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.
During apoptosis, DNA fragmentation and intracellular acidification occur concurrently. Previous results have shown that intracellular acidification is not required for DNA fragmentation, while the alternative, that acidification is a consequence of DNA fragmentation was analyzed here. To obviate the requirement of any nuclear function in acidification, apoptosis was induced by staurosporine in cytoplasts made from the breast tumor cell line MDA-MB-468. Both cells and cytoplasts demonstrated externalization of phosphatidylserine that was prevented by the pan-caspase inhibitor zVAD-fluoromethylketone or by expression of Bcl-2. Intracellular acidification was observed in both cells and cytoplasts and this was also inhibited by both zVAD-fluoromethylketone and Bcl-2. These results show that intracellular acidification and DNA fragmentation are independent consequences of caspase action during apoptosis.  相似文献   

2.
We have shown elsewhere that acidification is an early event in apoptosis, preceding DNA cleavage. Cells expressing the most common mutation (delF508) of the cystic fibrosis transmembrane regulator (CFTR) exhibit a higher resting intracellular pH and are unable to secrete chloride and bicarbonate in response to cAMP. We hypothesized that defective acidification in cells expressing delF508 CFTR would interfere with the acidification that accompanies apoptosis, which in turn, would prevent endonuclease activation and cleavage of DNA. We therefore determined whether the function of the CFTR would affect the process of apoptosis in mouse mammary epithelial C127 cells stably transfected with the wild-type CFTR (C127/wt) or the delF508 mutation of the CFTR (C127/508). C127 cells possessed an acid endonuclease capable of DNA degradation at low pH. Sixteen hours after treatment with cycloheximide, C127/wt cells underwent cytoplasmic acidification. In contrast, C127/508 cells failed to demonstrate acidification. Furthermore, the C127/508 cells did not show nuclear condensation or DNA fragmentation detected by in situ nick-end labeling after treatment with cycloheximide or etoposide, in contrast to the characteristic features of apoptosis demonstrated by the C127/wt cells. Measurement of cell viability indicated a preservation of cell viability in C127/508 cells but not in C127/wt cells. That this resistance to the induction of apoptosis depended upon the loss of CFTR activity is shown by the finding that inhibition of the CFTR with diphenylamine carboxylate in C127/wt cells conferred similar protection. These findings suggest a role for the CFTR in acidification during the initiation of apoptosis in epithelial cells and imply that a failure to undergo programmed cell death could contribute to the pathogenesis of cystic fibrosis.  相似文献   

3.
Cytoplasmic acidification is now recognized as a feature of apoptosis in a variety of systems. However, its relation to other events in the process of apoptosis is not yet characterized. In this work, we examined the effect of BCL-2 overexpression on acidification mediated by cycloheximide treatment or Fas ligation in Jurkat T-lymphoblasts. We find that BCL-2 overexpression attenuates cytoplasmic acidification and apoptosis detected by annexin V labeling. Acidification and phosphatidylserine externalization were found to occur concurrently. We also examined the requirement for protease activation for cytoplasmic acidification to occur and found that inhibition of interleukin-1beta converting enzyme/CED-3 family proteases (using carbobenzoxy-Val-Ala-Asp-fluoromethylketone, an inhibitor of these proteases) prevents acidification and apoptosis mediated by Fas ligation. These studies suggest that BCL-2 acts at a point upstream of acidification and that protease activation is also upstream of acidification.  相似文献   

4.
Apoptosis associated oligonucleosomal fragmentation of DNA can result from the activation of endonucleases that exhibit different pH optima and are either sensitive or insensitive to divalent cations. DNA fragmentation due to activation of cation sensitive endonucleases occurs in the absence of a change in intracellular pH whereas intracellular acidification is a feature of apoptosis characterized by activation of cation insensitive acidic endonuclease. We have reported earlier that somatostatin (SST) induced DNA fragmentation and apoptosis is signaled in a receptor subtype selective manner uniquely via human somatostatin receptor subtype 3 (hSSTR3). In the present study we investigated the pH dependence and cation sensitivity of endonuclease induced in hSSTR3 expressing CHO-K1 cells by the SST agonist octreotide (OCT) and its effect on intracellular pH. We show that OCT induced apoptosis is associated with selective stimulation of a divalent cation insensitive acidic endonuclease. The intracellular pH of of cells undergoing OCT induced apoptosis was 0.9 pH units lower than that of control cells. The effect of OCT on endonuclease and pH was inhibited by orthovanadate as well as by pretreatment with pertussis toxin, suggesting that hSSTR3 initiated cytotoxic signaling is protein tyrosine phosphatase mediated and is G protein dependent. These findings suggest that intracellular acidification and activation of acidic endonuclease mediate wild type p53 associated apoptosis signaled by hormones acting via G protein coupled receptors.  相似文献   

5.
Hydrogen peroxide (H2O2) is considered to be a mediator of apoptotic cell death but the mechanism by which it induces apoptosis is unclear. Here, we show that cells undergoing apoptosis from exposure to H2O2 display a significant decrease in intracellular concentration of superoxide (O2-) which is associated with a reduction of the intracellular milieu, as measured by an increase in the GSH/GSSG ratio and a decrease in intracellular pH. The notion that a decrease in intracellular O2- concentration triggers apoptosis is supported by the observation that H2O2-mediated apoptosis could be retarded in cells in which the intracellular O2- concentration is maintained at or above the cellular baseline level by inhibition of the major O2- scavenger superoxide dismutase (Cu/Zn SOD). Taken together, our observations indicate that a decrease in the intracellular O2- concentration, reduction and acidification of the intracellular milieu constitute a signal for H2O2-mediated apoptosis, thereby inducing a reductive as opposed to an oxidative stress.  相似文献   

6.
The induction of apoptosis by topoisomerase I inhibitors, camptothecin and SN38, was evaluated in drug-sensitive HL60 and multidrug-resistant (MDR) HL60-Vinc leukemic cells. MDR cells displayed a partial resistance to these apoptotic stimuli and this phenomenon was not modulated by verapamil. Basal free calcium concentrations were similar in both cell sublines and were not modified during treatment. Cytoplasmic pH was more acidic in sensitive cells than in MDR cells. Moreover, a significant acidification was obtained during the early stage of apoptosis in sensitive HL60 cells only. Basal Bcl-2 protein expression was found to be greater in MDR than in sensitive cells and was not modulated by apoptosis inducers. This increase of Bcl-2 in MDR cells could be due to the selection process as vincristine enhances Bcl-2 phosphorylation and expression in HL60 sensitive cells. MDR HL60-Vincristine cells therefore display a resistance to apoptosis induced by non-MDR drugs, possibly by Bcl-2 overexpression and inability of these drugs to mediate intracellular pH changes in these drug-resistant cells.  相似文献   

7.
Several laboratories have reported that overexpression of the multidrug resistance (MDR) protein is associated with intracellular alkalinization, and several investigators have reported that cells induced to undergo programmed cell death (apoptosis) acidify quite significantly. Because it is difficult to fully explain the resistance to apoptosis-inducing chemotherapeutic drugs that is exhibited by MDR tumor cells solely via altered drug transport alone [Hoffman et al. (1996) J. Gen. Physiol. 108, 295-313], we have investigated whether overexpression of the hu MDR 1 protein alters progression of the apoptotic cascade. LR73 fibroblasts induced to undergo apoptosis either via treatment with the chemotherapeutic drug colchicine or by serum withdrawal exhibit cellular volume changes, intracellular acidification, nuclear condensation, and chromosomal digestion ("ladder formation"), characteristic of apoptosis, in a temporally well-defined pattern. However, multidrug resistant LR73/20E or LR73/27 hu MDR 1 transfectants recently created in our laboratory without selection on chemotherapeutic drug are significantly delayed in the onset of apoptosis as defined by the above criteria, regardless of whether apoptosis is induced by colchicine treatment or by serum withdrawal. Thus, the delay cannot simply be due to the well-known ability of MDR protein overexpression to lower chemotherapeutic drug accumulation in MDR cells. LR73/27V500 "selectants", exhibiting similar levels of MDR protein overexpression but higher multidrug resistance due to selection with the chemotherapeutic drug vincristine, exhibit a slightly longer delay in the progression of apoptosis. Normal apoptotic cascade kinetics are partially restored by pre-treatment of the MDR cells with the MDR protein inhibitor verapamil. Untransfected LR73 cells not expressing MDR protein but elevated in pHi via manipulation of CO2/HCO3- as described [Hoffman et al. (1996) J. Gen. Physiol. 108, 295-313] are inhibited in DNA ladder formation, similar to LR73/hu MDR 1 transfectants. These results uncover an additional mechanism whereby MDR protein overexpression may promote the survival of tumor cells and further support the notion that in some systems intracellular acidification may be either causal or permissive for proper progression of the apoptotic cascade.  相似文献   

8.
It is not well appreciated that nutritional status can modulate apoptosis, a process that eliminates unwanted or damaged cells. Choline is an essential nutrient, and its absence induces apoptosis. When PC12 cells were cultivated in a choline-free medium, apoptosis was induced (27.4% of cells apoptotic at 72 h as compared to 4.4% in control medium). In choline-free medium at 72 h, there was a 49% decrease in phosphatidylcholine concentration (P<0.01) and a 34% decrease in sphingomyelin concentration (P<0.01); however, there was no change in phosphatidylethanolamine concentration. Before detecting increased apoptosis in choline-deficient cells, we measured a significant increase in ceramide (218% control) and diacyglycerol (155% control) concentrations. The addition of a cell-permeable ceramide to cells in control medium induced apoptosis; however, adding a cell-permeable diacyglycerol did not induce apoptosis. Caspase is a common mediator of apoptosis, and choline deficiency-induced apoptosis was prevented completely by replacing choline or adding a caspase inhibitor into the medium within 48 h of initial choline deprivation. In those cells rescued by replacing choline at 36 h, the concentrations of phosphatidylcholine, sphingomyelin, ceramide, and diacyglycerol returned to levels of control cells. In those cells rescued by adding a caspase inhibitor at 36 h, the concentrations of sphingomyelin and ceramide returned to control levels, but the concentrations of phosphatidylcholine and diacyglycerol did not return to normal. We propose that availability of dietary factors (choline in this model) can modulate apoptosis. Mechanisms that we identify using this model may help us to explain why dietary choline influences brain development.  相似文献   

9.
Acridine orange (AO) has been used as a vital fluorescent stain to identify apoptotic cells in Drosophila, but little is known about what structures are stained. We explored the specificity of AO staining while studying nuclear apoptosis in Tetrahymena. Using AO alone or together with the vital nuclear stain Hoechst 33342 (HO), we find that lysosomes are generally clustered around the degenerating nucleus and that such nuclei are stained an orange-red color, like lysosomes. Significantly, the combined dyes, more so than with AO alone, distinguish between apoptotic and normal (or necrotic) nuclei by a clear color difference. Moreover, these dyes differentially stain apoptotic and normal nuclei in avian chondrocytes. The differential staining results are nullified in fixed cells or in cytoskeletal preparations treated with RNAse. Similarly, lysosomotrophic agents eliminate the differential staining. Our results are consistent with acidification of the apoptotic nucleus, possibly by fusion with lysosomes. However, even under basic conditions, the macronucleus condenses and is eliminated, suggesting that, if the nucleus is becoming acidified, acidification by itself is not essential for nuclear elimination. The differential staining procedure may provide a useful method for specifically identifying apoptotic cells and separating them for further analysis.  相似文献   

10.
During embryo implantation, apoptosis is observed morphologically at the implantation site of endometrium. The objectives of this study were to demonstrate biochemical evidence of apoptosis and quantitative assessment of DNA fragmentation in uterine epithelial cells using a mouse implantation model, and to investigate the autocrine/paracrine regulation of apoptosis in uterine epithelial cells during blastocyst outgrowth. Blastocysts from day 4 pregnant mice were cultured on uterine epithelial cells for 96 h. Uterine epithelial cells dislodged by trophoblasts in endometrium-trophoblast unit demonstrated morphological features of apoptosis by Acridine Orange staining. Electrophoresis demonstrated DNA ladder and DNA fragmentation by enzyme-linked immunosorbent assay markedly increased after 48 h period of incubation. Apoptosis increased in an exponential way in accordance with trophoblast outgrowth. In addition, DNA fragmentation was shown in the epithelial cells by adding embryo-conditioned medium (CM) and the effect of embryo CM on apoptosis was significantly inhibited by anti-transforming growth factor (TGF)-beta antibody. Delayed outgrowth was observed after 48 h of incubation in the blastocysts cultured with anti-TGF-beta antibody. These results suggest there is autocrine/paracrine regulation of apoptosis in uterine epithelial cells at mouse embryo implantation and that TGF-beta might play an important role in the occurrence of apoptosis in the endometrium-trophoblast unit.  相似文献   

11.
We found that exposure to room temperature (RT/21 degrees C) causes apoptosis in HL-60 cells. Here we characterized RT-induced apoptosis in HL-60. After exposure to RT, apoptosis starts within 6 h and more than 80% of the cells underwent apoptosis within 20 h. All cells, however, were committed to apoptosis after 16 h and no viable cells could be recovered. The caspase-1 inhibitor (YVAD-CHO) effectively blocked apoptosis, whereas the caspase-3 inhibitor (DEVD-CHO) did not. About 20% of newly obtained early passage HL-60 cells (passage 10) also underwent apoptosis by RT treatment. These data suggest that some population in HL-60 which responds to RT with apoptosis became dominant during passaging.  相似文献   

12.
Several papers have recorded that Epstein-Barr virus (EBV) infection can protect B-cells from apoptosis. In the present paper, we record an increase in apoptosis in an EBV-infected human Burkitt's lymphoma cell line (AW-Ramos), compared with its virus free counterpart (Ramos). The viral-infected cells died more rapidly by apoptosis during normal growth in culture: at 72h, Ramos cells had 8% apoptosis and AW-Ramos had 27% apoptosis. The EBV-infected cells were particularly sensitive to treatment with phorbol-12, 13-dibutyrate (PDBu, 0.05 microgram/ml): at 72h, Ramos cells had 12% apoptosis and AW-Ramos had 42% apoptosis. DNA gel electrophoresis supported the morphological findings. Our results serve as a caveat against generalizations that may be made about prevention of apoptosis by EBV infection.  相似文献   

13.
The mechanisms which drive initiated cells to progress to form carcinomas are poorly understood. CWSV-1 rat hepatocytes, in which p53 protein is inactivated by SV40 large T antigen, respond by inducing p53-independent apoptosis when acutely switched to medium containing low choline (16% apoptotic at 48 h in 5 microM choline) as compared with controls (1% apoptotic at 48 h in 70 microM choline). The rate of apoptosis was inversely correlated with cellular phosphatidylcholine content. Choline deficiency (CD)-induced apoptosis is probably mediated by TGFbeta1 and reactive oxygen species, since immunoneutralization of TGFbeta1 in the medium or treatment with N-acetylcysteine (an antioxidant) or addition of neocuproine (a transition metal chelator) prevented CD-induced apoptosis. CWSV-1 hepatocytes could be gradually adapted to survive in 5 microM choline. CD-adapted cells had increased membrane phosphatidylcholine concentrations (compared with acute CD cells). Adapted cells acquired relative resistance to CD-induced apoptosis (7% of adapted cells compared with 19% of non-adapted cells were apoptotic at 48 h in 5 microM choline). They also became relatively resistant to another p53-independent form of apoptosis (TGFbeta1-induced). CD-adapted hepatocytes developed increased capability for anchorage-independent growth and formed tumors when transplanted into nude mice; passage-matched control hepatocytes did not possess these properties. Cell transformation was dependent on exposure to the selective pressure of CD apoptosis, as we observed that when CD apoptosis was inhibited with an antioxidant during adaptation, cells did not become anchorage independent. Acquisition by p53-deficient cells of resistance to p53-independent inducers of apoptosis (CD, TGFbeta1 and reactive oxygen species) may leave cells without another important apoptotic defensive barrier and may be responsible for the progression of initiated cells to frank carcinomas.  相似文献   

14.
Many parallel processes occur during the final stages of apoptosis. It is not clear which of these processes occur in all or most models of apoptosis and which occur only in some. In addition, the temporal relationship of these events is not always well understood. Correlated flow cytometric measurements were used to address these questions. Several models of apoptosis were studied, including thymocytes treated with dexamethasone. MOLT-4 cells treated with etoposide, U937 cells treated with anti-Fas, HL-60 cells treated with camptothecin, Raji cells grown in low serum, and aged neutrophils. All models showed a decrease in LDS-751 and fluorescein diacetate (FDA) staining, an increase in staining with dihydrorhodamine 123 (dhR123) or dihydroethidium, and an acidification of the cytoplasm. In each model, these changes were highly correlated, appearing simultaneously as multiparameter measurements. Changes in membrane status detected with merocyanin 540 (MC540) and annexin V behaved differently. A population with LDS-751 and FDA changes but without annexin V or MC540 changes could be demonstrated in some models. Several models did not show any change in annexin V binding, and HL-60 did not show a change in MC540 binding during apoptosis. The loss of cell surface antigens (CD45 and CD16) from aged neutrophils occurred in the entire LDS-751 and FDA dim population, even though other membrane changes (including the appearance of annexin V binding sites) were only apparent in a subset of these cells. These results suggest a model for the ordering of some of the terminal processes in apoptosis, with annexin V and MC540 changes trailing other events in apoptosis. These results confirm the need for caution in using a single-parameter measurement as an indicator of apoptosis for any new model being studied.  相似文献   

15.
We demonstrated previously that the activation of v-Abl protein tyrosine kinase (PTK) in IC.DP murine pre-mast cells resulted in suppression of apoptosis after withdrawal of interleukin 3 (IL-3), that protein kinase C (PKC) translocated to the nucleus 6 h after v-Abl PTK activation and that inhibition of PKC restored apoptosis after IL-3 deprivation in the presence of v-Abl PTK activity. Here we demonstrate that v-Abl PTK activation is followed by an approximately twofold increase in mRNA level of Bcl-XL by 6 h and a corresponding increase in Bcl-XL protein level by 24 h. Bcl-xL RNA and protein decreased in IL-3 deprived cells in the absence of v-Abl PTK activity. Exposure of cells with v-Abl PTK active to the PKC inhbitor calphostin C (125 ng/ml) prevented the increase in Bcl-xL protein and resulted in apoptosis. No changes in Bax or Bcl-2 protein level were noted after IL-3 withdrawal and/or activation of v-Abl PTK. Bak was barely detectable and Bad protein level decreased in cells undergoing apoptosis. The data suggest that suppression of apoptosis by v-Abl PTK in the absence of IL-3 is associated with PKC signalling and the upregulation of Bcl-xL in IC.DP cells.  相似文献   

16.
目的 选用人类单核细胞白血病细胞株THP-1,体外加入紫草素二甲醚衍生物5,8-二甲基-2-β-羟基异戊酰紫草素(SK36),研究其在THP-1细胞株增殖和凋亡中的作用,并对其作用机制进行初步探讨.方法 CCK法检测SK36对THP-1细胞的增殖抑制作用;流式细胞术检测细胞早期凋亡标记Annexin V/PI及细胞凋亡通路Caspase活性;激光共聚焦显微镜观察细胞凋亡和坏死.结果 流式细胞术检测结果显示,1.02μg/ml SK36作用24、48 h后的细胞凋亡率分别为(40.61±2.13)%和(67.40 ±9.15)%,明显高于对照组的(16.97±0.61)%,差异有统计学意义(t=18.444,t=9.528,P<0.01);SK36诱导THP-1细胞凋亡且经历了Caspase-3的激活(F=323.61,P<0.01).结论 紫草素二甲醚衍生物SK36能有效地诱导THP-1细胞凋亡,其作用机制主要是激活Caspase-3.  相似文献   

17.
18.
Flavopiridol (L86-8275) is a synthetic flavone currently undergoing Phase I clinical trials. It is active against a series of human cancer cell lines and has been shown to inhibit a broad range of protein kinases, including cyclin-dependent kinases and protein kinase C (PKC). Previous studies have shown that the PKC-specific inhibitor safingol significantly enhances the induction of apoptosis by mitomycin-C (MMC) in gastric cancer cells. Because flavopiridol can potentially inhibit PKC, we elected to determine the extent to which flavopiridol would promote MMC-induced apoptosis in both gastric and breast cancer cells. For these studies, MKN-74 gastric cancer cells and MDA-MB-468 breast cancer cells were exposed to either no drug, 1 microgram/ml MMC alone, 300 nM flavopiridol alone, or a combination of chemotherapy with flavopiridol for 24 h. Sequence specificity was also examined by first exposing cells to MMC for 24 h followed by flavopiridol for 24 h or to the same drugs in the reverse order. Apoptosis was measured by quantitative fluorescence microscopy of nuclear chromatin condensation in cells stained with the dye, bisbenzimide trihydrochloride. Exposure of MKN-74 cells to flavopiridol alone induced apoptosis in 12 +/- 1% of the cells, and exposure to MMC alone induced apoptosis in 10 +/- 1%. However, the combination of flavopiridol and MMC increased the induction of apoptosis to 55 +/- 3% of the cells (P < 0.005 for the drug combination versus flavopiridol alone). Pretreatment with the PKC activator 3-phorbol 12-myristate 13-acetate only partially reversed this effect (43 +/- 1%; P < 0.025). In MDA-MB-468 cells, flavopiridol alone induced apoptosis in 17 +/- 1% of the cells, and MMC alone induced apoptosis in 10 +/- 1% of the cells. The combination of flavopiridol and MMC increased the percentage of MDA-MB-468 cells undergoing apoptosis to 58 +/- 4% (P < 0.005 for the drug combination versus flavopiridol alone). Sequential treatment with MMC followed by flavopiridol induced apoptosis in 63 +/- 2% of the MKN-74 cells (P < 0.05 versus the concomitant drug combination) and in 76 +/- 2% of the MDA-MB-468 cells (P < 0.025 versus the concomitant drug combination), whereas flavopiridol followed by MMC did not increase the induction of apoptosis in either cell line. As determined by the terminal deoxynucleotidyl transferase labeling of the 3' ends of DNA fragments produced in apoptotic cells, the induction of apoptosis with the combination of flavopiridol and MMC occurred to MKN-74 cells in all phases of the cell cycle (i.e., G0-G1, S, and G2-M). These results indicate that flavopiridol potentiates the cytotoxic effect of the chemotherapeutic agent MMC by promoting drug-induced apoptosis in tumor cells. Sequencing studies suggest that MMC followed by flavopiridol or simultaneous treatment is superior to flavopiridol followed by MMC. The enhancement of MMC-induced apoptosis by flavopiridol may be partially PKC dependent and is not associated with one specific region of the cell cycle.  相似文献   

19.
Testis torsion-induced aspermatogenesis is not necessarily due to permanent loss of blood flow nor to dysfunctional Leydig cells or Sertoli cells. This investigation was undertaken to gain further insight into the mechanism underlying torsion-induced germ cell loss. Male rats were subjected to 1-h or 2-h ischemia-inducing torsion, and testes were examined at either 1, 2, 4, 24, or 48 h after torsion, depending on the study. Testes were examined for evidence of 1) in situ apoptosis by the terminal deoxynucleotidyl transferase-mediated deoxyuridine triphosphate (dUTP)-biotin nick-end labeling (TUNEL) technique, 2) apoptosis by the DNA "laddering" technique, 3) leukocyte margination and diapedesis in testicular vessels by immunocytochemical and histological techniques, and 4) testicular lipid peroxidation by the thiobarbituric acid reactive substances assay. The first TUNEL evidence for torsion-induced apoptosis was at 4 h after repair of 1-h torsion. Induction of apoptosis was confirmed by the electrophoretic laddering of DNA fragments. It was hypothesized that apoptosis was induced by reactive oxygen species arising from reperfusing leukocytes. A significant increase in both leukocyte margination and diapedesis occurred 4 h after torsion repair as did a significant increase in intratesticular lipid peroxidation products. These events were contemporaneous with the first appearance of apoptosis and consistent with the hypothesis that post-torsion, germ cell-specific apoptosis is induced by reactive oxygen species.  相似文献   

20.
T-84 and Caco-2 human colon carcinoma cells and Madin-Darby canine kidney (MDCK) cells were used to study binding and transcytosis of iodinated Clostridium botulinum neurotoxin serotypes A, B, and C, as well as tetanus toxin. Specific binding and transcytosis were demonstrated for serotypes A and B in intestinal cells. Using serotype A as an example, the rate of transcytosis by T-84 cells was determined in both apical to basolateral (11.34 fmol/h/cm2) as well as basolateral to apical (8.98 fmol/h/cm2) directions, and by Caco-2 cells in the apical to basolateral (8.42 fmol/h/cm2) direction. Serotype A retained intact di-chain structure during transit through T-84 or Caco-2 cells, and when released on the basolateral side was toxic in vivo to mice and in vitro on mouse phrenic nerve-hemidiaphragm preparations. Serotype C and tetanus toxin did not bind effectively to T-84 cells, nor were they efficiently transcytosed (8-10% of serotype A). MDCK cells did not bind or efficiently transcytose (0.32 fmol/h/cm2) botulinum toxin. Further characterization demonstrated that the rate of transcytosis for serotype A in T-84 cells was increased 66% when vesicle sorting was disrupted by 5 microM brefeldin A, decreased 42% when microtubules were disrupted by 10 microM nocodazole, and decreased 74% at 18 degreesC. Drugs that antagonize toxin action at the nerve terminal, such as bafilomycin A1 (which prevents acidification of endosomes) and methylamine HCl (which neutralizes acidification of endosomes), produced only a modest inhibitory effect on the rate of transcytosis (17-22%). These results may provide an explanation for the mechanism by which botulinum toxin escapes the human gastrointestinal tract, and they may also explain why specific serotypes cause human disease and others do not.  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