, leads to B-cell hyporesponsiveness.  相似文献   

12.
Human B cells can be activated with monoclonal antibodies (mAb) to surface IgM receptors or mAb to a 35-kilodalton B cell differentiation antigen, Bp35 (CD20). We compared anti-Ig-induced B cell activation with B cell triggering by anti-Bp35. Both anti-Ig- and anti-Bp35-dependent proliferation were augmented by the same co-stimulants, including a partially purified BCGF, recombinant IL 1, TPA, or each other. When anti-Bp35 and anti-Ig were used together to induce proliferation of tonsillar B cells, the strongest response was observed when anti-Bp35 was added 12 to 24 hr before anti-Ig. Anti-Bp35 also was found to act most effectively when added before the BCGF. Blood and tonsillar B cells differed in their proliferative response to anti-Ig or anti-Bp35: unlike dense tonsillar B cells, which consistently proliferated in response to either stimulus, blood B cells from many donors proliferated in response to anti-Ig but not to anti-Bp35 even in the presence of other co-stimuli. Dense tonsillar B cells that proliferate in response to anti-Bp35 appeared to be at a more activated stage than unresponsive blood B cells because they expressed higher levels of HLA class II molecules than blood B cells. Pretreatment of blood B cells with anti-Bp35 converted them to an HLA-DR(bri) phenotype and made them more responsive to anti-Ig-induced proliferation. These results suggest that B cells at different stages of differentiation differ in their response to anti-Bp35 and anti-Ig. The Bp35 surface polypeptide may play an early role in the activation of B cells prior to antigen or other signals.  相似文献   

13.
Salmonella typhimurium mitogen (STM) is a polyclonal activator of rat B lymphocytes, triggering them to proliferate, but not differentiate, to antibody-secreting cells. When lymphokines in the form of a supernatant from Con A-stimulated splenocytes (CAS) are added to B cell cultures activated by STM, only a small number of cells are driven to differentiate. Only with the addition of a third signal provided by the polyanionic polysaccharide dextran sulfate (DXS) is significant rat B cell differentiation observed. In this study, we have shown that this requirement for DXS is not unique to the STM mitogen. LPS, Staphylococcus aureus Cowan I-fixed cells, and anti-Ig antibody all induced rat B cell proliferation with little differentiation, even in the presence of CAS. DXS was necessary to induce differentiation in all cultures costimulated with mitogen and CAS. The requirement for DXS for optimal B cell differentiation is also observed with other lymphokine preparations such as the supernatants from PMA-stimulated EL-4 cells and PHA-stimulated human T cells. Furthermore, this augmentative effect of DXS in rat B cell differentiation was not confined to polyclonal activation systems. Ag-specific IgG secretion was also increased when DXS was added to Ag and CAS costimulated cultures of B cells harvested from the draining lymph nodes of rats immunized with DNP-keyhole limpet hemocyanin. Within the polyclonal activation system, a method of staged additions of STM, DXS, and CAS to B cell cultures was used to investigate the role of DXS during B cell differentiation. Optimal differentiation occurred only when DXS was present in the B cell cultures in conjunction with CAS. The augmentation in differentiation seen with DXS did not appear to be due to the recruitment of an additional CAS-responsive B cell subset, because cycling, low density B cell blasts showed large increases in IgM secretion with subsequent exposure to DXS and CAS. These studies suggest tha DXS acts as a cofactor to various differentiation factors, augmenting polyclonal and Ag-specific rat B cell differentiation. The relevance of DXS to in vivo immune responses is discussed.  相似文献   

14.
We have examined the requirement for major histocompatibility complex (MHC)-restricted T-cell help in the secondary in vivo antibody response to phosphocholine (PC). The memory response to PC has been demonstrated previously to be comprised of T15-dominant IgM and IgG3 plaque-forming cells (PFC) derived primarily from the Lyb-5+ B-cell subset, and IgG1 and IgG2 PFC, few of which bear the T15 idiotype and are predominantly derived from the Lyb-5- B-cell subset. Using carrier-primed (A X B)F1 T cells which have matured in a parentA chimeric environment so that "self" recognition is of the MHC determinants of parentA but not parentB, we have found that parentA PC-primed B cells, but not parentB PC-primed B cells, are activated. Even in the presence of an ongoing parentA anti-PC response, parentB PC-primed B cells were not activated, indicating that the restriction was between the helper T cell and the B cell, not between T-helper and accessory cells. MHC-restricted T-cell help was required by B cells producing T15+ and T15- IgM, IgG3, IgG1, and IgG2 responses.  相似文献   

15.
The abilities of B cells from 24 young (mean 26 yr) and 24 elderly (mean 86 yr) humans to proliferate and differentiate into immunoglobulin-secreting cells (ISC) were investigated. Initial studies in young subjects demonstrated that a Staph protein A (SpA)-driven system could simultaneously assess the proliferative and differentiative capabilities of B cells resulting in IgM production. B cell proliferative responses were found to be partially T cell-dependent, whereas differentiation was absolutely T cell-dependent. Also, no significant differences could be detected in the abilities of nonproliferating allogeneic and autologous T cells to support B cell responsiveness. Although B cells from elderly subjects continuously exposed to SpA displayed proliferative responses equal to young subjects, the differentiation of B cells from elderly subjects into IgM ISC was markedly reduced as compared to young subjects. Analyses of results from co-culture experiments showed that the differentiation impairments of B cells from some elderly subjects could be partially corrected by allogeneic T cells from young subjects, whereas the impairments of others were more refractory. Moreover, T cells from elderly subjects were able to promote the differentiation of B cells from young subjects. Other experiments in elderly subjects showed that significant impairments of B and T cell functions rarely coexisted and that compensatory increases in B or T cell function were not evident. Thus, B cells from certain elderly humans have intrinsic impairments of differentiation required for optimal IgM production even though activation and proliferation remain normal in the presence of SpA. These impairments in differentiation are sometimes improved by T cells from young subjects, although in some elderly individuals, the differentiative impairments fail to be reversed.  相似文献   

16.
Anti-mu antibody preparations have been found to exert both positive and negative effects on B cell activation and differentiation. To explore these paradoxical influences of IgM cross-linkage on human B cells, three gamma 1 kappa murine monoclonal antibodies specific for human mu-chains (DA4.4, AB6.4, 145.8) were examined for their comparative effects on activation of B cells and inhibition of terminal plasma cell differentiation. All three antibodies appeared equally efficient in immunoprecipitation of surface IgM molecules; however, fluorescence-activated cell sorter analysis revealed that the DA4.4 and AB6.4 antibodies saturated the B cell surface IgM at slightly lower concentrations than did the 145.8 antibody. When the affinity-purified antibodies were added in varying concentrations to cultures of small resting B cells, all three antibodies induced B cell enlargement and DNA synthesis, but with varying degrees of efficiency (DA4.4 greater than AB6.4 much greater than 145.8). In striking contrast, large B cells isolated either by FACS or density gradient separation were unresponsive. The anti-mu-induced proliferative response of small B cells required relatively high B cell densities, but not T cells or the Fc portion of the antibody molecules. The maximal proliferative response was obtained during the third day of culture, and the response curve suggested that anti-mu induced only one round of B cell replication. All three antibodies were capable of completely inhibiting T cell factor-induced differentiation of large B cells into IgM plasma cells; both F(ab')2 fragments and intact anti-mu antibodies were effective in final concentrations as low as 1 microgram/ml. Significant suppression of IgG and IgA plasma cell differentiation was also achieved, but required higher concentrations of the anti-mu antibodies. For each antibody, there was a close correlation between the efficiency of inducing small B cell proliferation and of inhibiting large B cell differentiation into plasma cells. The results show that the B cell response to cross-linkage of cell surface IgM varies according to the differentiation stage. We postulate that the mature resting B cell represents the only stage in the life history of the B cell during which surface Ig cross-linkage leads to a positive signal, negative signals being the rule at other stages in B cell replication and differentiation.  相似文献   

17.
To further characterize a human B-cell growth factor (BCGF) produced by phytohemagglutinin (PHA) P-stimulated peripheral blood T cells, a partially purified preparation of this material was tested in a number of murine assays for B-cell stimulatory factors (BSF). Human BCGF lacked murine BSF-1 activity as assessed via the induction of polyclonal proliferation of anti-IgM-stimulated murine B cells; however, this material consistently augmented the proliferative response of murine B cells to anti-IgM and a saturating dose of murine BSF-1. Human BCGF also induced proliferation in unstimulated murine B cells, and augmented the proliferative response of dextran sulfate activated murine B cells. Human BCGF is therefore capable of causing proliferation of unstimulated and activated murine B cells, and by these criteria closely resembles murine BCGF II. In contrast to murine BCGF II, however, human BCGF failed to stimulate proliferation or immunoglobulin (Ig) secretion by murine BCL1 B lymphoma cells. A murine analog of this human BCGF showing the same pattern of biological responses was found in concanavalin A-stimulated supernatants of the murine MB2.1 T-cell line and D9-Cl T-cell hybridoma. The active component of the human BCGF preparation was not due to contaminating PHA, interleukin 1, interleukin 2; interferon-gamma, or endotoxin. Comparison between the above human BCGF and a commonly used source of murine BCGF II, i.e., supernatant from antigen-stimulated D10.G4.1 T cells, provided information suggestive of BCGF II heterogeneity. Both human BCGF and D10.G4.1 supernatant caused proliferation of unstimulated and dextran sulfate-stimulated murine B cells; however, only the human BCGF preparation augmented the proliferative response of murine B cells to anti-IgM and a saturating dose of murine BSF-1, and only the D10.G4.1 supernatant stimulated BCL1 cell proliferation and immunoglobulin secretion. The data therefore indicate that the different assays for BCGF II used in this study respond to different factors, and suggest the existence of two BCGF II-like activities.  相似文献   

18.
Genomic-scale gene expression analysis provides views of biological processes as a whole that are difficult to obtain using traditional single-gene experimental approaches. In the case of differentiating systems, gene expression profiting can define a stage of differentiation by the characteristic expression of hundreds of genes. Using specialized DNA microarrays termed 'Lymphochips', gene expression during mature B-cell differentiation has been defined. Germinal centre B cells represent a stage of differentiation that can be defined by a gene expression signature that is not shared by other highly proliferative B-cell populations such as mitogenically activated peripheral blood B cells. The germinal centre gene expression signature is maintained to a significant degree in lymphoma cell lines derived from this stage of differentiation, demonstrating that this gene expression programme does not require ongoing interactions with other germinal centre cell types. Analysis of representative cDNA libraries prepared from resting and activated peripheral blood B cells, germinal centre centroblasts, centrocytes and tonsillar memory B cells has confirmed and extended the results of DNA microarray gene expression analysis.  相似文献   

19.
Human blood lymphocytes were separated into T and non-T cells and cultured with pokeweed mitogen (PWM). It was found that in the absence of T cells no differentiation of B cells into immunoglobulin-containing blasts and plasma cells took place. Moreover, the cell yields and the rate of DNA synthesis and blast transformation were very low. The influence of T cells on PWM-induced B-lymphocyte differentiation was studied in mixtures of T/non-T cells at various ratios. Addition of even a few T lymphocytes caused a considerable stimulation of B cells by all parameters used. The responses of T/non-T mixtures of the original cellular composition were of the same order as those of cultures of unseparated cells. It is concluded that the differentiation of human blood B lymphocytes into cells actively synthesizing immunoglobulins, as induced by PWM, is strongly dependent upon the presence of T cells.  相似文献   

20.
B-cell antigen receptor signaling is initiated upon binding of the antigen to membrane-bound immunoblobulin (Ig), and the anti-Ig antibody (Ab) mimics this signaling. In B cells latently infected with Epstein-Barr virus (EBV), the same signals induce virus activation. We examine here whether rheumatoid factors (RFs), autoantibodies directed against the Fc portion of IgG, induce EBV and B-cell activation. As a source of RFs, RF-producing lymphoblastoid cell line (LCL) clones were isolated from peripheral blood mononuclear cells (PBMC) and synovial cells from patients with rheumatoid arthritis (RA) by EBV transformation. Burkitt's lymphoma-derived Akata cells, which are highly responsive to EBV activation by anti-Ig Abs, were used for the assay of EBV activation. Akata cells expressed IgG3 as membrane-bound Ig. RFs from a synovium-derived LCL were directed to IgG3 and induced EBV activation in 16 to 18% of Akata cells, whereas RFs from another synovium-derived LCL were directed to IgG1 and did not induce EBV activation. Pretreatment of RFs with the purified Fc fragment of human IgG completely abolished EBV activation. Furthermore, B-cell activation was assessed by incorporation of [3H]thymidine. RFs from synovium-derived LCLs efficiently induced B-cell activation, and the addition of CD40 ligand had a synergistic effect. On the other hand, RFs from PBMC-derived LCLs were polyreactive, had a lower affinity to IgG, and did not induce EBV and B-cell activation. The present findings imply a possible role for RFs as EBV and B-cell activators.  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