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.
The GIVA phospholipase A(2) (PLA(2)) contains two domains: a calcium-binding domain (C2) and a catalytic domain. These domains are linked via a flexible tether. GIVA PLA(2) activity is Ca(2+)-dependent in that calcium binding promotes protein docking to the phospholipid membrane. In addition, the catalytic domain has a lid that covers the active site, presumably regulating GIVA PLA(2) activity. We now present studies that explore the dynamics and conformational changes of this enzyme in solution utilizing peptide amide hydrogen/deuterium (H/D) exchange coupled with liquid chromatography-mass spectrometry (DXMS) to probe the solvent accessibility and backbone flexibility of the C2 domain, the catalytic domain, and the intact GIVA PLA(2). We also analyzed the changes in H/D exchange of the intact GIVA PLA(2) upon Ca(2+) binding. The DXMS results showed a fast H/D-exchanging lid and a slow exchanging central core. The C2 domain showed two distinct regions: a fast exchanging region facing away from the catalytic domain and a slow exchanging region present in the "cleft" region between the C2 and catalytic domains. The slow exchanging region of the C2 domain is in tight proximity to the catalytic domain. The effects of Ca(2+) binding on GIVA PLA(2) are localized in the C2 domain and suggest that binding of two distinct Ca(2+) ions causes tightening up of the regions that surround the anion hole at the tip of the C2 domain. This conformational change may be the initial step in GIVA PLA(2) activation.  相似文献   

2.
The activity of the major isoform of porcine pancreatic phospholipase A2 (PLA2), designated B-PLA2, against micellar substrates is inhibited by heparin. Inhibition is a consequence of binding of the enzyme to heparin, documented by a heparin-induced alteration in the intrinsic fluorescence of B-PLA2 and in the 8-anilino-1-naphthalene sulfonate fluorescence and by the enhanced rate of chemical modification of the active site residue His-48. As a consequence of heparin binding, the conformation of B-PLA2 at the active site and at the amino-terminus is altered, and the enzyme does not bind to phospholipid micelles. In spite of the heparin-induced conformational changes, B-PLA2 retains its ability to catalyze the hydrolysis of monomeric phospholipid. Other glycosaminoglycans can bind to and inhibit the activity of B-PLA2 toward organized phospholipids, but none tested is as effective as heparin. An isoform of the pancreatic enzyme, designated UB-PLA2 and which corresponds to iso-pig PLA2, does not bind to nor is its catalytic activity influenced by heparin. A peptide corresponding to the amino-terminal 26 residues of B-PLA2 can rescue PLA2 from heparin inhibition. A similar peptide corresponding to the amino-terminus of UB-PLA2 has no effect on heparin inhibition. A model for the inhibition of B-PLA2 by heparin is proposed in which the catalytically significant effect of heparin is to interact directly with the amino-terminus of B-PLA2, the interfacial recognition site, to prevent the enzyme from binding to micellar substrates.  相似文献   

3.
Previously, we reported a growth-dependent change in prostaglandin production as a consequence of a marked growth-dependent alteration in arachidonic acid (AA) mobilization from phospholipids. Our present results show that fetal calf serum (FCS) and 4 beta-phorbol-12-myristate acetate (PMA) caused an enhancement of phospholipase A(2) (PLA(2)) activity in the membrane fraction of non-confluent cells allowing PLA(2) access to its substrate and the release of AA. Western blot analysis has shown that FCS and PMA increased secreted PLA(2) (sPLA(2)) expression in non-confluent 3T6 fibroblast cultures. Moreover, FCS and PMA induced dithiothreitol-sensitive and bromoenol lactone-sensitive PLA(2) activities in cytosol and membrane fraction. However, these stimuli did not modify significantly the PLA(2) activity in both fractions when 3T6 fibroblasts reached a high cell density. This could be associated with the impairment of AA mobilization in these cell culture conditions. On the other hand, we observed that FCS and PMA induced the same prostaglandin H synthase-2 induction in non-confluent and confluent culture conditions. Moreover, the prostaglandin E(2) levels reached in cell culture supernatants were independent of the degree of confluence when AA was added exogenously. These results suggest that the changes of intracellular distribution of PLA(2) activity of sPLA(2) and iPLA(2) stimulated by exogenous stimuli may be controlled by cell density conditions which constitute an important mechanism in the regulation of prostaglandin release.Copyright 2001 Wiley-Liss, Inc.  相似文献   

4.
The phospholipase A(2) (PLA(2)) enzymes are activated by binding to phospholipid membranes. Although the N-terminal alpha-helix of group I/II PLA(2)s plays an important role in the productive mode membrane binding of the enzymes, its role in the structural aspects of membrane-induced activation of PLA(2)s is not well understood. In order to elucidate membrane-induced conformational changes in the N-terminal helix and in the rest of the PLA(2), we have created semisynthetic human group IB PLA(2) in which the N-terminal decapeptide is joined with the (13)C-labeled fragment, as well as a chimeric protein containing the N-terminal decapeptide from human group IIA PLA(2) joined with a (13)C-labeled fragment of group IB PLA(2). Infrared spectral resolution of the unlabeled and (13)C-labeled segments suggests that the N-terminal helix of membrane-bound IB PLA(2) has a more rigid structure than the other helices. On the other hand, the overall structure of the chimeric PLA(2) is more rigid than that of the IB PLA(2), but the N-terminal helix is more flexible. A combination of homology modeling and polarized infrared spectroscopy provides the structure of membrane-bound chimeric PLA(2), which demonstrates remarkable similarity but also distinct differences compared with that of IB PLA(2). Correlation is delineated between structural and membrane binding properties of PLA(2)s and their N-terminal helices. Altogether, the data provide evidence that the N-terminal helix of group I/II PLA(2)s acts as a regulatory domain that mediates interfacial activation of these enzymes.  相似文献   

5.
Evidence is presented that the zymogen of porcine pancreatic phospholipase A2 (prophospholipase A2) interacts with a lipid-water interface provided that the interface has a net negative surface charge. Fluorescence spectroscopy and non-equilibrium gel filtration indicate that binding of prophospholipase A2 (proPLA) to mixed detergent micelles is dependent on the presence of an anionic detergent. Prophospholipase binding is accompanied by a change in the environment of the single tryptophan residue qualitatively similar to that observed when the active enzyme, phospholipase A2 (PLA), binds to micelles. In addition, the rate of tryptic activation of prophospholipase is significantly reduced in the presence of negatively-charged mixed micelles, whereas no change in rate occurs when neutral micelles are present. These observations suggest that the lack of catalytic activity of the zymogen toward organized substrates carrying a negative surface charge cannot be explained by a failure to bind at the lipid-water interface.  相似文献   

6.
Phospholipase A(2) (PLA(2)) binds to membranes and catalyzes phospholipid hydrolysis, thus initiating the biosynthesis of lipid-derived mediators of inflammation. A snake-venom PLA(2) was completely inhibited by covalent modification of the catalytic histidine 48 by p-bromophenacyl bromide. Moreover, His(48) modification affected PLA(2) structure, its membrane-binding affinity, and the effects of PLA(2) on the membrane structure. The native PLA(2) increased the order parameter of fluid membranes, whereas the opposite effect was observed for gel-state membranes. The data suggest membrane dehydration by PLA(2) and the formation of PLA(2)-membrane hydrogen bonding. The inhibited PLA(2) had lower membrane-binding affinity and exerted weaker effects on membrane hydration and on the lipid-order parameter. Although membrane binding resulted in formation of more flexible alpha-helices in the native PLA(2), which corresponds to faster amide hydrogen exchange, the modified enzyme was more resistant to hydrogen exchange and experienced little structural change upon membrane binding. The data suggest that 1), modification of a catalytic residue of PLA(2) induces conformational changes that propagate to the membrane-binding surface through an allosteric mechanism; 2), the native PLA(2) acquires more dynamic properties during interfacial activation via membrane binding; and 3), the global conformation of the inhibited PLA(2), including the alpha-helices, is less stable and is not influenced by membrane binding. These findings provide further evidence for an allosteric coupling between the membrane-binding (regulatory) site and the catalytic center of PLA(2), which contributes to the interfacial activation of the enzyme.  相似文献   

7.
Secretory low molecular weight phospholipase A(2)s (PLA(2)s) are believed to be involved in the release of arachidonic acid, a precursor for the biosynthesis of pro-inflammatory eicosanoids. Therefore, the specific inhibitors of these enzymes may act as potent anti-inflammatory agents. Similarly, the compounds with known anti-inflammatory properties should act as specific inhibitors. Two plant compounds, (a) anisic acid (4-methoxy benzoic acid) and (b) atropine (8-methyl-8-azabicyclo oct-3-hydroxy-2-phenylpropanoate), have been used in various inflammatory disorders. Both compounds (a) and (b) have been found to inhibit PLA(2) activity having binding constants of 4.5 x 10(-5) M and 2.1 x 10(-8) M, respectively. A group IIA PLA(2) was isolated and purified from the venom of Daboia russelli pulchella (DRP) and its complexes were made with anisic acid and atropine. The crystal structures of the two complexes (i) and (ii) of PLA(2) with compounds (a) and (b) have been determined at 1.3 and 1.2 A resolutions, respectively. The high-quality observed electron densities for the two compounds allowed the accurate determinations of their atomic positions. The structures revealed that these compounds bound to the enzyme at the substrate - binding cleft and their positions were stabilized by networks of hydrogen bonds and hydrophobic interactions. The most characteristic interactions involving Asp 49 and His 48 were clearly observed in both complexes, although the residues that formed hydrophobic interactions with these compounds were not identical because their positions did not exactly superimpose in the large substrate-binding hydrophobic channel. Owing to a relatively small size, the structure of anisic acid did not alter upon binding to PLA(2), while that of atropine changed significantly when compared with its native crystal structure. The conformation of the protein also did not show notable changes upon the bindings of these ligands. The mode of binding of anisic acid to the present group II PLA(2) is almost identical to its binding with bovine pancreatic PLA(2) of group I. On the other hand, the binding of atropine to PLA(2) is similar to that of another plant alkaloid aristolochic acid.  相似文献   

8.
The pathogenesis and pathophysiology of Acanthamoeba infections remain incompletely understood. Phospholipases are known to cleave phospholipids, suggesting their possible involvement in the host cell plasma membrane disruption leading to host cell penetration and lysis. The aims of the present study were to determine phospholipase activities in Acanthamoeba and to determine their roles in the pathogenesis of Acanthamoeba. Using an encephalitis isolate (T1 genotype), a keratitis isolate (T4 genotype), and an environmental isolate (T7 genotype), we demonstrated that Acanthamoeba exhibited phospholipase A(2) (PLA(2)) and phospholipase D (PLD) activities in a spectrophotometry-based assay. Interestingly, the encephalitis isolates of Acanthamoeba exhibited higher phospholipase activities as compared with the keratitis isolates, but the environmental isolates exhibited the highest phospholipase activities. Moreover, Acanthamoeba isolates exhibited higher PLD activities compared with the PLA(2). Acanthamoeba exhibited optimal phospholipase activities at 37℃ and at neutral pH indicating their physiological relevance. The functional role of phospholipases was determined by in vitro assays using human brain microvascular endothelial cells (HBMEC), which constitute the blood-brain barrier. We observed that a PLD-specific inhibitor, i.e., compound 48/80, partially inhibited Acanthamoeba encephalitis isolate cytotoxicity of the host cells, while PLA(2)-specific inhibitor, i.e., cytidine 5'-diphosphocholine, had no effect on parasite-mediated HBMEC cytotoxicity. Overall, the T7 exhibited higher phospholipase activities as compared to the T4. In contract, the T7 exhibited minimal binding to, or cytotoxicity of, HBMEC.  相似文献   

9.
Protobothrops flavoviridis venom contains plural phospholipase A(2) (PLA(2)) isozymes. A [Lys(49)]PLA(2) called BPII induced cell death in human leukemia cells. PLA2, an [Asp(49)]PLA(2) that has much stronger lipolytic activity than BPII, failed to induce cell death. BPII-treated cells showed morphological changes, DNA fragmentation, and nuclear condensation. This BPII-induced apoptotic cell death was neither inhibited by inhibitors of caspases 3 and 6 nor accompanied by activation of procaspase 3, indicating that BPII-induced cell death is caspase independent. Since inactive p-bromophenacylated BPII induced cell death, BPII-induced apoptotic cell death is independent of PLA(2) lipolytic activity. Rapid externalization of phosphatidylserine in BPII-treated cells was observed for fluorescein isothiocyanate (FITC)-labeled annexin V. In the cells treated with BPII, this spread over the cell membranes, implying that the cell toxicity of BPII is mediated via its cell-surface receptor.  相似文献   

10.
Fourier transform infrared spectroscopy has been used to investigate the secondary structure of porcine and bovine pancreatic phospholipase A2 (PLA2) and the zymogen of porcine PLA2, prophospholipase A2 (proPLA2), in both H2O and D2O media. Detailed qualitative analysis was made of these proteins using second derivative and deconvolution techniques. Quantitative studies of the proteins in solution made using Factor Analysis gave average values of 54% alpha-helix, 15% beta-sheet and 23% beta-turns. These values agree well with the secondary structures deduced from previous studies of single crystals using X-ray techniques. No significant differences in secondary structure were observed for porcine pancreatic (pro)phospholipase A2 in the presence or absence of Ca2+ ions, or in the temperature range 10-45 degrees C. The binding of the non-degradable phospholipid analogue, n-alkylphosphocholine, in monomeric form produced no significant difference in the secondary structure of either enzyme. Conformational differences were, however, observed between the enzyme lyophilised in a solid film and in aqueous solution. The change is probably due to the formation of beta-sheet upon hydration, coupled with a loss of random structures. Conformational differences in both porcine and bovine pancreatic PLA2 were observed on binding to n-alkylphosphocholine micelles. This change may be due to a small increase in the alpha-helical structure and a decrease in the beta-sheet, and/or possibly beta-turn content. Similar conformational changes were observed for the interaction of porcine and bovine PLA2 with the substrate analogue inhibitor 1-heptanoyl-2-heptanoylamino-2-deoxy-sn-glycero-3-phospho glycol in micellar form.  相似文献   

11.
Given the potent hydrolyzing activity toward phosphatidylcholine, group X secretory phospholipase A(2) (sPLA(2)-X) elicits a marked release of arachidonic acid linked to the potent production of lipid mediators in various cell types. We have recently shown that sPLA(2)-X can also act as a ligand for mouse phospholipase A(2) receptor (PLA(2)R). Here, we found that sPLA(2)-X was internalized and degraded via binding to PLA(2)R associated with the diminished prostaglandin E(2) (PGE(2)) formation in PLA(2)R-expressing Chinese hamster ovary (CHO) cells compared to CHO cells. Indirect immunocytochemical analysis revealed that internalized sPLA(2)-X was co-localized with PLA(2)R in the punctate structures in PLA(2)R-expressing CHO cells. Moreover, in mouse osteoblastic MC3T3-E(1) cells that endogenously express the PLA(2)R, the internalized sPLA(2)-X was localized in lysosomes. These findings demonstrate that PLA(2)R acts as a clearance receptor for sPLA(2)-X to suppress its strong enzymatic activity.  相似文献   

12.
An NMR study has been made of porcine pancreatic phospholipase A2 (PLA) in three environments: free in solution, in a binary complex with dodecylphosphocholine micelles, and in a ternary complex with a micelle and the substrate-like inhibitor (R)-1-octyl-2-(N-dodecanoylamino)-2-deoxyglycero-3-phosph oglycol. 1H and 15N chemical shifts, amide exchange rates, and NOE intensities are compared for the enzyme in different environments. From these data, structural differences are found for the N-terminal part, the end of the surface loop at residues Tyr69 and Thr70, and the active site residue His48, and also for the Ca-binding loop (residues 28-32). Specifically, when binding to a micelle, the side chains of residues Ala1, Trp3, and Tyr69, as well as all protons of Thr70, are found to be closer together. After subsequent introduction of the competitive inhibitor, further changes are found for these residues. The N-terminus is flexible in PLA free in solution, in contrast with the crystal structures where it adopts an alpha-helical conformation. According to the NMR data, this helix is rigidly formed only in the ternary complex. Furthermore, in the ternary complex, the N-terminal amino group and the exchangeable hydrogen at N3 of the ring of His48 are observed. We propose that PLA is activated in two steps. An initial conformational change occurs upon binding to a micellar interface. The catalytically active conformation of the enzyme, which has an extensive network of hydrogen bonds, is formed only when binding a substrate or competitive inhibitor at a lipid-water interface.  相似文献   

13.
The use of conformationally restricted phospholipids 1 and 2 has been employed to understand the conformational preference of phospholipase A2 (PLA2) for substrate phospholipids. Inhibition of porcine pancreatic PLA2 with 1 and 2 indicated a two- to fivefold preference for the distal isomer 2 over the proximal isomer 1. Based upon these studies, both side-chains of the substrate phospholipid appear to occupy the lipid binding domains near the active site with the side-chains further apart most preferred by PLA2.  相似文献   

14.
We have shown previously and confirmed in this study that the phospholipase A(2) (PLA(2)) activity of peroxiredoxin 6 (Prdx6) is markedly increased by phosphorylation. This report evaluates the conformation and thermodynamic stability of Prdx6 protein after phosphorylation to understand the physical basis for increased activity. Phosphorylation resulted in decreased negative far-UV CD, strengthened ANS binding, and a lack of rigid tertiary structure, compatible with a change in conformation to that of a molten globule. The ΔG°(D) was 3.3 ± 0.3 kcal mol(-1) for Prdx6 and 1.7 ± 0.7 kcal mol(-1) for pPrdx6, suggesting that phosphorylation destabilizes the protein. Phosphorylation of Prdx6 changed the conformation of the N-terminal domain exposing Trp 33, as determined by tryptophan fluorescence and NaI fluorescence quenching. The kinetics of interaction of proteins with unilamellar liposomes (50:25:15:10 DPPC:egg PC:cholesterol:PG molar ratio) were evaluated with tryptophan fluorescence. pPrdx6 bound to liposomes with a higher affinity (K(d) = 5.6 ± 1.2 μM) than Prdx6 (K(d) = 24.9 ± 4.5 μM). By isothermal titration calorimetry, pPrdx6 bound to liposomes with a large exothermic heat loss (ΔH = -31.49 ± 0.22 kcal mol(-1)). Correlating our conformational studies with the published crystal structure of oxidized Prdx6 suggests that phosphorylation results in exposure of hydrophobic residues, thereby providing accessibility to the sites for liposome binding. Because binding of the enzyme to the phospholipid substrate interface is a requirement for PLA(2) activity, these results indicate that a change in the conformation of Prdx6 upon its phosphorylation is the basis for enhancement of PLA(2) enzymatic activity.  相似文献   

15.
It has been reported that the cooperative binding of calcium ions indicated a local conformational change of the human cytosolic phospholipase A2 (cPLA2) C2 domain (Nalefski et al., (1997) Biochemistry 36, 12011-12018). However its structural evidence is less known (Malmberg et al., (2003) Biochemistry 42, 13227-13240). In this letter, life-time decay and fluorescence quenching techniques were employed to compare the calcium-induced conformational changes. The life-time decay parameters and fluorescence quenching constant changes were small between the apo- and holo-C2 domains when tryptophan residue was excited at 295 nm. In contrast, the quenching constant change was large, from 0.52 M(-1) for the apo-C2 to 8.8 M(-1) for the holo-C2 domain, when tyrosine residues were excited at 284 nm. Our results provide new information on amino acid side chain orientation change at calcium binding loop 3, which is necessary for Ca2+ binding regulated membrane targeting of human cytosolic phospholipase A2.  相似文献   

16.
Pigment epithelium-derived factor (PEDF) is an extracellular multifunctional protein belonging to the serpin superfamily with demonstrable neurotrophic, gliastatic, neuronotrophic, antiangiogenic, and antitumorigenic properties. We have previously provided biochemical evidence for high affinity PEDF-binding sites and proteins in plasma membranes of retina, retinoblastoma, and CNS cells. This study was designed to reveal a receptor involved in the biological activities of PEDF. Using a yeast two-hybrid screening, we identified a novel gene from pigment epithelium of the human retina that codes for a PEDF-binding partner, which we term PEDF-R. The derived polypeptide has putative transmembrane, intracellular and extracellular regions, and a phospholipase domain. Recently, PEDF-R (TTS-2.2/independent phospholipase A(2) (PLA(2))zeta and mouse desnutrin/ATGL) has been described in adipose cells as a member of the new calcium-independent PLA(2)/nutrin/patatin-like phospholipase domain-containing 2 (PNPLA2) family that possesses triglyceride lipase and acylglycerol transacylase activities. Here we describe the PEDF-R gene expression in the retina and its heterologous expression by bacterial and eukaryotic systems, and we demonstrate that its protein product has specific and high binding affinity for PEDF, has a potent phospholipase A(2) activity that liberates fatty acids, and is associated with eukaryotic cell membranes. Most importantly, PEDF binding stimulates the enzymatic phospholipase A(2) activity of PEDF-R. In conclusion, we have identified a novel PEDF-R gene in the retina for a phospholipase-linked membrane protein with high affinity for PEDF, suggesting a molecular pathway by which ligand/receptor interaction on the cell surface could generate a cellular signal.  相似文献   

17.
The binding of heparin to basic fibroblast growth factor (bFGF) induces a small but highly reproducible conformational change observable in the amide I region of the protein's infrared spectrum. The observed spectral changes suggest that the conformational change is highly localized most likely in the beta-turn regions of the bFGF molecule. Heparan sulfate, a component of the endothelial extracellular matrix, was also observed to bind to bFGF and induce a similar conformational change to that observed for heparin. Further, sucrose octasulfate, a compound which mimics the effects of heparin biologically, was also observed to induce this same conformational change. This spectroscopically observable change has allowed us to probe the functional determinants necessary for heparin to bind the bFGF and to induce the observed conformational change. We have determined the effects of binding of various monomeric and polymeric, sulfated and nonsulfated glycosaminoglycans and carbohydrate compounds. The results indicate that the binding of heparin involves highly specific interactions. Further, heparin was observed to greatly increase the thermal stability of bFGF, raising the Tm by 25 degrees C. Sucrose octasulfate was also able to enhance the thermal stability of bFGF, but not to the same extent as heparin.  相似文献   

18.
VEGF (vascular endothelial growth factor) regulates neovascularization through binding to its receptor KDR (kinase insert domain-containing receptor; VEGF receptor-2). We recently identified a catalytically inactive PLA(2) (phospholipase A(2)) homologue (KDR-bp) in the venom of eastern cottonmouth (Agkistrodon piscivorus piscivorus) as a third KDR-binding protein, in addition to VEGF(165) and tissue inhibitor of metalloproteinase-3. KDR-bp binds to the extracellular domain of KDR with a K(d) of 10(-8) M, resulting in specific blockade of endothelial cell growth induced by VEGF(165). Inactive PLA(2) homologues are widely distributed in the venoms of Viperidae snakes and are known to act as myotoxins. In the present study, we demonstrated that KDR-binding ability is a common characteristic for inactive PLA(2) homologues in snake venom, but not for active PLA(2)s such as neurotoxic and platelet aggregation-modulating PLA(2)s. To understand better the KDR and KDR-bp interaction, we resolved the binding region of KDR-bp using eight synthetic peptides designed based on the structure of KDR-bp. A synthetic peptide based on the structure of the C-terminal loop region of KDR-bp showed high affinity for KDR, but other peptides did not, suggesting that the C-terminal loop region of KDR-bp is involved in the interaction with KDR. The results of the present study provide insight into the binding of inactive PLA(2) homologues to KDR, and may also assist in the design of novel anti-KDR molecules for anti-angiogenic therapy.  相似文献   

19.
We reported previously that human fetal skin fibroblast migration into a denuded area was stimulated by an autocrine factor, basic fibroblast growth factor (bFGF). Since the signal transduction pathway of this migration is unknown, we attempted to clarify it by comparing this fibroblast migration with a previously reported bovine endothelial cell migration into a wounded area stimulated by an addition of bFGF, in which the bFGF signal was mediated by phospholipase A(2)-coupled G-protein and phospholipase A(2) (PLA(2)) via arachidonic acid. Our study demonstrated that pertussis toxin, a specific inhibitor of PLA(2)-coupled G-protein, did not suppress human fetal skin fibroblast migration, but 2-(p-amylcinnamyl)amino-4-chlorobensoic acid (ONO-RS-082), a PLA(2) inhibitor, did. Since ONO-RS-082 is a non-specific PLA(2) inhibitor, a cytoplasmic, Ca-dependent PLA(2) (cPLA(2)) inhibitor, AACOCF3, was examined. AACOCF3 suppressed cell migration in certain concentrations. The PLA(2) inhibitor-suppressed cell migration was restored by adding arachidonic acid, and cell migration suppressed by anti-bFGF antibodies was restored by adding arachidonic acid. In addition, pertussis toxin did not suppress arachidonic acid release, which shows an action of PLA(2), but AACOCF3 did. These results indicate that human fetal skin fibroblast migration stimulated by an autocrine factor, bFGF, was mediated by PLA(2) via arachidonic acid without the involvement of PLA(2)-coupled G-protein.  相似文献   

20.
A key feature of prion diseases is the conversion of the cellular prion protein (PrP(C)) into disease-related isoforms (PrP(Sc)), the deposition of which is thought to lead to neurodegeneration. In this study a pharmacological approach was used to determine the metabolic pathways involved in the formation of protease-resistant PrP (PrP(res)) in three prion-infected cell lines (ScN2a, SMB, and ScGT1 cells). Daily treatment of these cells with phospholipase A(2) (PLA(2)) inhibitors for 7 days prevented the accumulation of PrP(res). Glucocorticoids with anti-PLA(2) activity also prevented the formation of PrP(res) and reduced the infectivity of SMB cells. Treatment with platelet-activating factor (PAF) antagonists also reduced the PrP(res) content of cells, while the addition of PAF reversed the inhibitory effect of PLA(2) inhibitors on PrP(res) formation. ScGT1 cells treated with PLA(2) inhibitors or PAF antagonists for 7 days remained clear of detectable (PrPres) when grown in control medium for a further 12 weeks. Treatment of non-infected cells with PLA(2) inhibitors or PAF antagonists reduced PrP(C) levels suggesting that limiting cellular PrP(C) may restrict prion formation in infected cells. These data indicate a pivotal role for PLA(2) and PAF in controlling PrP(res) formation and identify them as potential therapeutic agents.  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