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 31 毫秒
1.
Normal human lymphocytes resisted the hydrolytic action of secretory phospholipase A2 but became susceptible to the enzyme following treatment with a calcium ionophore, ionomycin. To test the hypothesis that this susceptibility requires exposure of the anionic lipid phosphatidylserine on the external face of the cell membrane, experiments were repeated with a human Burkitt's lymphoma cell line (Raji cells). In contrast to normal lymphocytes or S49 mouse lymphoma cells, most of the Raji cells (83%) did not translocate phosphatidylserine to the cell surface upon treatment with ionomycin. Those few that did display exposed phosphatidylserine were hydrolyzed immediately upon addition of phospholipase A2. Interestingly, the remaining cells were also completely susceptible to the enzyme but were hydrolyzed at a slower rate and after a latency of about 100 s. In contradistinction to the defect in phosphatidylserine translocation, Raji cells did display other physical membrane changes upon ionomycin treatment that may be relevant to hydrolysis by phospholipase A2. These changes were detected by merocyanine 540 and trimethylammonium diphenylhexatriene fluorescence and were common among normal lymphocytes, S49 cells, and Raji cells. The levels of these latter effects corresponded well with the relative rates of hydrolysis among the three cell lines. These results suggested that while phosphatidylserine enhances the rate of cell membrane hydrolysis by secretory phospholipase A2, it is not an absolute requirement. Other physical properties such as membrane order contribute to the level of membrane susceptibility to the enzyme independent of phosphatidylserine.  相似文献   

2.
During cell death, plasma membranes of cells become vulnerable to attack by extracellular secretory phospholipase A(2). The purpose of this study was to identify the timing of this phenomenon relative to other events that occur during the process of cell death. Death was induced in S49 murine lymphoma cells by treatment with dexamethasone, dibutyryl cAMP, ionomycin, thapsigargin, or heat shock (1 h at 43 degrees C). The appearance of membrane susceptibility to secretory phospholipase A(2) was compared to the following apoptotic events: loss of mitochondrial membrane potential, phosphatidylserine exposure in the outer leaflet of the cell membrane, early DNA damage assessed by the comet assay, and changes in cell size and internal complexity assessed by flow cytometry. Each inducer of death was distinct in the time course of events produced. Although dead cells were susceptible to the action of phospholipase A(2), live cells (impermeable to propidium iodide) also became vulnerable to the enzyme during characteristic time courses after exposure to each inducer. In fact, susceptibility to sPLA(2) was observed in each case prior to or concurrent with the earliest of the markers of apoptosis. These results demonstrate that the onset of susceptibility to sPLA(2) is an early event in apoptosis suggesting that changes in membrane structure may be relevant to initial aspects of the apoptotic process.  相似文献   

3.
Exposure of S49 lymphoma cells to exogenous group IIA or V secretory phospholipase A2 (sPLA2) caused an initial release of fatty acid followed by resistance to further hydrolysis by the enzyme. This refractoriness was overcome by exposing cells to palmitoyl lysolecithin. This effect was specific in terms of lysophospholipid structure. Induction of membrane susceptibility by lysolecithin involved an increase in cytosolic calcium and was duplicated by incubating the cells with calcium ionophores such as ionomycin. Lysolecithin also activated cytosolic phospholipase A2 (cPLA2). Inhibition of this enzyme attenuated the ability of lysolecithin (but not ionomycin) to induce susceptibility to sPLA2. Lysolecithin or ionomycin caused concurrent hydrolysis of both phosphatidylethanolamine and phosphatidylcholine implying that transbilayer movement of phosphatidylethanolamine occurred upon exposure to these agents but that susceptibility is not simply due to exposure of a preferred substrate (i.e. phosphatidylethanolamine) to the enzyme. Microvesicles were apparently released from the cells upon addition of lysolecithin or ionomycin. Both these vesicles and the remnant cell membranes were susceptible to sPLA2. Together these data suggest that lysolecithin induces susceptibility through both cPLA2-dependent and -independent pathways. Whereas elevated cytosolic calcium was required for both pathways, it was sufficient only for the cPLA2-independent pathway. This cPLA2-independent pathway involved changes in cell membrane structure associated with transbilayer phospholipid migration and microvesicle release.  相似文献   

4.
During apoptosis, a number of physical changes occur in the cell membrane including a gradual increase in permeability to vital stains such as propidium iodide. This study explored the possibility that one consequence of membrane changes concurrent with early modest permeability is vulnerability to degradation by secretory phospholipase A(2). The activity of this hydrolytic enzyme toward mammalian cells depends on the health of the cell; healthy cells are resistant, but they become susceptible early during programmed death. Populations of S49 lymphoma cells during programmed death were classified by flow cytometry based on permeability to propidium iodide and susceptibility to secretory phospholipase A(2). The apoptotic inducers thapsigargin and dexamethasone caused modest permeability to propidium iodide and increased staining by merocyanine 540, a dye sensitive to membrane perturbations. Various secretory phospholipase A(2) isozymes (human groups IIa, V, X, and snake venom) preferentially hydrolyzed the membranes of cells that displayed enhanced permeability. In contrast, cells exposed briefly to a calcium ionophore showed the increase in cell staining intensity by merocyanine 540 without accompanying uptake of propidium iodide. Under that condition, only the snake venom and human group X enzymes hydrolyzed cells that were dying. These results suggested that cells showing modest permeability to propidium iodide during the early phase of apoptosis are substrates for secretory phospholipase A(2) and that specificity among isoforms of the enzyme depends on the degree to which the membrane has been perturbed during the death process. This susceptibility to hydrolysis may be important as part of the signal to attract macrophages toward apoptotic cells.  相似文献   

5.
The membranes of healthy lymphocytes normally resist hydrolysis by secretory phospholipase A2. However, they become susceptible during the process of apoptosis. Previous experiments have demonstrated the importance of certain physical changes to the membrane during cell death such as a reduction in membrane lipid order and exposure of phosphatidylserine on the membrane surface. Nevertheless, those investigations also showed that at least one additional factor was required for rapid hydrolysis by the human group IIa phospholipase isozyme. This study was designed to test the possibility that oxidation of membrane lipids is the additional factor. Flow cytometry and confocal microscopy with a fluorescent probe of oxidative potential suggested that oxidation of the plasma membrane occurs during apoptosis stimulated by thapsigargin. When oxidative potential was high, the activity of human group IIa secretory phospholipase A2 was enhanced 30- to 100-fold compared to that observed with conditions sufficient for maximal hydrolysis by other secretory phospholipase A2 isoforms. Direct oxidation of cell membranes with either of two oxidizing agents also stimulated hydrolysis by secretory phospholipase A2. Both oxidizers caused externalization of phosphatidylserine, but a change in lipid order did not always occur. These results demonstrated that membrane oxidation strongly stimulates human group IIa secretory phospholipase A2 activity toward apoptotic cells. Interestingly, the change in membrane order, previously thought to be imperative for high rates of hydrolysis, was not required when membrane lipids were oxidized. Whether phosphatidylserine exposure is still necessary with oxidation remains unresolved since the two events could not be deconvoluted.  相似文献   

6.
During apoptosis, physical changes in the plasma membrane prepare the cell for clearance by phagocytes and hydrolysis by secretory phospholipase A2 (sPLA2). The relationships among these changes have not been adequately established, especially for hormone-stimulated apoptosis. This study addresses these issues for glucocorticoid-induced apoptosis in S49 lymphoma cells. Flow cytometry, microscopy, and fluorescence spectroscopy were used to assess merocyanine 540 emission, laurdan generalized polarization, phosphatidylserine exposure, caspase activation, and membrane permeability to propidium iodide in the absence and presence of sPLA2. The earliest event observed was activation of cellular caspases. Results with membrane probes suggest that interlipid spacing also increases early during apoptosis and precedes transbilayer migration of phosphatidylserine, DNA fragmentation, and a general increase in lipid order associated with blebbing and dissolution of the cells. The activity of sPLA2 appeared to be linked more to lipid spacing than to loss of membrane asymmetry. The early nature of some of these events and their ability to promote activity of a proinflammatory enzyme suggests the possibility of an inflammatory response during T-lymphocyte apoptosis.  相似文献   

7.
A diminution in the order of membrane lipids, which occurs during apoptosis, has been shown to correlate with increased membrane susceptibility to hydrolysis by secretory phospholipase A2. Studies with artificial membranes, however, have demonstrated that the relationship between membrane order and hydrolysis is more complex than suggested thus far by cell studies. To better resolve this relationship, this study focused on comparisons between increasing temperature and calcium ionophore as means of decreasing membrane order in S49 cells. Although these two treatments caused comparable changes in apparent membrane order as detected by steady-state fluorescence measurements, only ionophore treatment enhanced phospholipase activity. Experiments with exogenously-added phosphatidylserine indicated that the difference was not due to the presence of that anionic phospholipid in the outer membrane leaflet. Instead, analysis of the equilibration kinetics of various cationic membrane probes revealed that the difference could relate to the spacing of membrane lipids. Specifically, ionophore treatment increased that spacing while temperature only affected overall membrane order and fluidity. To consider the possibility that the distinction with ionophore might relate to the actin cytoskeleton, cells were stained with phalloidin and imaged via confocal microscopy. Ionophore caused disruption of actin fibers while increased temperature did not. This apparent connection between membrane hydrolysis and the cytoskeleton was further corroborated by examining the relationship among these events during apoptosis stimulated by thapsigargin.  相似文献   

8.
Arachidonoyl-hydrolyzing phospholipase A2 plays a central role in providing substrate for the synthesis of the potent lipid mediators of inflammation, the eicosanoids, and platelet-activating factor. Although Ca2+ is required for arachidonic acid release in vivo and most phospholipase A2 enzymes require Ca2+ for activity in vitro, the role of Ca2+ in phospholipase A2 activation is not understood. We have found that an arachidonoyl-hydrolyzing phospholipase A2 from the macrophage-like cell line, RAW 264.7, exhibits Ca2(+)-dependent association with membrane. The intracellular distribution of the enzyme was studied as a function of the Ca2+ concentration present in homogenization buffer. The enzyme was found almost completely in the 100,000 x g soluble fraction when cells were homogenized in the presence of Ca2+ chelators and there was a slight decrease in soluble fraction activity when cells were homogenized at the level of Ca2+ in an unstimulated cell (80 nM). When cells were homogenized at Ca2+ concentrations expected in stimulated cells (230-450 nM), 60-70% of the phospholipase A2 activity was lost from the soluble fraction and became associated with the particulate fraction in a manner that was partly reversible with EGTA. Membrane-associated phospholipase A2 activity was demonstrated by [3H]arachidonic acid release both from exogenous liposomes and from radiolabeled membranes. With radiolabeled particulate fraction as substrate, this enzyme hydrolyzed arachidonic acid but not oleic acid from membrane phospholipid, and [3H]arachidonic acid was derived from phosphatidylcholine, phosphatidylethanolamine, and phosphatidylinositol/phosphatidylserine. We suggest a mechanism in which the activity of phospholipase A2 is regulated by Ca2+: in an unstimulated cell phospholipase A2 is found in the cytosol; upon receptor ligation the cytosolic Ca2+ concentration increases, and the enzyme becomes membrane-associated which facilitates arachidonic acid hydrolysis.  相似文献   

9.
Best KB  Ohran AJ  Hawes AC  Hazlett TL  Gratton E  Judd AM  Bell JD 《Biochemistry》2002,41(47):13982-13988
Normally, cell membranes resist hydrolysis by secretory phospholipase A(2). However, upon elevation of intracellular calcium, the cells become susceptible. Previous investigations demonstrated a possible relationship between changes in lipid order caused by increased calcium and susceptibility to phospholipase A(2). To further explore this relationship, we used temperature as an experimental means of manipulating membrane physical properties. We then compared the response of human erythrocytes to calcium ionophore at various temperatures in the range of 20-50 degrees C using fluorescence spectroscopy and two-photon fluorescence microscopy. The steady state fluorescence emission of the environment-sensitive probe, laurdan, revealed that erythrocyte membrane order decreases systematically with temperature throughout this range, especially between 28 and 45 degrees C. Furthermore, the ability of calcium ionophore to induce increased membrane order and susceptibility to phospholipase A(2) depended similarly on temperature. Both responses to calcium influx were enhanced as membrane fluidity increased. Analysis of the spatial distribution of laurdan fluorescence at several temperatures indicated that the ordering effect of intracellular calcium on fluid membranes generates an increase in the number of fluid-solid boundaries. Hydrolysis of the membrane appeared to progress outward from these boundaries. We conclude that phospholipase A(2) prefers to hydrolyze lipids in fluid regions of human erythrocyte membranes, but primarily when those regions coexist with domains of ordered lipids.  相似文献   

10.
The hemolytic actions of three kinds of phospholipase C on horse and sheep erythrocytes were studied in relation to their hydrolytic activities on the phospholipid components of these red cells. Clostridium novyi (oedematiens) type A phospholipase C hemolyzed horse red cells by hydrolyzing phosphatidylcholine. However, the enzyme did not lyse sheep cells nor did it hydrolyze any phospholipid under the same conditions, although this enzyme hydrolyzed both sphingomyelin and phosphatidylethanolamine in the phospholipid mixture extracted from sheep red cells. Clostridium perfringens phospholipase C hemolyzed not only horse red cells by hydrolyzing phosphatidylcholine but also sheep red cells by hydrolyzing sphingomyelin. Sphingomyelin on sheep red cell membrane was hydrolyzed 10 times faster by this enzyme than that on horse red cell membrane. Pseudomonas aureofaciens phospholipase C hemolyzed horse red cells by attacking phosphatidylcholine and phosphatidylethanolamine. The enzyme did not attack sheep red cells but it did hydrolyze phosphatidylethanolamine in the extracted phospholipid mixture from sheep cells. The hemolytic activity of phospholipase C depends not only on the enzyme and the asymmetric distribution of phospholipids in the erythrocyte membrane but also on the accessibility of the enzymes to the phospholipids in the surface of the membranes. Hemolysis by phospholipase C belongs to a hot-cold type of lysis.  相似文献   

11.
The phospholipids of intact microsomal membranes were hydrolysed 50% by phospholipase C of Clostridium welchii, without loss of the secretory protein contents of the vesicle, which are therefore not permeable to the phospholipase. Phospholipids extracted from microsomes and dispersed by sonication were hydrolysed rapidly by phospholipase C-Cl. welchii with the exception of phosphatidylinositol. Assuming that only the phospholipids of the outside of the bilayer of the microsomal membrane are hydrolysed in intact vesicles, the composition of this leaflet was calculated as 84% phosphatidylcholine, 8% phosphatidylethanolamine, 9% sphingomyelin and 4% phosphatidylserine, and that of the inner leaflet 28% phosphatidylcholine, 37% phosphatidylethanolamine, 6% phosphatidylserine and 5% sphingomyelin. Microsomal vesicles were opened and their contents released in part by incubation with deoxycholate (0.098%) lysophosphatidylcholine (0.005%) or treatment with the French pressure cell. Under these conditions, hydrolysis of the phospholipids by phospholipase C-Cl. welchii was increased and this was mainly due to increased hydrolysis of those phospholipids assigned to the inner leaflet of the bilayer, phosphatidylethanolamine and phosphatidylserine. Phospholipase A2 of bee venom and phospholipase C of Bacillus cereus caused rapid loss of vesicle contents and complete hydrolysis of the membrane phospholipids, with the exception of sphingomyelin which is not hydrolysed by the former enzyme.  相似文献   

12.
Exposure of human erythrocytes to the calcium ionophore ionomycin rendered them susceptible to the action of secretory phospholipase A(2) (sPLA(2)). Analysis of erythrocyte phospholipid metabolism by thin-layer chromatography revealed significant hydrolysis of both phosphatidylcholine and phosphatidylethanolamine during incubation with ionomycin and sPLA(2). Several possible mechanisms for the effect of ionomycin were considered. Involvement of intracellular phospholipases A(2) was excluded since inhibitors of these enzymes had no effect. Assessment of membrane oxidation by cis-parinaric acid fluorescence and comparison to the oxidants diamide and phenylhydrazine revealed that oxidation does not participate in the effect of ionomycin. Incubation with ionomycin caused classical physical changes to the erythrocyte membrane such as morphological alterations (spherocytosis), translocation of aminophospholipids to the outer leaflet of the membrane, and release of microvesicles. Experiments with phenylhydrazine, KCl, quinine, merocyanine 540, the calpain inhibitor E-64d, and the scramblase inhibitor R5421 revealed that neither phospholipid translocation nor vesicle release was required to induce susceptibility. Results from fluorescence spectroscopy and two-photon excitation scanning microscopy using the membrane probe laurdan argued that susceptibility to sPLA(2) is a consequence of increased order of membrane lipids.  相似文献   

13.
The role of phospholipids in the binding of 125I-choriogonadotropin to bovine corpus luteum plasma membranes has been investigated with the use of purified phospholipase A and phospholipase C to alter membrane phospholipids. The phospholipase C-digested plasma membrane preparation showed 85 to 90% inhibition of 125I-choriogonadotropin binding activity when 70% of the membrane phospholipid was hydrolyzed. Similarly treatment of plasma membranes with phospholipase A resulted in 45 to 55% hydrolysis of membrane phospholipid and almost 75% inhibition of receptor activity. Both these enzymes hydrolyzed membrane-associated phosphatidylcholine to a greater extent than phosphatidylethanolamine and phosphatidylserine. Phosphorylaminoalcohols of phospholiphase C end products were completely released into the medium, while phospholipase A by-products remained associated with plasma membranes. Addition of a phospholipids suspension or liposomes to plasma membranes pretreated with phospholipase A and C did not restore gonadotropin binding activity. Soluble phosphorylcholine, phosphorylethanolamine, and phosphorylserine and insoluble diglyceride products of phospholipase C action had no effect on receptor activity. In contrast, end products of the phospholipase A action, such as lysophosphatides and fatty acids, inhibited both on the membrane-associated and solubilized receptor activity. Lysophosphatidylcholine was the most effective end product inhibiting the binding of gonadotropin to the receptor, followed by lysophosphatidylethanolamine and lysophosphatidylserine. The inhibitory effects of phospholipase A or lysophosphatides were completely reversed upon removal of membrane-bound phospholipid end products by washing the membranes with defatted bovine serum albumin. However, phospholipase C inhibition could not be overcome by defatted albumin washings. Solubilization of plasma membranes with detergents which had been pretreated with phospholipase C partially restored the inhibited activity. It is concluded that the phospholipase-mediated inhibition of gonadotropin binding activity was due to hydrolysis and alterations of the phospholipid environment in the case of phospholipase C and by direct inhibition by end products in the case of phospholipase A.  相似文献   

14.
Elevated concentrations of intracellular calcium in erythrocytes increase membrane order and susceptibility to secretory phospholipase A2. We hypothesize that calcium aids the formation of domains of ordered lipids within erythrocyte membranes by interacting directly with the inner leaflet of the cell membrane. The interface of these domains with regions of more fluid lipids may create an environment with weakened neighbor-neighbor interactions that would facilitate phospholipid migration into the active site of bound secretory phospholipase A2. This hypothesis was investigated by determining the effects of seven other divalent ions on erythrocyte membrane properties. Changes in membrane order were assessed with steady-state fluorescence spectroscopy and two-photon microscopy with an environment-sensitive probe, laurdan. Each ion increased apparent membrane order in model membranes and in erythrocytes when introduced with an ionophore, suggesting that direct binding to the inner face of the membrane accounts for the effects of calcium on membrane fluidity. Furthermore, the degree to which ions affected membrane properties correlated with the ionic radius and electronegativity of the ions. Lastly, erythrocytes became more susceptible to enzyme hydrolysis in the presence of elevated intracellular levels of nickel and manganese, but not magnesium. These differences appeared related to the ability of the ions to induce a transition in erythrocyte shape.  相似文献   

15.
S Gatt  B Morag    S Rottem 《Journal of bacteriology》1982,151(3):1095-1101
Mycoplasma gallisepticum strains have a membrane-bound lysophospholipase which hydrolyzes lysophospholipid generated in these membranes by treatment with an external phospholipase. This paper studies the hydrolysis of the membranous lysophospholipids by an enzyme residing in the same membrane (intramembrane utilization) or in adjacent membranes (intermembrane utilization). To study intermembrane hydrolysis, the phospholipids of M. gallisepticum were labeled with [3H]oleic acid. Membranes were prepared, heated at 65 degrees C, and subsequently treated with pancreatic phospholipase A2. This resulted in membranes whose enzyme was heat inactivated, but which contained lysophospholipid. When these membranes were mixed with M. gallisepticum cells or membranes, the lysophospholipid was hydrolyzed by the membranous lysophospholipase. To study intramembrane hydrolysis, [3H]oleyl-labeled membranes of M. gallisepticum were treated with pancreatic phospholipase A2 at pH 5.0. At this pH, lysophospholipid was generated but not hydrolyzed. Adjustment of the pH to 7.4 resulted in hydrolysis of the lysophospholipid by the membranous lysophospholipase. These procedures permitted measuring the initial rates of intramembrane and intermembrane hydrolysis of the lysophospholipid, showing that the time course and dependence on endogenous substrate concentration were different in the intramembrane and intermembrane modes of utilization. They also permitted calculation of the molar concentration of the lysophospholipid in the membrane and its rate of hydrolysis, expressed as moles per minute per cell or per square centimeter of cell surface.  相似文献   

16.
Apoptosis is generally accompanied by a late phase of ceramide (Cer) production, the significance of which is unknown. This study describes a previously unrecognized link between Cer accumulation and phosphatidylserine (PS) exposure at the cell surface, a characteristic of the execution phase of apoptosis resulting from a loss of plasma membrane phospholipid asymmetry. Using a fluorescent sphingomyelin (SM) analogue, N-(N-[6-[(7-nitrobenz-2-oxa-1, 3-diazol-4-yl)amino]caproyl]-sphingosylphosphorylcholine (C(6)-NBD-SM), we show that Cer is derived from SM, initially located in the outer leaflet of the plasma membrane, which gains access to a cytosolic SMase by flipping to the inner leaflet in a process of lipid scrambling paralleling PS externalization. Lipid scrambling is both necessary and sufficient for SM conversion: Ca(2+) ionophore induces both PS exposure and SM hydrolysis, whereas scrambling-deficient Raji cells do not show PS exposure or Cer formation. Cer is not required for mitochondrial or nuclear apoptotic features since these are still observed in Raji cells. SM hydrolysis facilitates cholesterol efflux to methyl-beta-cyclodextrin, which is indicative of a loss of tight SM-cholesterol interaction in the plasma membrane. We provide evidence that these biophysical alterations in the lipid bilayer are essential for apoptotic membrane blebbing/vesiculation at the cell surface: Raji cells show aberrant apoptotic morphology, whereas replenishment of hydrolyzed SM by C(6)- NBD-SM inhibits blebbing in Jurkat cells. Thus, SM hydrolysis, during the execution phase of apoptosis, results from a loss of phospholipid asymmetry and contributes to structural changes at the plasma membrane.  相似文献   

17.
Rat platelets released phospholipase A2 and lysophospholipase upon activation with thrombin or ADP. The release of phospholipases was energy-dependent and was not in parallel with that of a known lysosomal marker enzyme, N-acetyl-beta-D-glucosaminidase. The phospholipases are derived from other granules (dense granules or alpha-granules) rather than lysosomal granules of the cells. All of the activities of both phospholipases in the cell free fraction obtained from the activated platelet reaction mixture was recovered in the supernatant after centrifugation at 105,000 X g. The degree of hydrolysis of phospholipids by the phospholipase A2 followed the order: phosphatidylethanolamine (PE) greater than phosphatidylserine (PS) greater than phosphatidylcholine (PC). Phospholipase A2 shows a broad pH optimum (greater than pH 7.0) and absolutely requires Ca2+. Lysophospholipase was specific to lysophosphatidylserine (lysoPS), and neither lysophosphatidylethanolamine (lysoPE) nor lysophosphatidylcholine (lysoPC) was hydrolyzed appreciably. Both 1-acyl- and 2-acyl-lysophosphatidylserine were equally hydrolyzed. Lysophospholipase activity shows similar pH optimum to phospholipase A2. The lysophospholipase activity was lost easily at 60 degrees C. The activity was reduced by the presence of EDTA, though low but distinct activity was observed even in the presence of EDTA. Addition of Ca2+ to the mixtures restores the full activity.  相似文献   

18.
Protoplasts prepared from Bacillus subtilis by lysozyme digestion lysed in the presence of pure pancreatic phospholipase A(2). The phospholipids cardiolipin, phosphatidylethanolamine, phosphatidylglycerol and lysylphosphatidylglycerol, which are present in the membrane, are degraded by phospholipase A(2) only after removal of the cell wall, giving free fatty acids and lyso derivatives. The four phospholipids are hydrolyzed equally well at a given enzyme concentration. Differences in the phospholipid composition of the protoplasts were obtained by variations in the growth medium, time of harvesting, and preincubation time with lysozyme. The extent of hydrolysis appeared to depend on the initial phospholipid composition. A relative increase in acidic phospholipids in the membrane facilitated the action of phospholipase A(2), whereas the rate of hydrolysis was diminished when protoplasts were tested which contained a relatively high amount of positively charged phospholipid. Pure phospholipase C from B. cereus preferentially hydrolyzed phosphatidyl-ethanolamine in the B. subtilis membrane. More than 80% of this phospholipid was converted into diglyceride, whereas only 30% of the cardiolipin was hydrolyzed. Such a loss of phospholipids, however, was not followed by lysis of the protoplasts. Liposomes were prepared from the lipid extracts of B. subtilis and incubated with both phospholipases. The hydrolysis pattern of the phospholipids in these model membrane systems was identical to the hydrolysis pattern of the phospholipids in the protoplast membrane. Phospholipase A(2) hydrolyzed all the phospholipids in the liposomes equally well, whereas phospholipase C preferentially degraded phosphatidylethanolamine.  相似文献   

19.
Although cell membranes normally resist the hydrolytic action of secretory phospholipase A(2) (sPLA(2)), they become susceptible during apoptosis or after cellular trauma. Experimentally, susceptibility to the enzyme can be induced by loading cells with calcium. In human erythrocytes, the ability of the calcium ionophore to cause susceptibility depends on temperature, occurring best above approximately 35 degrees C. Considerable evidence from experiments with artificial bilayers suggests that hydrolysis of membrane lipids requires two steps. First, the enzyme adsorbs to the membrane surface, and second, a phospholipid diffuses from the membrane into the active site of the adsorbed enzyme. Analysis of kinetic experiments suggested that this mechanism can explain the action of sPLA(2) on erythrocyte membranes and that temperature and calcium loading promote the second step. This conclusion was further supported by binding experiments and assessment of membrane lipid packing. The adsorption of fluorescent-labeled sPLA(2) was insensitive to either temperature or ionophore treatment. In contrast, the fluorescence of merocyanine 540, a probe sensitive to lipid packing, was affected by both. Lipid packing decreased modestly as temperature was raised from 20 to 60 degrees C. Calcium loading enhanced packing at temperatures in the low end of this range, but greatly reduced packing at higher temperatures. This result was corroborated by measurements of the rate of extraction of a fluorescent phosphatidylcholine analog from erythrocyte membranes. Furthermore, drugs known to inhibit susceptibility in erythrocytes also prevented the increase in phospholipid extraction rate. These results argue that the two-step model applies to biological as well as artificial membranes and that a limiting step in the hydrolysis of erythrocyte membranes is the ability of phospholipids to migrate into the active site of adsorbed enzyme.  相似文献   

20.
After incubation of human erythrocytes at 37 degrees C in the absence of glucose (A) for 24 h, (B) for 4 h with 8 mM hexanol or (C) for 3 h with SH reagents, phosphatidylethanolamine becomes partly susceptible to hydrolysis by phospholipase A2 from Naja naja. The presence of glucose during the pretreatments suppresses this effect, except in the case of SH reagents that inhibit glycolysis. After incubation with tetrathionate, up to 45% of the phosphatidylethanolamine is degraded by the enzyme, an amount considerably in excess of the 20% attacked in fresh erythrocytes. Pancreatic phospholipase A2, an enzyme unable to hydrolyse the phospholipids of intact erythrocytes, partially degrades phosphatidylcholine and phosphatidylethanolamine of erythrocytes pretreated with hexanol or SH reagents. Reagents capable of oxidizing SH groups to disulfides (tetrathionate, o-iodosobenzoate and hydroquinone) even render susceptible to pancreatic phospholipase A2 phosphatidylserine, a phospholipid supposed to be entirely located in the inner lipid layer of the membrane. Alkylating or acylating SH reagents have no such effect. It is postulated that disulfide bond formation between membrane protein SH groups leads to an alteration in protein-phospholipid interactions and consequently induces a reorientation of phospholipids between the inner and the outer membrane lipid layer.  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