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.
Adjuvanticity of nystatin, one of the polyenic antifungal antibiotics having as its primary target the membrane sterol of eukaryotic cells, was investigated by examining its effect on several functions of mouse spleen cells relevant to immunological phenomena in vitro. Nystatin was found to stimulate significantly DNA synthesis in thymus-independent (B) cells but not in thymus-dependent (T) cells. Like the other B-cell mitogens such as bacterial lipopolysaccharide (LPS), nystatin elicited nonspecifically polyclonal antibody synthesis in mouse spleen cell cultures, and also restored antibody response of T cell-deficient spleen cells of congenitally athymic nude mice to heterologous erythrocytes (RBC; thymus-dependent antigen). Thus, nystatin and LPS appeared to cause similar changes in the functions of spleen cells relevant to immunological events. However, antagonism but no additive effect in the adjuvanticity was revealed between the two adjuvants. As an interesting finding, the polyclonal generation of anti-RBC antibody-forming cells (AFC) in the spleen cell cultures by stimulation with B-cell mitogen, i.e., either nystatin or LPS, was not inhibited at all by inclusion of any anti-RBC antiserum, whereas, as is well known, the generation of AFC by stimulation with the antigen was specifically suppressed by the corresponding antiserum, indicating a difference in the genesis between the mitogen-induced AFC and the antigen-induced AFC.  相似文献   

2.
Phospholipase Cgamma2 (PLCgamma2) is a critical signaling effector of the B-cell receptor (BCR). Here we show that PLCgamma2 deficiency impedes early B-cell development, resulting in an increase of B220+ CD43+ BP-1+ CD24hi pre-BCR+ large pre-B cells. PLCgamma2 deficiency impairs pre-BCR-mediated functions, leading to enhanced interleukin-7 (IL-7) signaling and elevated levels of RAGs in the selected large pre-B cells. Consequently, PLCgamma2 deficiency renders large pre-B cells susceptible to transformation, resulting in dramatic acceleration of Myc-induced lymphomagenesis. PLCgamma2(-/-) Emu-Myc transgenic mice mainly develop lymphomas of B220+ CD43+ BP-1+ CD24hi pre-BCR+ large pre-B-cell origin, which are uncommon in wild-type Emu-Myc transgenics. Furthermore, lymphomas from PLCgamma2(-/-) Emu-Myc transgenic mice exhibited a loss of p27Kip1 and often displayed alterations in Arf or p53. Thus, PLCgamma2 plays an important role in pre-BCR-mediated early B-cell development, and its deficiency leads to markedly increased pools of the most at-risk large pre-B cells, which display hyperresponsiveness to IL-7 and express high levels of RAGs, making them prone to secondary mutations and Myc-induced malignancy.  相似文献   

3.
Changes in suppressor, helper, and B-cell functions in aging mice   总被引:2,自引:0,他引:2  
Suppressor, T-helper, and B-cell functions were measured in spleen cells derived from individual mice of varying ages. Suppressor activity was assessed by the ability of test cells to suppress the secondary anti-hapten response of young-adult spleen cells in mixed cultures. T-Helper-cell activity was assessed by measuring the ability of T cells activated with carrier-anti-carrier immune complexes to restore the response of hapten-primed, T-depleted syngeneic cells. B-Cell activity was assessed by measuring the response to the T-independent antigen DNP-Ficoll. It was found that the suppressor function increased rapidly, reached a peak in middle-aged mice, and remained elevated thereafter. In contrast, T-helper and B-cell functions declined at a constant rate from young adulthood to old age. An exception to this pattern was found in a distinct subpopulation of aged mice, which had splenomegaly and extremely low suppressor, T-helper, and B-cell functions.  相似文献   

4.
The error frequency and mutational specificity associated with Escherichia coli uracil-initiated base excision repair were measured using an M13mp2 lacZalpha DNA-based reversion assay. Repair was detected in cell-free extracts utilizing a form I DNA substrate containing a site-specific uracil residue. The rate and extent of complete uracil-DNA repair were measured using uracil-DNA glycosylase (Ung)- or double-strand uracil-DNA glycosylase (Dug)-proficient and -deficient isogenic E. coli cells. In reactions utilizing E. coli NR8051 (ung(+) dug(+)), approximately 80% of the uracil-DNA was repaired, whereas about 20% repair was observed using NR8052 (ung(-) dug(+)) cells. The Ung-deficient reaction was insensitive to inhibition by the PBS2 uracil-DNA glycosylase inhibitor protein, implying the involvement of Dug activity. Under both conditions, repaired form I DNA accumulated in conjunction with limited DNA synthesis associated with a repair patch size of 1-20 nucleotides. Reactions conducted with E. coli BH156 (ung(-) dug(+)), BH157 (ung(+) dug(-)), and BH158 (ung(-) dug(-)) cells provided direct evidence for the involvement of Dug in uracil-DNA repair. The rate of repair was 5-fold greater in the Ung-proficient than in the Ung-deficient reactions, while repair was not detected in reactions deficient in both Ung and Dug. The base substitution reversion frequency associated with uracil-DNA repair was determined to be approximately 5.5 x 10(-)(4) with transversion mutations dominating the mutational spectrum. In the presence of Dug, inactivation of Ung resulted in up to a 7.3-fold increase in mutation frequency without a dramatic change in mutational specificity.  相似文献   

5.
B-cell clonal expansions (BCE) in young mice are transient, detectable for less than 4 weeks. In contrast, BCE in old mice persist more than 2 months. The greater persistence of BCE in old mice does not appear to be due to the age of the host as the survival of phenyloxazolone chicken serum albumin-induced BCE in most old mice was shorter than in young mice. This raises the possibility that persistent BCE seen in old mice develop over time from transient BCE present earlier in life. To test this hypothesis, young C57BL/6 mice were immunized with hen egg lysozyme (HEL) during the first year of life. By 28 months of age, the majority of these mice had developed a benign, persistent BCE associated with a HEL-specific serum mIg. We also investigated whether benign, persistent BCE, present in 18-month-old mice, can evolve into B-cell lymphomas. We observed that four of eight C57BL/6 mice that survived to 29 months of age had developed diffuse large cell lymphomas. In three of these mice, this diagnosis was made by microscopic analysis of the lymphoid organs. In one mouse, a macroscopic lymphoma was present that permitted us to demonstrate that the IgH mRNA CDR3 length and sequence in the malignant lymphoma was derived from a persistent BCE present 11 months earlier. Together these observations are consistent with the hypothesis that stepwise accumulation of genetic alterations combined with Darwinian selection underlies the evolution of B cells from transient BCE in young mice into persistent BCE, serum mIg, and B-cell lymphomas observed in older mice.  相似文献   

6.
BACKGROUND: CD2 is expressed by T and natural killer (NK) cells and has been reported in T/NK cell lineage neoplasms as well as in immature B-lymphoblastic and myeloid leukemias. Although CD2+ B-cells have been identified in normal fetal and postnatal thymus, they have not been reported in adults. METHODS: We retrospectively reviewed flow cytometric immunophenotypic data on consecutive low-grade B-cell leukemias and lymphomas to investigate the frequency of CD2 expression. We also reviewed samples from normal healthy donors to determine whether there is a normal CD2+ B-cell population. RESULTS: CD2 expression (partial or complete) was observed in 13 of 83 (16%) chronic lymphocytic leukemias (CLL), 16 of 29 (55%) follicle center lymphomas (FCL), 3 of 12 (25%) hairy cell leukemias (HCL), 0 of 6 mantle cell lymphomas (MCL), 8 of 28 (29%) large cell lymphomas (LCL), and in 0 of 5 marginal zone/mucosa-associated lymphoid tissue lymphomas (MZL/MALT). We determined that 5.74 +/- 2.46% (mean +/- SD) of normal peripheral blood B cells and 6.48 +/- 1.62 % (mean +/- SD) of normal bone marrow B cells coexpress CD2. CONCLUSIONS: CD2 expression in B-cell neoplasia is a more prevalent phenomenon than previously appreciated. Normal CD2+ B-cell populations are observed in adults and may represent the nonmalignant counterpart of CD2+ B-cell neoplasms.  相似文献   

7.
The inbred mouse strain CWD/Agl has a high incidence of spontaneous B-cell lymphomas characterized by gross splenomegaly and lymph node enlargement. The endogenous ecotropic retrovirus of CWD/Agl mice is expressed in the spleen within the first 2 weeks of age and in the thymus by 1 month of age. Endogenous xenotropic virus is expressed in the spleen and bone marrow of the earliest age group examined (4 months). Restriction enzyme analysis of DNA extracted from tumorous tissues suggests that mink cell focus-forming viruses are not required for B-cell lymphomagenesis in CWD/Agl mice. CWD/Agl mice provide an important new experimental model for the study of B-cell lymphoma.  相似文献   

8.
During embryogenesis, the development of tissues, organs and systems, including the immune system, is particularly susceptible to the effects of noxious agents. We examined the effects of prenatal (in utero) exposure to WiFi signals on pregnancy outcome and the immune B-cell compartment, including antibody production. Sixteen mated (plug-positive) female mice were assigned to each of the following groups: cage control, sham-exposed and microwave-exposed (WiFi signals at 2.45 GHz, whole body, SAR 4 W/kg, 2 h/day, 14 consecutive days starting 5 days after mating). No effects due to exposure to WiFi signals during pregnancy on mating success, number of newborns/mother and body weight at birth were found. Newborn mice were left to grow until 5 or 26 weeks of age, when immunological analyses were performed. No differences due to exposure were found in spleen cell number, B-cell frequency or antibody serum levels. When challenged in vitro with LPS, B cells from all groups produced comparable amounts of IgM and IgG, and proliferated at a similar level. All these findings were consistently observed in the female and male offspring at both juvenile (5 weeks) and adult (26 weeks) ages. Stress-associated effects as well as age- and/or sex-related differences were observed for several parameters. In conclusion, our results do not show any effect on pregnancy outcome or any early or late effects on B-cell differentiation and function due to prenatal exposure to WiFi signals.  相似文献   

9.
The effect of erythropoietic stimulation and suppression on the production of antibody-forming cells (AFC) in the spleen and on the B-cell migration from the bone marrow to the spleen was investigated in the CBA mice. Erythropoiesis stimulation proved to sharply increase the AFC count in the spleen and the B-cell migration from the bone marrow to the spleen 1 and 4 days after the bleeding. Erythropoiesis suppression resulted in a slight increase of the AFC count in the spleen 4 and 7 days after the transfusion of syngeneic red blood cells. However, the erythropoietic suppression inhibited the B-cell migration from the bone marrow to the spleen. Possible mechanisms of the effect of the erythropoietic stimulation and suppression on the antibody production are discussed.  相似文献   

10.
Xenoengraftment of human cells in mice with severe combined immunodeficiency (SCID) has been used as a model system to study the mechanisms of B-cell lymphomagenesis. In the study reported here, we determined that SCID mice can also be used as a model to study angiogenesis in B-cell lymphomas. The C.B-17 scid/scid mice were xenotransplanted with Epstein-Barr virus (EBV)-transformed lymphoblastoid cell lines (LCL), and we determined whether CD31, a marker found on endothelial cells, was detected in the human B-cell lymphomas that developed in these mice. Microvessel formation was identified by use of immunohistochemical staining for CD31. To assess possible mechanisms of angiogenic stimulus, we analyzed the expression of interleukin 8 (IL-8), a chemokine documented to promote angiogenesis, in non-small-cell lung cancer and bronchogenic carcinomas. We observed that a panel of LCL and LCL-lymphomas expressed IL-8 mRNA and protein. Neutralization of IL-8, however, did not inhibit lymphomagenesis, suggesting that IL-8 is not essential for angiogenesis in this model. To examine other parameters of angiogenesis, we identified expression of vascular endothelial growth factor in the lymphomas. These data suggest that angiogenesis accompanies EBV-associated B-cell lymphoma development, but IL-8 is not essential for this process. Thus, the SCID mouse model is amenable to testing of anti-angiogenic factors.  相似文献   

11.
Mice lacking c-fos develop severe osteopetrosis with deficiencies in bone remodeling and exhibit extramedullary hematopoiesis, thymic atrophy, and altered B-cell development. In this study, we have used these mice to characterize in detail the developmental potential of hematopoietic stem cells lacking c-fos and to analyze how the lymphoid differentiation is altered. In c-fos -/- mice, B-cell numbers are reduced in the spleen, lymph nodes, and the peripheral blood as a result of a marked reduction (> 90%) in the number of clonogenic B-cell precursors. In contrast, the number and lineage distribution of myeloid progenitor cells are not affected. The thymic defects observed in a large number of these mice correlate with their health status, suggesting that this may be an indirect effect of the c-fos mutation. In vitro differentiation and bone marrow reconstitution experiments demonstrated that hematopoietic stem cells lacking c-fos can give rise to all mature myeloid as well as lymphoid cells, suggesting that the observed B lymphopenia in the mutant mice is due to an altered environment. Transplantation of wild-type bone marrow cells into newborn mutant mice resulted in the establishment of a bone marrow space and subsequent correction of the B-cell defect. These results demonstrate that hematopoietic stem cells lacking Fos have full developmental potential and that the observed defect in B-cell development is most likely due to the impaired bone marrow environment as a consequence of osteopetrosis.  相似文献   

12.
Kaposi sarcoma-associated herpesvirus (KSHV) has been linked to Kaposi sarcoma and B-cell malignancies. Mechanisms of KSHV-induced oncogenesis remain elusive, however, in part due to lack of reliable in vivo models. Recently, we showed that transgenic mice expressing the KSHV latent genes, including all viral microRNAs, developed splenic B cell hyperplasia with 100% penetrance, but only a fraction converted to B cell lymphomas, suggesting that cooperative oncogenic events were missing. Myc was chosen as a possible candidate, because Myc is deregulated in many B cell lymphomas. We crossed KSHV latency locus transgenic (latency) mice to Cα Myc transgenic (Myc) mice. By itself these Myc transgenic mice develop lymphomas only rarely. In the double transgenic mice (Myc/latency) we observed plasmacytosis, severe extramedullary hematopoiesis in spleen and liver, and increased proliferation of splenocytes. Myc/latency mice developed frank lymphoma at a higher rate than single transgenic latency or Myc mice. These data indicate that the KSHV latency locus cooperates with the deregulated Myc pathways to further lymphoma progression.  相似文献   

13.
Graft-versus-host disease (GVHD) is a prevalent and potential complication of hematopoietic stem cell transplantation. An animal model, xenogeneic GVHD (X-GVHD), that mimics accurately the clinical presentation of GVHD would provide a tool for investigating the mechanism involved in disease pathogenesis. Murine models indicated that inhibiting IL-21 signaling was a good therapy to reduce GVHD by impairing T cell functions. We sought to investigate the effect of exogenous human IL-21 on the process of X-GVHD. In this study, human IL-21 was expressed by hydrodynamic gene delivery in BALB/c-Rag2-/- IL-2Rγc-/- (BRG) immunodeficient mice which were intravenously transplanted human peripheral blood mononuclear cells (hPBMCs). We found that human IL-21 exacerbated X-GVHD and resulted in rapid fatality. As early as 6 days after hPBMCs transplanted to BRG mice, a marked expansion of human CD19+ B cells, but not T cells, was observed in spleen of IL-21-treated mice. Compared with control group, IL-21 induced robust immunoglobulin secretion, which was accompanied by increased accumulation of CD19+ CD38high plasma cells in spleen. In addition, we demonstrated that B-cell depletion was able to ameliorate X-GVHD. These results are the first to find in vivo expansion and differentiation of human B cells in response to IL-21, and reveal a correlation between the expansion of B cells and the exacerbation of xenogeneic GVHD. Our findings show evidence of the involvement of B cells in X-GVHD and may have implications in the treatment of the disease  相似文献   

14.
The link between infection and autoimmunity is not yet well understood. This study was designed to evaluate if an acute viral infection known to induce type I interferon production, like influenza, can by itself be responsible for the breakdown of immune tolerance and for autoimmunity. We first tested the effects of influenza virus on B cells in vitro. We then infected different transgenic mice expressing human rheumatoid factors (RF) in the absence or in the constitutive presence of the autoantigen (human immunoglobulin G [IgG]) and young lupus-prone mice [(NZB x NZW)F(1)] with influenza virus and looked for B-cell activation. In vitro, the virus induces B-cell activation through type I interferon production by non-B cells but does not directly stimulate purified B cells. In vivo, both RF and non-RF B cells were activated in an autoantigen-independent manner. This activation was abortive since IgM and IgM-RF production levels were not increased in infected mice compared to uninfected controls, whether or not anti-influenza virus human IgG was detected and even after viral rechallenge. As in RF transgenic mice, acute viral infection of (NZB x NZW)F(1) mice induced only an abortive activation of B cells and no increase in autoantibody production compared to uninfected animals. Taken together, these experiments show that virus-induced acute type I interferon production is not able by itself to break down B-cell tolerance in both normal and autoimmune genetic backgrounds.  相似文献   

15.
In order to study the kinetics and composition of the polyclonal B-cell activation associated to malaria infection, antigen-specific and non-specific B-cell responses were evaluated in the spleens of mice infected with Plasmodium yoelii 17XL or injected with lysed erythrocytes or plasma from P. yoelii infected mice or with P. falciparum culture supernatants. Spleen/body weight ratio, numbers of nucleated spleen cells and Immunoglobulin-containing and Immunoglobulin-secreting cells increased progressively during the course of infection, in parallel to the parasitaemia. A different pattern of kinetics was observed when anti-sheep red blood cell and anti-trinitrophenylated-sheep red blood cell plaque forming cells response were studied: maximum values were observed at early stages of infection, whereas the number of total Immunoglobulin-containing and Immunoglobulin-secreting cells were not yet altered. Conversely, at the end of infection, when these latter values reached their maximum, the anti-sheep red blood cell and anti-trinitrophenylated-sheep red blood cell specific responses were normal or even infranormal. In mice injected with Plasmodium-derived material, a higher increase in antigen-specific PFC was observed, as compared to the increase of Immunoglobulin-containing and Immunoglobulin-secreting cell numbers. This suggested a "preferential" (antigen-plus mitogen-induced) stimulation of antigen-specific cells rather than a generalized non-specific (mitogen-induced) triggering of B-lymphocytes. On the basis of these and previous results, it is suggested that the polyclonal B-cell activation that takes place during the course of infection appears as a result of successive waves of antigen-specific B-cell activation.  相似文献   

16.
A loss of B-cell function in old mice was demonstrated by measuring the in vitro response of lymphoid cells to the B-cell polyclonal activator, LPS (lipopolysaccharide), and the in vivo response to the thymus-independent antigen, pneumococcal polysaccharide type III (SIII). The reduced mitogenic reactivity of lymphoid cells from old compared with young mice could not be explained by a shift in kinetics of the responding cells. When LPS cultures were carried out in the presence of colchicine, fewer cells from old mice were found to respond to the mitogenic signal. The total number of B cells assessed by labelling with either anti-immunoglobulin serum or antigen-antibody complexes was not decreased in old animals. Taken together, these results are consistent with a qualitative rather than a quantitative loss of B-cell function with age. They did not, however, exclude the possibility of depletion of an LPS-reactive sub-population of B cells. Since the number of LPS-reactive cells could not be determined directly, the antibody response of old mice to SIII was investigated. The decreased level of antibody production by old mice to SIII was not due to a shift in kinetics of the responding cells. Extracellular influences were excluded by showing that the reduced responsiveness of old spleen cells persisted after adoptive transfer into young irradiated recipients. Furthermore, pretreatment of cells from old mice with anti-Thy.1 serum and complement before transfer did not enhance their antibody-forming potential. The loss of B-cell activity with age could not, therefore, be explained in terms of an increase in T-cell-dependent suppressive effects. Support for an intrinsic defect in the B cell itself came from the demonstration of similar numbers of SIII-binding cells in normal spleens from old and young mice. Following immunisation, a shift toward low intensity binding cells was observed in spleens from both old and young mice. This shift was, however, less pronounced in the case of old cells, which is consistent with an age-related decline in transformation potential of antibody-forming-cell precursors. The conclusion was, therefore, reached that the reduction with age in B-cell as well as T-cell function is due to a qualitative rather than a quantitative defect in lymphocytes themselves.  相似文献   

17.
alpha-Galactosylceramide (alpha-GalCer), a glycolipid antigen, specifically activates natural killer T (NKT) cells by a CD1d-restricted mechanism. In this work, we found that in vivo administration of alpha-GalCer resulted in the activation of B cells in addition to NKT cells, namely, alpha-GalCer administration caused upregulation of the early activation marker, CD69, on both NKT and B cells. In addition, expression of B7.2 and I-A(b) on B cells was greatly upregulated by alpha-GalCer. However, serum levels of IgE, IgG1, and IgG2a were not significantly changed within 48 h. In the present experiments, it was also demonstrated that the upregulation of CD69 expression by alpha-GalCer was strongly blocked by anti-IL-4 monoclonal antibody. Moreover, B-cell activation by alpha-GalCer was not observed in NKT-deficient mice. These results suggested that antigen-stimulated NKT cells might play a critical role not only in early defense mechanisms but also in early B-cell activation through IL-4 production.  相似文献   

18.
During B-cell receptor (BCR) signaling, phosphoinositide-3 kinase (PI3K) is thought to function upstream of phospholipase Cgamma2 (PLCgamma2). PLCgamma2 deficiency specifically impedes transitional type 2 (T2) to follicular (FO) mature B-cell transition. Here, we demonstrate that PI3K deficiency specifically impaired T2-to-FO mature B-cell transition and marginal zone B-cell development. Furthermore, we investigated the functional relationship between PI3K and PLCgamma2 using PI3K-/-, PLCgamma2-/-, and PI3K-/- PLCgamma2-/- B cells. Interestingly, PLCgamma2 deficiency had no effect on BCR-mediated PI3K activation, whereas PI3K deficiency only partially blocked activation of PLCgamma2. Moreover, whereas PI3K-/- PLCgamma2-/- double deficiency did not affect hematopoiesis, it resulted in embryonic lethality. PI3K-/- PLCgamma2-/- fetal liver cells transplanted into B-cell null JAK3-/- mice failed to restore development of peripheral B cells and failed to progress through early B-cell development at the pro-B- to pre-B-cell transition, a more severe phenotype than was observed with either PI3K or PLCgamma2 single-deficiency B cells. Consistent with this finding, BCR signaling was more severely impaired in the absence of both PI3K and PLCgamma2 genes than in the absence of either one alone. Taken together, these results demonstrate that whereas PI3K functions upstream of PLCgamma2, activation of PLCgamma2 can occur independently of PI3K and that PI3K and PLCgamma2 also have distinct functions in BCR signal transduction.  相似文献   

19.
We examined the effects of in vivo exposure to a GSM-modulated 900 MHz RF field on B-cell peripheral differentiation and antibody production in mice. Our results show that exposure to a whole-body average specific absorption rate (SAR) of 2 W/kg, 2 h/day for 4 consecutive weeks does not affect the frequencies of differentiating transitional 1 (T1) and T2 B cells or those of mature follicular B and marginal zone B cells in the spleen. IgM and IgG serum levels are also not significantly different among exposed, sham-exposed and control mice. B cells from these mice, challenged in vitro with LPS, produce comparable amounts of IgM and IgG. Moreover, exposure of immunized mice to RF fields does not change the antigen-specific antibody serum level. Interestingly, not only the production of antigen-specific IgM but also that of IgG (which requires T-B-cell interaction) is not affected by RF-field exposure. This indicates that the exposure does not alter an ongoing in vivo antigen-specific immune response. In conclusion, our results do not indicate any effects of GSM-modulated RF radiation on the B-cell peripheral compartment and antibody production and thus provide no support for health-threatening effects.  相似文献   

20.
CBA and C57B1 mice (high and low responders to sheep red blood cells, respectively) were injected intravenously with syngeneic lymph node, marrow, spleen, or thymus cells together with sheep red blood cells (SRBC), and the production of antibody-forming cells (AFC) was assayed in the spleen. Transfer of lymph node, marrow, spleen, or thymus cells led to a significant enhancement of immune responsiveness in low-responding C57B1 mice. In contrast, transfer of marrow, lymph node, or spleen cells to high-responding CBA mice was accompanied by a decline in AFC production. These effects were magnified if syngeneic cell donors had been primed with SRBC; suppression in CBA mice and stimulation in C57B1 mice were especially pronounced after transfer of SRBC-primed lymphoid cells. Pretreatment of CBA donors with cyclophosphamide in a dose causing selective B-cell depletion completely abrogated the suppression of immune responsiveness. A large dose (107) of syngeneic B cells injected together with SRBC suppressed the accumulation of AFC in both CBA and C57B1 mice. No suppression of immune responsiveness was observed after transfer of intact thymus cells, hydrocortisone-resistant thymocytes, or activated T cells. We conclude that suppression of the immune response to SRBC is induced by B cells. At the same time, there is a possibility that the addition of “excess” B cells acts as a signal, triggering suppressor T cells.  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