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.
Cytosolic phospholipase A(2)(cPLA(2)), an enzyme responsible for the generation of arachidonic acid, is located in the cytosolic compartment in most tissues and it translocates to membrane compartments when activated. We found that cPLA(2) distribution in pancreatic beta-cells is different from that of most other mammalian cells: it is evenly distributed throughout the beta-cell, in both cytoplasmic and nuclear compartments. Agents that increased intracellular Ca(2+) in the MIN6 beta-cell line also stimulated a redistribution of cPLA(2) immunoreactivity such that the majority of the enzyme moved from the nucleus to the cytoplasm. The time course of events was compatible with the elevation in Ca(2+) being responsible for translocation of cPLA(2). These observations suggest that cPLA(2) may be compartmentalised in unstimulated beta-cells, perhaps to limit its access to substrate prior to elevations in intracellular Ca(2+).  相似文献   

2.
Although CpG containing DNA is an important regulator of innate immune responses via toll-like receptor 9 (TLR9), excessive activation of this receptor is detrimental to the host. Here, we show that cytosolic phospholipase A2 (cPLA2) activation is important for TLR9-mediated inducible nitric oxide synthase (iNOS) expression. Activation of TLR9 signaling by CpG induces iNOS expression and NO production. Inhibition of TLR9 blocked the iNOS expression and NO production. The CpG also stimulates cPLA2-hydrolyzed arachidonic acid (AA) release. Inhibition of cPLA2 activity by inhibitor attenuated the iNOS expression by CpG response. Additionally, knockdown of cPLA2 protein by miRNA also suppressed the CpG-induced iNOS expression. Furthermore, the CpG rapidly phosphorylates three MAPKs and Akt. A potent inhibitor for p38 MAPK or Akt blocked the CpG-induced AA release and iNOS expression. These results suggest that TLR9 activation stimulates cPLA2 activity via p38 or Akt pathways and mediates iNOS expression.  相似文献   

3.
Cytosolic phospholipase A2 (cPLA2) mediates agonist-induced arachidonic acid release, the first step in eicosanoid production. cPLA2 is regulated by phosphorylation and by calcium, which binds to a C2 domain and induces its translocation to membrane. The functional roles of phosphorylation sites and the C2 domain of cPLA2 were investigated. In Sf9 insect cells expressing cPLA2, okadaic acid, and the calcium-mobilizing agonists A23187 and CryIC toxin induce arachidonic acid release and translocation of green fluorescent protein (GFP)-cPLA2 to the nuclear envelope. cPLA2 is phosphorylated on multiple sites in Sf9 cells; however, only S505 phosphorylation partially contributes to cPLA2 activation. Although okadaic acid does not increase calcium, mutating the calcium-binding residues D43 and D93 prevents arachidonic acid release and translocation of cPLA2, demonstrating the requirement for a functional C2 domain. However, the D93N mutant is fully functional with A23187, whereas the D43N mutant is nearly inactive. The C2 domain of cPLA2 linked to GFP translocates to the nuclear envelope with calcium-mobilizing agonists but not with okadaic acid. Consequently, the C2 domain is necessary and sufficient for translocation of cPLA2 to the nuclear envelope when calcium is increased; however, it is required but not sufficient with okadaic acid.  相似文献   

4.
Aires V  Hichami A  Boulay G  Khan NA 《Biochimie》2007,89(8):926-937
We synthesized a diacylglycerol (DAG)-containing arachidonic acid, i.e., 1-stearoyl-2-arachidonyl-sn-glycerol (SAG), and studied its implication in the modulation of canonical transient receptor potential sub-type 6 (TRPC6) channels in stably-transfected HEK-293 cells. SAG induced the influx of Ca(2+), and also of other bivalent cations like Ba(2+) and Sr(2+), in these cells. SAG-evoked Ca(2+) influx was not due to its metabolites as inhibitors of DAG-lipase (RHC80267) and DAG-kinase (R50922) failed to inhibit the response of the same. To emphasise that SAG exerts its action via its DAG configuration, but not due to the presence of stearic acid at sn-1 position, we synthesized 1-palmitoyl-2-arachidonyl-sn-glycerol (PAG). PAG-induced increases in [Ca(2+)](i) were not significantly different from those induced by SAG. For the comparative studies, we also synthesized the DAG-containing docosahexaenoic acid, i.e., 1-stearoyl-2-docosahexaenoyl-sn-glycerol (SDG). We observed that SDG and 1,2-dioctanoyl-sn-glycerol (DOG), a DAG analogue, also evoked increases in [Ca(2+)](i), which were lesser than those evoked by SAG. However, activation of TRPC6 channels by all the DAG molecular species (SAG, DOG and SDG) required Src kinases as the tyrosine kinase inhibitors, PP2 and SU6656, significantly attenuated the increases in [Ca(2+)](i) evoked by these agents. Moreover, disruption of lipid rafts with methyl-beta-cyclodextrin completely abolished SAG-, DOG- and SDG-induced increases in [Ca(2+)](i). The present study shows that SAG as well as SDG and DOG stimulate Ca(2+) influx through the activation of TRPC6 calcium channels which are regulated by Src kinases and intact lipid raft domains.  相似文献   

5.
Here, we investigated the modulation of cytosolic phospholipase A2 (cPLA2)-mediated arachidonic acid (AA) release by the polyprenylated acylphloroglucinol hyperforin. Hyperforin increased AA release from human platelets up to 2.6 fold (maximal effect at 10 µM) versus unstimulated cells, which was blocked by cPLA2α-inhibition, and induced translocation of cPLA2 to a membrane compartment. Interestingly, these stimulatory effects of hyperforin were even more pronounced after depletion of intracellular Ca2+ by EDTA plus BAPTA/AM. Hyperforin induced phosphorylation of cPLA2 at Ser505 and activated p38 mitogen-activated protein kinase (MAPK), and inhibition of p38 MAPK by SB203580 prevented cPLA2 phosphorylation. However, neither AA release nor translocation of cPLA2 was abrogated by SB203580. In cell-free assays using liposomes prepared from different lipids, hyperforin failed to stimulate phospholipid hydrolysis by isolated cPLA2 in the presence of Ca2+. However, when Ca2+ was omitted, hyperforin caused a prominent increase in cPLA2 activity using liposomes composed of 1-palmitoyl-2-arachidonyl-sn-glycero-3-phosphoethanolamine but not of 1-palmitoyl-2-arachidonyl-sn-glycero-3-phosphocholine (PAPC) unless the PAPC liposomes were enriched in cholesterol (20 to 50%). Finally, two-dimensional 1H-MAS-NMR analysis visualized the directed insertion of hyperforin into POPC liposomes. Together, hyperforin, through insertion into phospholipids, may facilitate cPLA2 activation by enabling its access towards select lipid membranes independent of Ca2+ ions. Such Ca2+- and phosphorylation-independent mechanism of cPLA2 activation may apply also to other membrane-interfering molecules.  相似文献   

6.
Exposure of human peripheral blood monocytes to free arachidonic acid (AA) results in the rapid induction of lipid droplet (LD) formation by these cells. This effect appears specific for AA in that it is not mimicked by other fatty acids, whether saturated or unsaturated. LDs are formed by two different routes: (i) the direct entry of AA into triacylglycerol and (ii) activation of intracellular signaling, leading to increased triacylglycerol and cholesteryl ester formation utilizing fatty acids coming from the de novo biosynthetic route. Both routes can be dissociated by the arachidonyl-CoA synthetase inhibitor triacsin C, which prevents the former but not the latter. LD formation by AA-induced signaling predominates, accounting for 60–70% of total LD formation, and can be completely inhibited by selective inhibition of the group IVA cytosolic phospholipase A2α (cPLA2α), pointing out this enzyme as a key regulator of AA-induced signaling. LD formation in AA-treated monocytes can also be blocked by the combined inhibition of the mitogen-activated protein kinase family members p38 and JNK, which correlates with inhibition of cPLA2α activation by phosphorylation. Collectively, these results suggest that concomitant activation of p38 and JNK by AA cooperate to activate cPLA2α, which is in turn required for LD formation possibly by facilitating biogenesis of this organelle, not by regulating neutral lipid synthesis.  相似文献   

7.
Alpha-lipoic acid (ALA) and its reduced form dihydrolipoic acid (DHLA) are powerful antioxidants both in hydrophilic and lipophylic environments with diverse pharmacological properties including anti-inflammatory activity. The mechanism of anti-inflammatory activity of ALA and DHALA is not known. The present study describes the interaction of ALA and DHALA with pro-inflammatory secretory PLA(2) enzymes from inflammatory fluids and snake venoms. In vitro enzymatic inhibition of sPLA(2) from Vipera russellii, Naja naja and partially purified sPLA(2) enzymes from human ascitic fluid (HAF), human pleural fluid (HPF) and normal human serum (HS) by ALA and DHLA was studied using (14)C-oleate labeled Escherichia coli as the substrate. Biophysical interaction of ALA with sPLA(2) was studied by fluorescent spectral analysis and circular dichroism studies. In vivo anti-inflammatory activity was checked using sPLA(2) induced mouse paw edema model. ALA but not DHLA inhibited purified sPLA(2) enzymes from V. russellii, N. naja and partially purified HAF, HPF and HS in a dose dependent manner. This data indicated that ALA is critical for inhibition. IC(50) value calculated for these enzymes ranges from 0.75 to 3.0 microM. The inhibition is independent of calcium and substrate concentration. Inflammatory sPLA(2) enzymes are more sensitive to inhibition by ALA than snake venom sPLA(2) enzymes. ALA quenched the fluorescence intensity of sPLA(2) enzyme in a dose dependent manner. Apparent shift in the far UV-CD spectra of sPLA(2) with ALA indicated change in its alpha-helical confirmation and these results suggest its direct interaction with the enzyme. ALA inhibits the sPLA(2) induced mouse paw edema in a dose dependent manner and confirms the sPLA(2) inhibitory activity in vivo also. These data suggest that ALA may act as an endogenous regulator of sPLA(2) enzyme activity and suppress inflammatory reactions.  相似文献   

8.
In rat luteal cells labeled with (3H]oleic acid, PGF-stimulated phospholipase D (PLD) activation was investigated. The PLD activity was detected by measuring the accumulation of [3H]phosphatidylethanol (PtdEt) in the presence of ethanol. PGF stimulated PtdEt accumulation at concentrations of more than 100 nM in the presence of ethanol. However, PtdEt accumulation did not change in the absence of ethanol. PGF (1 μM) increased PtdEt accumulation after 1 min, and the accumulation reached a plateau by 2–3 min. These results indicate that PGF activates PLD in rat luteal cells. U-73122, a phospholipase C (PLC) inhibitor, and staurosporine, a protein kinase C (PKC) inhibitor, did not inhibit PGF-stimulated [3H]PtdEt accumulation. These results suggest that PGF-induced PLD activation is different from PLC-PKC systems. We reported previously that PGF stimulated the release of arachidonic acid. The effects of indomethacin, nordihydroguaiaretic acid (NDGA), and 5,8,11,14-eicosatetraynoic acid (ETYA), inhibitors of arachidonic acid metabolism, on PGF-stimulated PtdEt accumulation were examined. Pretreatment with indomethacin enhanced PGF-induced PtdEt accumulation. In contrast, pretreatment with NDGA and ETYA inhibited PGF-induced PtdEt accumulation. It is suggested that PGF-stimulated PLD activation is mediated via lipoxygenase products.  相似文献   

9.
Depolarization-evoked increases in intraterminal free Ca2+ are required for the induction of neurotransmitter release from nerve terminals. Although the mechanisms that regulate the voltage-induced accumulation of presynaptic Ca2+ remain obscure, there is evidence that the phospholipase-dependent accumulation of arachidonic acid, or its metabolites, may be involved. Therefore, fura-2 loaded hippocampal mossy fiber nerve endings were used to investigate the relationships between membrane depolarization, lipid metabolism and presynaptic Ca2+ availability. It was observed that depolarization of the nerve terminals with KCl induced an increase in intraterminal free calcium that was inhibited more than 90% by a combination of voltage-sensitive Ca2+ channel blockers. In addition, the K+-dependent effects on Ca2+ concentrations were attenuated in the presence of phospholipase A2 inhibitors, but were mimicked by the phospholipase A2 activator melittin and exogenous arachidonic acid. Both the melittin- and arachidonic acid-induced increases in presynaptic Ca2+ were reduced by voltage-sensitive Ca2+ channel blockers. The stimulatory effects of arachidonic acid appeared to be independent of its further metabolism to prostaglandins. In fact, inhibition of either cyclooxygenase or lipoxygenase pathways resulted in a potentiation of the depolarization-evoked increase in intraterminal free Ca2+. From these results, we propose that some portion of the depolarization-evoked increase in intraterminal free calcium depends on the activation of phospholipase A2 and the subsequent accumulation of unesterified arachidonic acid.  相似文献   

10.
The association of cytosolic phospholipase A2-α (cPLA2α) with intracellular membranes is central to the generation of free arachidonic acid and thromboxane A2 in activated platelets. Despite this, the site and nature of this membrane association has not been fully characterised upon platelet activation. High resolution imaging showed that cPLA2α was distributed in a partly structured manner throughout the resting platelet. Upon glass activation or thrombin stimulation, cPLA2α relocated to a peripheral region corresponding to the platelet plasma membrane. Upon thrombin stimulation of platelets a major pool of cPLA2α was associated with the plasma membrane in an EGTA-resistant manner. EGTA-resistant membrane binding was abolished upon de-polymerisation of actin filaments by DNase I and furthermore, cPLA2α co-immunoprecipitated with actin upon thrombin stimulation of platelets. Immunofluorescence microscopy studies revealed that, upon platelet activation, cPLA2α and actin co-localised at the plasma membrane. Thus we have identified a novel mechanism for the interaction of cPLA2α with its membrane substrate via interaction with actin.  相似文献   

11.
Liu Y  Taylor CW 《FEBS letters》2006,580(17):4114-4120
Arachidonic acid (AA) regulates many aspects of vascular smooth muscle behaviour, but the mechanisms linking receptors to AA release are unclear. In A7r5 vascular smooth muscle cells pre-labelled with (3)H-AA, vasopressin caused a concentration-dependent stimulation of 3H-AA release that required phospholipase C and an increase in cytosolic [Ca2+]. Ca2+ release from intracellular stores and Ca2+ entry via L-type channels or the capacitative Ca2+ entry pathway were each effective to varying degrees. Selective inhibitors of PLA2 inhibited the 3H-AA release evoked by vasopressin, though not the underlying Ca2+ signals, and established that cPLA2 mediates the release of AA. We conclude that in A7r5 cells vasopressin stimulates AA release via a Ca2+-dependent activation of cPLA2.  相似文献   

12.
Phospholipase A(2) catalyses hydrolysis of the ester bond at the C2 position of 3-sn-phosphoglycerides. Here we report the 1.9A resolution crystal structure of the triple mutant K56,120,121M of bovine pancreatic phospholipase A(2). The structure was solved by molecular replacement method using the orthorhombic form of the recombinant phospholipase A(2). The final protein model contains all the 123 amino acid residues, two calcium ions, 125 water molecules and one 2-methyl-2-4-pentanediol molecule. The model has been refined to a crystallographic R-factor of 19.6% (R(free) of 25.9%) for all data between 14.2A and 1.9A. The residues 62-66, which are in a surface loop, are always disordered in the structures of bovine pancreatic phospholipase A(2) and its mutants. It is interesting to note that the residues 62-66 in the present structure is ordered and the conformation varies substantially from those in the previously published structures of this enzyme. An unexpected and interesting observation in the present structure is that, in addition to the functionally important calcium ion in the active site, one more calcium ion is found near the N terminus. Detailed structural analyses suggest that binding of the second calcium ion could be responsible for the conformational change and the ordering of the surface loop. Furthermore, the results suggest a structural reciprocity between the k(cat)(*) allosteric site and surface loop at the i-face, which represents a newly identified structural property of secreted phospholipase A(2).  相似文献   

13.
We previously demonstrated that the oxysterol potentiation of arachidonic acid release and prostaglandin biosynthesis induced by foetal calf serum activation of normal rat kidney (NRK) cells (fibroblastic clone 49F) was not related to a direct effect of oxysterols on cell free Ca2+ level. Since both Ca2+ variations and protein C are involved in arachidonic acid release in some models, we looked for a possible modulation by protein C in the oxysterol effect on arachidonic acid release. We show that when the phorbol ester 12-O-tetradecanoyl-phorbol-13acetate (TPA), a protein kinase C activator, was added to the culture medium, the oxyterol effect on arachidonic acid release and prostaglandin synthesis clearly increased. Moreover, the effect of TPA was dose-dependent and TPA EC50 (4 × 10−9 M) was unchanged in the presence of the oxysterol. Preincubation of cells with TPA for 24 h prevented the arachidonic acid release induced by TPA alone, whereas the oxysterol effect was decreased but not abolished. In the absence of serum, TPA and ionomycin added together induced the same noticeable (arachidonic acid) release and PGE2 synthesis as serum alone. Nevertheless, the potentiating effect of cholest-5-ene-3β,25-diol was much higher when serum itself was used to activate NRK cells than it was in the present serum-mimicking experimental conditions. Thus, the presence of growth factors is probably required to obtain a full oxysterol effect. We conclude that the oxysterol effect was synergistic with, but not fully dependent on, protein kinase C and Ca2+ ion fluxes, therefore oxysterols could affed earlier events triggered by serum growth factor binding to their cell membrane receptors.  相似文献   

14.
Previous studies demonstrated that melittin, the main peptide in bee venom, could cause persistent spontaneous pain, primary heat and mechanical hyperalgesia, and enhance the excitability of spinal nociceptive neurons. However, the underlying mechanism of melittin-induced cutaneous hypersensitivity is unknown. Effects of melittin applied topically to acutely dissociated rat dorsal root ganglion neurons were studied using whole-cell patch clamp and calcium imaging techniques. Melittin induced intracellular calcium increases in 60% of small (<25 μm) and medium (<40 μm) diameter sensory neurons. In current clamp, topical application of melittin evoked long-lasting firing in 55% of small and medium-sized neurons tested. In voltage clamp, melittin evoked inward currents in sensory neurons in a concentration-dependent manner. Repeated application of melittin caused increased amplitude of the inward currents. Most melittin-sensitive neurons were capsaicin-sensitive, and 65% were isolectin B4 positive. Capsazepine, the TRPV1 receptor inhibitor, completely abolished the melittin-induced inward currents and intracellular calcium transients. Inhibitions of signaling pathways showed that phospholipase A2, but not phospholipase C, was involved in producing the melittin-induced inward currents. Inhibitors of cyclooxygenases (COX) and lipoxygenases (LOX), two key components of the arachidonic acid metabolism pathway, each partially suppressed the inward current evoked by melittin. Inhibitors of protein kinase A (PKA), but not of PKC, also abolished the melittin-induced inward currents. These results indicate that melittin can directly excite small and medium-sized sensory neurons at least in part by activating TRPV1 receptors via PLA2-COXs/LOXs cascade pathways.  相似文献   

15.
Treatment of U937 cells with a sublethal concentration of tert-butylhydroperoxide generates DNA single strand breakage in U937 cells and this response is increased by caffeine, ATP, pyruvate or antimycin A. As we previously reported (Guidarelli, Clementi, Brambilla and Cantoni, (1997) Biochem. J. 328, 801–806), the enhancing effects of antimycin A are mediated by inhibition of complex III and the ensuing formation of superoxides and hydrogen peroxide in a reaction in which ubisemiquinone serves as an electron donor. Active electron transport was required in pyruvate-supplemented cells since the increased genotoxic response occurred as a consequence of enforced mitochondrial Ca2+ accumulation, a process driven by the increased electrochemical gradient. The enhancing effects of caffeine or ATP were also the consequence of mitochondrial Ca2+ accumulation but these responses were independent on electron transport. The increased formation of DNA lesions resulting from exposure to tert-butylhydroperoxide associated with the Ca2+-mobilizing agents or the respiratory substrate was mediated by arachidonic acid generated by Ca2+-dependent activation of phospholipase A2. Melittin, a potent phospholipase A2 activator, and reagent arachidonic acid mimicked the effects of caffeine, ATP or pyruvate on the tert-butylhydroperoxide-induced DNA single strand breakage.  相似文献   

16.
Docosahexaenoic acid (DHA) and arachidonic acid (AA) are found in high concentrations in brain cell membranes and are important for brain function and structure. Studies suggest that AA and DHA are hydrolyzed selectively from the sn-2 position of synaptic membrane phospholipids by Ca2+-dependent cytosolic phospholipase A2 (cPLA2) and Ca2+-independent phospholipase A2 (iPLA2), respectively, resulting in increased levels of the unesterified fatty acids and lysophospholipids. Cell studies also suggest that AA and DHA release depend on increased concentrations of Ca2+, even though iPLA2 has been thought to be Ca2+-independent. The source of Ca2+ for activation of cPLA2 is largely extracellular, whereas Ca2+ released from the endoplasmic reticulum can activate iPLA2 by a number of mechanisms. This review focuses on the role of Ca2+ in modulating cPLA2 and iPLA2 activities in different conditions. Furthermore, a model is suggested in which neurotransmitters regulate the activity of these enzymes and thus the balanced and localized release of AA and DHA from phospholipid in the brain, depending on the primary source of the Ca2+ signal.  相似文献   

17.
Duchenne muscular dystrophy is an inherited disease caused by the absence of dystrophin, a structural protein normally located under the sarcolemma of skeletal muscle fibers. Muscle degeneration occurring in this disease is thought to be partly caused by increased Ca2+ entry through sarcolemmal cationic channels. Using the Mn2+ quench method, we show here that Mn2+ entry triggered by Ca2+ store depletion but not basal Mn2+ entry relies on Ca2+-independent PLA2 (iPLA2) activity in dystrophic fibers isolated from a murine model of Duchenne muscular dystrophy, the mdx5cv mouse. iPLA2 was found to be localized in the vicinity of the sarcolemma and consistently, the iPLA2 lipid product lysophosphatidylcholine was found to trigger Ca2+ entry through sarcolemmal channels, suggesting that it acts as an intracellular messenger responsible for store-operated channels opening in dystrophic fibers. Our results suggest that inhibition of iPLA2 and lysophospholipid production may be of interest to reduce Ca2+ entry and subsequent degeneration of dystrophic muscle.  相似文献   

18.
Therapy with interleukin-2 (IL-2) induces remissions in some forms of cancer. This treatment however, is accompanied by side-effects which, in part, may be mediated by the formation of eicosanoids and plateletactivating factor. We investigated the systemic release of phospholipase A2 (PLA2), a rate-limiting enzyme in the formation of these lipid mediators, in patients receiving IL-2. In a pilot study of 4 patients we observed an increase in PLA2 activity in serial plasma samples obtained during the first day after a bolus infusion of IL-2, which increase closely correlated with that of antigen levels of secretory phospholipase A2 (sPLA2) as measured by enzyme-linked immunosorbent assay (r=0.92;P<0.001). In 20 patients, receiving 12×106–18×106 IU IL-2/m2, we then investigated the course of antigenic levels of sPLA2 in relation to those of the cytokines tumour necrosis factor (TNF) and interleukin-6 (IL-6) (both cytokines may induce sPLA2 in vivo). From 4 h on, sPLA2 levels significantly increased, reaching a peak 24 h after the IL-2 infusion. Subsequent IL-2 infusions even induced a further increase of sPLA2. This increase of sPLA2 was presumably not due to a direct effect of IL-2 on, for example, hepatocytes, since this cytokine, in contrast to IL-1, IL-6, TNF and interferon , was not able to induce the synthesis of sPLA2 by Hep G2 cells in vitro. Consistent with this, plasma levels of TNF and IL-6 in the patients rose, reaching peak levels before a zenith of sPLA2 occurred, i.e at 2 h and 4 h after the start of the IL-2 infusion respectively. sPLA2 levels significantly correlated with the development of the side-effects increase in body weight (r=0.49;P<0.0001) and decrease in mean arterial blood pressure (r=0.40;P<0.0001). Moreover, maximum sPLA2 levels induced by IL-2 were higher in patients who had progressive disease after therapy than in patients who had stable disease or a partial response.  相似文献   

19.
Effects of L-histidine and related compounds on protein synthesiswere studied in cultured L6 myotubes. L-Histidine specifically stimulated protein synthesis, whereas D-histidine, histamine, L-arginine and L-lysine did not. Inhibitors of phospholipase A2, phospholipase C and cyclooxygenase intercepted the stimulatory action of L-histidine on protein synthesis, while inhibitors of protein kinase C and 5-lipoxygenase did not. These results suggest an involvement of phospholipase A2 and cyclooxygenase in the stimulatory action of L-histidine on protein synthesis in L6 myotubes. This revised version was published online in August 2006 with corrections to the Cover Date.  相似文献   

20.
R180, isolated from porcine brain cortex, is a high-affinity membrane receptor for ammodytoxin A (AtxA), a secreted phospholipase A(2) (sPLA(2)) and presynaptically active neurotoxin from venom of the long-nosed viper (Vipera ammodytes ammodytes). As a member of the M-type sPLA(2) receptors, present on the mammalian plasma membrane, R180 has been proposed to be responsible for one of the first events in the process of presynaptic neurotoxicity, the binding of the toxin to the nerve cell. To test this hypothesis, we prepared and analyzed three N-terminal fusion proteins of AtxA possessing a 12 or 5 amino acid residue peptide. The presence of such an additional "propeptide" prevented interaction of the toxin with the M-type receptor but not its lethality in mouse and neurotoxic effects on a mouse phrenic nerve-hemidiaphragm preparation. In addition, antibodies raised against the sPLA(2)-binding C-type lectin-like domain 5 of the M-type sPLA(2) receptor were unable to abolish the neurotoxic action of AtxA on the neuromuscular preparation. The specific enymatic activities of the fusion AtxAs were two to three orders of magnitude lower from that of the wild type, yet resulting in a similar but less pronounced neurotoxic profile on the neuromuscular junction. This is in accordance with other data showing that a minimal enzymatic activity suffices for presynaptic toxicity of sPLA(2)s to occur. Our results indicate that the interaction of AtxA with the M-type sPLA(2) receptor at the plasma membrane is not essential for presynaptic activity of the toxin. Interaction of AtxA with two intracellular proteins, calmodulin and the R25 receptor, was affected but not prevented by the presence of the N-terminal fusion peptides, implying that these proteins may play a role in the sPLA(2) neurotoxicity.  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