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.
Bacterial LPS induces endothelial cell (EC) injury both in vivo and in vitro. We studied the effect of Escherichia coli 0111:B4 LPS on movement of 14C-BSA across bovine pulmonary artery EC monolayers. In the presence of serum, a 6-h LPS exposure augmented (P < 0.001) transendothelial 14C-BSA flux compared with the media control at concentrations > or = 0.5 ng/ml, and LPS (10 ng/ml) exposures of > or = 2-h increased (P < 0.005) the flux. In the absence of serum, LPS concentrations of up to 10 micrograms/ml failed to increase 14C-BSA flux at 6 h. The addition of 10% serum increased EC sensitivity to the LPS stimulus by > 10,000-fold. LPS (10 ng/ml, 6 h) failed to increase 14C-BSA flux at serum concentrations < 0.5%, and maximum LPS-induced increments could be generated in the presence of > or = 2.5%. LPS-binding protein (LBP) and soluble CD14 (sCD14) could each satisfy this serum requirement; either anti-LBP or anti-CD14 antibody each totally blocked (P < 0.00005) the LPS-induced changes in endothelial barrier function. LPS-LBP had a more rapid onset than did LPS-sCD14. The LPS effect in the presence of both LBP and sCD14 exceeded the effect in the presence of either protein alone. These data suggest that LBP and sCD14 each independently functions as an accessory molecule for LPS presentation to the non-CD14-bearing endothelial surface. However, in the presence of serum both molecules are required.  相似文献   

2.
Lipoproteins isolated from normal human plasma can bind and neutralize bacterial lipopolysaccharide (LPS) and may represent an important mechanism in host defense against gram-negative septic shock. Recent studies have shown that experimentally elevating the levels of circulating high-density lipoproteins (HDL) provides protection against death in animal models of endotoxic shock. We sought to define the components of HDL that are required for neutralization of LPS. To accomplish this we have studied the functional neutralization of LPS by native and reconstituted HDL using a rapid assay that measures the CD14- dependent activation of leukocyte integrins on human neutrophils. We report here that reconstituted HDL particles (R-HDL), prepared from purified apolipoprotein A-I (apoA-I) combined with phospholipid and free cholesterol, are not sufficient to neutralize the biologic activity of LPS. However, addition of recombinant LPS binding protein (LBP), a protein known to transfer LPS to CD14 and enhance responses of cells to LPS, enabled prompt binding and neutralization of LPS by R- HDL. Thus, LBP appears capable of transferring LPS not only to CD14 but also to lipoprotein particles. In contrast with R-HDL, apoA-I containing lipoproteins (LpA-I) isolated from plasma by selected affinity immunosorption (SAIS) on an anti-apoA-I column, neutralized LPS without addition of exogenous LBP. Several lines of evidence demonstrated that LBP is a constituent of LpA-I in plasma. Passage of plasma over an anti-apoA-I column removed more than 99% of the LBP detectable by ELISA, whereas 31% of the LBP was recovered by elution of the column. Similarly, the ability of plasma to enable activation of neutrophils by LPS (LBP/Septin activity) was depleted and recovered by the same process. Furthermore, an immobilized anti-LBP monoclonal antibody coprecipitated apoA-I. The results described here suggest that in addition to its ability to transfer LPS to CD14, LBP may also transfer LPS to lipoproteins. Since LBP appears to be physically associated with lipoproteins in plasma, it is positioned to play an important role in the neutralization of LPS.  相似文献   

3.
Lipopolysaccharide binding protein (LBP) is a plasma protein known to facilitate the diffusion of bacterial LPS (endotoxin). LBP catalyzes movement of LPS monomers from LPS aggregates to HDL particles, to phospholipid bilayers, and to a binding site on a second plasma protein, soluble CD14 (sCD14). sCD14 can hasten transfer by receiving an LPS monomer from an LPS aggregate, and then surrendering it to an HDL particle, thus acting as a soluble "shuttle" for an insoluble lipid. Here we show that LBP and sCD14 shuttle not only LPS, but also phospholipids. Phosphatidylinositol (PI), phosphatidylcholine, and a fluorescently labeled derivative of phosphatidylethanolamine (R-PE) are each transferred by LBP from membranes to HDL particles. The transfer could be observed using recombinant LBP and sCD14 or whole human plasma, and the plasma-mediated transfer of PI could be blocked by anti-LBP and partially inhibited by anti-CD14. sCD14 appears to act as a soluble shuttle for phospholipids since direct binding of PI and R-PE to sCD14 was observed and because addition of sCD14 accelerated transfer of these lipids. These studies define a new function for LBP and sCD14 and describe a novel mechanism for the transfer of phospholipids in blood. In further studies, we show evidence suggesting that LBP transfers LPS and phospholipids by reciprocal exchange: LBP-catalyzed binding of R-PE to LPS x sCD14 complexes was accompanied by the exit of LPS from sCD14, and LBP-catalyzed binding of R-PE to sCD14 was accelerated by prior binding of LPS to sCD14. Binding of one lipid is thus functionally coupled with the release of a second. These results suggest that LBP acts as a lipid exchange protein.  相似文献   

4.
Bacterial endotoxin (lipopolysaccharide [LPS]) causes fatal shock in humans and experimental animals. The shock is mediated by cytokines released by direct LPS stimulation of cells of monocytic origin (monocyte/macrophage [MO]). Recent studies have supported the concept that the plasma protein, LPS binding protein (LBP), plays an important role in controlling MO responses to LPS. Specifically, evidence has been presented to suggest that CD14, a membrane protein present in MO, serves as a receptor for complexes of LPS and the plasma protein LPS binding protein (LBP). In this function CD14 mediates attachment of LPS-bearing particles opsonized with LBP and appears to play an important role in regulating cytokine production induced by complexes of LPS and LBP. The CD14-, murine pre-B cell line 70Z/3 responds to LPS by synthesis of kappa light chains and consequent expression of surface IgM. To better understand the role of CD14 in controlling cellular responses to LPS, we investigated the effect of transfection of CD14 into 70Z/3 cells on LPS responsiveness. We report here that transfection of human or rabbit CD14 cDNA into 70Z/3 cells results in membrane expression of a glycosyl-phosphatidylinositol-anchored CD14. When LPS is complexed with LBP, CD14-bearing 70Z/3 cells bind more LPS than do the parental or 70Z/3 cells transfected with vector only. Remarkably, the expression of CD14 lowers the amount of LPS required to stimulate surface IgM expression by up to 10,000-fold when LPS dose-response curves in the CD14-, parental and CD14-bearing, transfected 70Z/3 cells are compared. In contrast, the response of CD14-bearing 70Z/3 cells and the parental 70Z/3 cell line (CD14-) to interferon gamma is indistinguishable. LPS stimulation of the parental and CD14-bearing 70Z/3 cells results in activation of NF-kB. These data provide evidence to support the concept that the LPS receptor in cells that constitutively express CD14 may be a multiprotein complex containing CD14 and membrane protein(s) common to a diverse group of LPS-responsive cells.  相似文献   

5.
Lipopolysaccharides (LPS) that lack acyloxyacyl groups can antagonize responses to LPS in human cells. Although the site and mechanism of inhibition are not known, it has been proposed that these inhibitory molecules compete with LPS for a common cellular target such as a cell-surface binding receptor. In the present study, we used an in vitro model system to test this hypothesis and to evaluate the role of CD14 in cellular responses to LPS. Cells of the THP-1 human monocyte-macrophage cell line were exposed to 1,25 dihydroxyvitamin D3 to induce adherence to plastic and expression of CD14, a binding receptor for LPS complexed with LPS-binding protein (LBP). The uptake of picograms of [3H]LPS (agonist) and enzymatically deacylated LPS [3H]dLPS (antagonist) was measured by exposing the cells to the radiolabeled ligands for short incubation periods. The amounts of cell-associated LPS and dLPS were then correlated with cellular responses by measuring the induction of nuclear NF-kappa B binding activity and the production of cell-associated interleukin (IL)-1 beta. We found that similar amounts of [3H]LPS or [3H]dLPS were taken up by the cells. The rate of cellular accumulation of the ligands was greatly enhanced by LBP and blocked by a monoclonal antibody to CD14 (mAb 60b), yet no cellular responses were induced by dLPS or dLPS-LBP complexes. In contrast, LPS stimulated marked increases of NF-kappa B binding activity and IL-1 beta. These responses were enhanced by LBP and inhibited by mAb 60b. dLPS and its synthetic lipid A counterpart, LA-14-PP (also known as lipid Ia, lipid IVa, or compound 406) strongly inhibited LPS-induced NF-kappa B and IL-1 beta, yet neither antagonist inhibited the uptake of LPS via CD14. dLPS did not inhibit NF-kappa B responses to tumor necrosis factor (TNF) alpha or phorbol ester. Our results indicate that (a) both stimulatory and nonstimulatory ligands can bind to CD14 in the presence of LBP; (b) the mechanism of inhibition by dLPS is LPS-specific, yet does not involve blockade of LPS binding to CD14; and (c) in keeping with previous results of others, large concentrations of LPS can stimulate the cells in the absence of detectable binding to CD14. The findings indicate that the site of dLPS inhibition is distal to CD14 binding in the LPS signal pathway in THP-1 cells, and suggest that molecules other than CD14 are important in LPS signaling.  相似文献   

6.
Exposure of rabbit peritoneal exudate macrophages (PEM) or whole blood to picomolar concentrations of LPS induces adaptation or hyporesponsiveness to LPS. Because of the importance of plasma LPS-binding protein (LBP) and the macrophage cell membrane protein CD14 in recognition of LPS, we examined the effect of LBP on LPS-induced adaptation in PEM. PEM exposed to LPS in the presence of LBP for 8 h were markedly less responsive to subsequent stimulation by LPS than monocytes/macrophages (M phi) adapted in the absence of LBP. LPS-induced expression of TNF was sharply reduced in LBP-LPS-adapted PEM, but in contrast these cells remained fully responsive to Staphylococcus aureus peptidoglycan. We considered that specific hyporesponsiveness in LPS-adapted M phi or in blood monocytes could be due to decreased expression of CD14 or diminished binding of LBP-LPS complexes to CD14. However, flow cytometry analysis revealed only minimal reduction of CD14 expression or CD14-dependent binding of a fluorescent LPS derivative when normo- and hyporesponsive cells were compared. These results show that complexes of LPS and LBP are more effective than LPS alone in inducing adaptation to LPS, and LPS-induced hyporesponsiveness probably results from changes in cellular elements distinct from CD14 that are involved in either LPS recognition or LPS-specific signal transduction.  相似文献   

7.
Lipopolysaccharide binding protein (LBP) is an acute-phase reactant that binds bacterial LPS. We show that LBP binds to the surface of live Salmonella and to LPS coated erythrocytes (ELPS), and strongly enhances the attachment of these particles to macrophages. LBP bridges LPS-coated particles to macrophages (MO) by first binding to the LPS, then binding to MO. Pretreatment of ELPS with LBP enabled binding to MO, but pretreatment of MO had no effect. Moreover, MO did not recognize erythrocytes coated with LBP unless LPS was also added, thus suggesting that interaction of LBP with LPS results in a conformational change in LBP that allows recognition by MO. Binding of LBP-coated particles appears to be mediated by a receptor found on blood monocytes and MO but not on other leukocytes or umbilical vein endothelium. The receptor is mobile in the plane of the membrane since binding activity on MO was downmodulated upon spreading of cells on surfaces coated with LBP-LPS complexes. The receptor appears to be distinct from other opsonic receptors since downmodulation of CR1, CR3, Fc gamma RI, Fc gamma RII, and Fc gamma RIII with mAbs did not affect binding of LBP-coated particles, and leukocytes from CD18-deficient patients bound LBP-coated particles normally. Coating of erythrocytes with LBP-LPS complexes strongly enhanced phagocytosis observed in the presence of suboptimal amounts of anti-erythrocyte IgG. However, binding mediated by LBP-LPS complexes alone caused neither phagocytosis of the LBP-coated erythrocytes nor initiation of an oxidative burst. The results of our studies define LBP as an opsonin. During the acute phase, LBP can be expected to bind gram-negative bacteria and bacterial fragments and promote the interaction of coated bacteria with phagocytes.  相似文献   

8.
Gram-negative bacterial lipopolysaccharide (LPS) stimulates phagocytic leukocytes by interacting with the cell surface protein CD14. Cellular responses to LPS are markedly potentiated by the LPS-binding protein (LBP), a lipid-transfer protein that binds LPS aggregates and transfers LPS monomers to CD14. LBP also transfers LPS to lipoproteins, thereby promoting the neutralization of LPS. LBP present in normal plasma has been shown to enhance the LPS responsiveness of cells in vitro. The role of LBP in promoting LPS responsiveness in vivo was tested in LBP-deficient mice produced by gene targeting in embryonic stem cells. Whole blood from LBP-deficient animals was 1,000-fold less responsive to LPS as assessed by the release of tumor necrosis factor (TNF)-α. Blood from gene-targeted mice was devoid of immunoreactive LBP, essentially incapable of transferring LPS to CD14 in vitro, and failed to support cellular responses to LPS. These activities were restored by the addition of exogenous recombinant murine LBP to the plasma. Despite these striking in vitro findings, no significant differences in TNF-α levels were observed in plasma from wild-type and LBP-deficient mice injected with LPS. These data suggest the presence of an LBP-independent mechanism for responding to LPS. These LBP knockout mice may provide a tool for discovering the nature of the presumed second mechanism for transferring LPS to responsive cells.  相似文献   

9.
We have recently shown that monomeric bacterial LPS is rapidly delivered from the plasma membrane to an intracellular site and that agents that block vesicular transport block responses of neutrophils to lipopolysaccharide (LPS) (Detmers, P.A., N. Thiéblemont, T. Vasselon, R. Pironkova, D.S. Miller, and S.D. Wright. 1996. J. Immunol. 157:5589–5596). To examine further the connection between intracellular transport of LPS and signaling, we observed internalization of fluorescently labeled LPS in cells from LPS-hyporesponsive (Lpsd) mice. Binding of fluorescent LPS from LPS–soluble CD14 (sCD14) complexes by peritoneal macrophages from Lpsd and control (Lpsn) mice was quantitatively similar, and confocal images obtained from these cells exhibited an identical appearance immediately after labeling. Incubation of labeled Lpsn macrophages at 37°C caused movement of the fluorescence from the cell perimeter in one or two spots in the perinuclear region. However, in Lpsd cells the fluorescence remained dispersed, suggesting a defect in vesicular transport. LPS resembles ceramide, and Lpsd mice fail to respond to ceramide. As with LPS, we found that binding of fluorescent ceramide by Lpsd and Lpsn macrophages was quantitatively similar, and the label moved rapidly to one to two spots in the perinuclear region in Lpsn mice. However, in Lpsd macrophages the fluorescence remained dispersed. These results show that cells deficient in responses to LPS exhibit defective vesicular transport of LPS and ceramide and point to a role for vesicular transport in responses to these mediators.  相似文献   

10.
T Cell Stimulation In Vivo by Lipopolysaccharide (LPS)   总被引:14,自引:0,他引:14       下载免费PDF全文
Lipopolysaccharide (LPS) from gram-negative bacteria causes polyclonal activation of B cells and stimulation of macrophages and other APC. We show here that, under in vivo conditions, LPS also induces strong stimulation of T cells. As manifested by CD69 upregulation, LPS injection stimulates both CD4 and CD8+ T cells, and, at high doses, stimulates naive (CD44lo) cells as well as memory (CD44hi) cells. However, in terms of cell division, the response of T cells after LPS injection is limited to the CD44hi subset of CD8+ cells. In contrast with B cells, proliferative responses of CD44hi CD8+ cells require only very low doses of LPS (10 ng). Based on studies with LPS-nonresponder and gene-knockout mice, LPS-induced proliferation of CD44hi CD8+ cells appears to operate via an indirect pathway involving LPS stimulation of APC and release of type I (α, β) interferon (IFN-I). Similar selective stimulation of CD44hi CD8+ cells occurs in viral infections and after injection of IFN-I, implying a common mechanism. Hence, intermittent exposure to pathogens (gram-negative bacteria and viruses) could contribute to the high background proliferation of memory–phenotype CD8+ cells found in normal animals.  相似文献   

11.
CD14, a myeloid cell-surface receptor and soluble plasma protein, binds LPS and other microbial molecules and initiates the innate immune response to bacterial invasion. The blood concentration of soluble CD14 (sCD14) increases during the systemic response to infection. Although high sCD14 blood levels have correlated with increased risk of dying from severe sepsis, sCD14 can diminish cell responses to LPS. We show here that in human serum, sCD14 increases the rate at which cell-bound LPS is released from the monocyte surface and binds to plasma lipoproteins. This enhanced rate of LPS efflux is associated with a significant reduction in the ability of monocytes to produce cytokines in response to LPS. Serum from septic patients reduced the LPS-monocyte interaction by as much as tenfold, and depletion of sCD14 from the serum restored LPS-monocyte binding and release kinetics to near normal levels. In serum from septic patients, monocyte-bound LPS also moved more rapidly into lipoproteins, which completely neutralized the biologic activity of the LPS that bound to them. In human plasma, sCD14 thus diminishes monocyte responses to LPS by transferring cell-bound LPS to lipoproteins. Stress-related increases in plasma sCD14 levels may help prevent inflammatory responses within the blood.  相似文献   

12.
A series of 700 blood donor sera were screened for IgG antibodies to the core of Gram-negative bacterial endotoxin with a quantitative enzyme-linked immunosorbent-assay (ELISA), based on a cocktail of incomplete-core R-LPS from four different Gram-negative bacterial species, and further serum samples were obtained from donors exhibiting a range of different reactivity for isolation of serum IgG. Analysis of the different IgG samples by ELISA employing a panel of individual LPS from 31 different Gram-negative bacteria covering a range of species, serotypes and R-LPS chemotyes showed that high-titer sera from the screening ELISA expressed IgG with multiple reactivity to LPS in the complex ELISA. We investigated this multiple reactivity in three serum IgGs by inhibition and absorption of isolated serum IgG ELISA reactivity to R-LPS, employing purified LPS and whole bacteria respectively. In two cases the ELISA reactivity appeared to be predominantly attributable to a single antibody component directed to the inner LPS core structure in the lipid A to KDO region. For the third serum IgG, the results suggested that the cross-reactivity may be attributable to more than one specificity-group of cross-reactive antibodies, although still restricted to the LPS inner core structures.  相似文献   

13.
Lipopolysaccharide binding protein in acute pancreatitis   总被引:14,自引:0,他引:14  
OBJECTIVE: To assess the expression of plasma lipopolysaccharide binding protein (LBP) concentrations and its relationship to markers of the systemic inflammatory response syndrome during acute pancreatitis. DESIGN: A prospective study. SETTING: General surgical units of university teaching hospitals in the Belfast area. PATIENTS: The study included 18 patients admitted with established diagnosis of acute pancreatitis on the basis of elevated serum amylase or by contrast radiology. Patients were retrospectively stratified using the Modified Glasgow Criteria into severe (n = 7) and mild (n = 11) disease. INTERVENTIONS AND MEASUREMENTS: Blood samples were obtained at admission (day 1) and for a further 3 days for the measurement of LBP, C-reactive protein (CRP), tumor necrosis factor, and interleukin (IL)-6. Acute Physiology and Chronic Health Evaluation (APACHE) II scores were calculated on day 1 and day 2. MAIN RESULTS: LBP and CRP concentrations were significantly increased from healthy control values in acute pancreatitis patients at presentation. In the mild group LBP, CRP and IL-6 concentrations remained relatively constant throughout the study period. By comparison, severe acute pancreatitis was associated with significantly higher LBP concentrations and a marked systemic inflammatory response as evidenced by increased CRP, IL-6, and APACHE II scores. The rise in LBP occurred after the observed increase of these markers. Significant correlations were found among CRP and LBP, IL-6 and LBP, and IL-6 and APACHE II scores. There were no fatalities in the mild group, whereas four of the seven patients with severe disease died. CONCLUSIONS: LBP was significantly raised in patients with severe acute pancreatitis but would seem to be of limited use in predicting disease severity. This acute phase protein may have a role in the progression of systemic complications associated with acute pancreatitis.  相似文献   

14.
We have recently shown that lipopolysaccharide (LPS)-binding protein (LBP) is a lipid transfer protein that catalyzes two distinct reactions: movement of bacterial LPS (endotoxin) from LPS micelles to soluble CD14 (sCD14) and movement of LPS from micelles to reconstituted high density lipoprotein (R-HDL) particles. Here we show that LBP facilitates a third lipid transfer reaction: movement of LPS from LPS- sCD14 complexes to R-HDL particles. This action of LBP is catalytic, with one molecule of LBP enabling the movement of multiple LPS molecules into R-HDL. LBP-catalyzed movement of LPS from LPS-sCD14 complexes to R-HDL neutralizes the capacity of LPS to stimulate polymorphonuclear leukocytes. Our findings show that LPS may be transferred to R-HDL either by the direct action of LBP or by a two- step reaction in which LPS is first transferred to sCD14 and subsequently to R-HDL. We have observed that the two-step pathway of LPS transfer to R-HDL is strongly favored over direct transfer. Neutralization of LPS by LBP and R-HDL was accelerated more than 30- fold by addition of sCD14. Several observations suggest that sCD14 accelerates this reaction by serving as a shuttle for LPS: addition of LBP and sCD14 to LPS micelles resulted in LPS-sCD14 complexes that could diffuse through a 100-kD cutoff filter; LPS-sCD14 complexes appeared transiently during movement of LPS to R-HDL facilitated by purified LBP; and sCD14 could facilitate transfer of LPS to R-HDL without becoming part of the final LPS-R-HDL complex. Complexes of LPS and sCD14 were formed transiently when LPS was incubated in plasma, suggesting that these complexes may play a role as intermediates in the neutralization of LPS under physiological conditions. These findings detail a new activity for sCD14 and suggest a novel mechanism for lipid transfer by LBP.  相似文献   

15.
We investigated whether pretreatment with geranylgeranylacetone (GGA), a potent heat shock protein (HSP) inducer, could inhibit proinflammatory cytokine liberation and nitric oxide (NO) production in lipopolysaccharide (LPS)-treated murine macrophages. The levels of NO and tumor necrosis factor-alpha (TNF-alpha) released from murine macrophage RAW 264 cells were increased dose- and time-dependently following treatment with LPS (1 microg/ml). GGA (80 microM) treatment 2 h before LPS addition significantly suppressed TNF-alpha and NO productions at 12 h and 24 h after LPS, respectively, indicating that GGA inhibits activation of macrophages. However, replacement by fresh culture medium before LPS treatment abolished the inhibitory effect of GGA on NO production in LPS-treated cells. Furthermore, GGA inhibited both HSP70 and inducible NO synthase expressions induced by LPS treatment despite an HSP inducer. When it was examined whether GGA interacts with LPS and/or affects expression of Toll-like receptor 4 (TLR4) and CD14 on the cell surface, GGA inhibited the binding of LPS to the cell surface, while GGA did not affect TLR4 and CD14 expressions. These results indicate that GGA suppresses the binding of LPS to the cell surface of macrophages, resulting in inhibiting signal transduction downstream of TLR4.  相似文献   

16.
LPS binding protein (LBP) is an acute-phase glycoprotein that facilitates LPS activation of immune cells through interactions with CD14 and Toll-like receptor 4. Initially, LBP production was thought to occur exclusively in the liver in response to stimulation with TNF-alpha, IL-1, and IL-6. More recently, it has been shown that type II pneumocytes are also capable of LBP production. Little is known, however, regarding the regulation and or distribution of this protein in response to localized intrapulmonary infection. We performed time-course experiments challenging C3H mice intratracheally with LPS (10 mug). In separate experiments, mice deficient in IL-6 were given the same dose of intratracheal LPS and euthanized 8 h later. Despite the intratracheal route of LPS administration, an increase in plasma LBP concentrations occurred earlier and was of greater magnitude than the increase observed in bronchoalveolar lavage fluid. Liver LBP mRNA increased to a greater extent than did lung LBP mRNA. Whereas the TNF-alpha response remained localized within the alveolar space, IL-6 was increased both locally and in plasma. Of several tissues analyzed, the lung was the greatest producer of IL-6 mRNA. Plasma LBP was significantly decreased in the IL-6-deficient mice compared with wild-type controls challenged with intratracheal LPS. We conclude that lung-derived IL-6 is an important mediator of hepatic LBP up-regulation. We speculate that the disruption of these lung-liver signaling pathways may be important to host response efforts to confine infection to the lung. If impaired, this may be one mechanism underlying the increased mortality observed in patients with liver disease who develop pneumonia.  相似文献   

17.
18.
It has been suggested that the antibiotic-induced release of lipopolysaccharide (LPS) is an important cause of the development of septic shock in patients treated for severe infections caused by gram-negative bacteria. β-Lactam antibiotics change the integrity of the bacterial cell envelope by binding to penicillin-binding proteins (PBP) in the membrane and thus may affect the amount of LPS that is released and the kinetics of that release. In this respect, ceftazidime at intermediate concentrations binds with a high affinity to PBP 3 and PBP 1a and thus can induce filament formation in addition to killing, whereas imipenem preferentially binds to PBP 2 and PBP 1b, leading to spheroplast formation and rapid cell lysis. We investigated the effects of these antibiotics on the killing and the release of the radioactively labelled LPS of Salmonella typhi Ty 21A. A mathematical model was developed to calculate the delay between bacterial killing and LPS release, designated the lag time. At antibiotic concentrations inducing equal killing, the amount of LPS released was the same for both antibiotics. Only after 6 h of incubation at antibiotic concentrations above 0.5 μg/ml, the amount of 3H-LPS released was slightly higher (~1.2-fold) in incubations with ceftazidime than in those with imipenem, and the maximum releases of the total label were 33.2% ± 0.89% and 27.1% ± 0.45%, respectively. Despite the clear concentration-dependent effect on the bacterial killing and subsequent LPS release, the lag time was independent of the antibiotic concentration. For ceftazidime as well as imipenem the lag time amounted to approximately 60 min. In conclusion, our findings imply that the mechanism of antibiotic-induced LPS release is independent of the PBP affinities for these β-lactam antibiotics. Furthermore, once the organism is killed by either imipenem or ceftazidime, the rate of LPS release from S. typhi does not differ according to the antibiotic with which the organism is killed, and there is little difference in the relative amount of LPS released.  相似文献   

19.
The neurotrophic property of the immunosuppressant drug FK506 (tacrolimus) is believed to depend on the 12-kDa FK506-binding protein (FKBP-12). Here, we show that FK506 maintains its neurotrophic activity in primary hippocampal cell cultures from FKBP-12 knockout mice. In human neuroblastoma SH-SY5Y cells, the neurotrophic action of FK506 (10 pM to 10 nM) is completely prevented by the addition of a monoclonal antibody (50-100 nM) to the immunophilin FKBP-52 (also known as FKBP-59 or heat shock protein 56), a component of mature steroid receptor complexes. By itself, the FKBP-52 antibody is also neurotrophic. The neurotrophic activity of dexamethasone (50 nM) is potentiated by FK506, whereas that of beta-estradiol (50 nM) is not altered, suggesting a common mechanisms of action. Geldanamycin (which disrupts mature steroid receptor complexes) is also neurotrophic (0.1-10 nM), whereas it reduces the neurotrophic activity of FK506 and steroid hormones (dexamethasone and beta-estradiol). Conversely, 20 mM molybdate (which prevents the disruption of mature steroid receptor complexes) decreases the neurotrophic activity of FK506, FKBP-52 antibody, dexamethasone, and beta-estradiol. In rats, FK506 (10 mg/kg s.c.) augments the regenerative response of regenerating motor and sensory neurons to nerve injury as shown by its ability to increase the axotomy-induced induction of c-jun expression. A model is proposed to account for the neurotrophic action of both neuroimmunophilin ligands (FK506) and steroid hormones. Components of steroid receptor complexes represent novel targets for the rational design of new neurotrophic drugs.  相似文献   

20.
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