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.
N E Gabriel  M F Roberts 《Biochemistry》1987,26(9):2432-2440
Asymmetric unilamellar vesicles are produced when short-chain phospholipids (fatty acyl chain lengths of 6-8 carbons) are mixed with long-chain phospholipids (fatty acyl chain lengths of 14 carbons or longer) in ratios of 1:4 short-chain/long-chain component. Short-chain lecithins are preferentially distributed on the outer monolayer, while a short-chain phosphatidylethanolamine derivative appears to localize on the inner monolayer of these spontaneously forming vesicles. Lanthanide NMR shift experiments clearly show a difference in head-group/ion interactions between the short-chain and long-chain species. Two-dimensional 1H NMR studies reveal efficient spin diffusion networks for the short-chain species embedded in the long-chain bilayer matrix. The short-chain lecithin is considerably more mobile than the long-chain component but has hindered motion compared to short-chain lecithin micelles. This differentiation in physical characteristics of the two phospholipid components is critical to understanding the activity of phospholipases toward these binary systems.  相似文献   

2.
N E Gabriel  M F Roberts 《Biochemistry》1986,25(10):2812-2821
Stable unilamellar vesicles formed spontaneously upon mixing aqueous suspensions of long-chain phospholipid (synthetic, saturated, and naturally occurring phosphatidylcholine, phosphatidylethanolamine, and sphingomyelin) with small amounts of short-chain lecithin (fatty acid chain lengths of 6-8 carbons) have been characterized by using NMR spectroscopy, negative staining electron microscopy, differential scanning calorimetry, and Fourier transform infrared (FTIR) spectroscopy. This method of vesicle preparation can produce bilayer vesicles spanning the size range 100 to greater than 1000 A. The combination of short-chain lecithin and long-chain lecithin in its gel state at room temperature produces relatively small unilamellar vesicles, while using long-chain lecithin in its liquid-crystalline state produces large unilamellar vesicles. The length of the short-chain lecithin does not affect the size distribution of the vesicles as much as the ratio of short-chain to long-chain components. In general, additional short-chain decreases the average vesicle size. Incorporation of cholesterol can affect vesicle size, with the solubility limit of cholesterol in short-chain lecithin micelles governing any size change. If the amount of cholesterol is below the solubility limit of micellar short-chain lecithin, then the addition of cholesterol to the vesicle bilayer has no effect on the vesicle size; if more cholesterol is added, particle growth is observed. Vesicles formed with a saturated long-chain lecithin and short-chain species exhibit similar phase transition behavior and enthalpy values to small unilamellar vesicles of the pure long-chain lecithin prepared by sonication. As the size of the short-chain/long-chain vesicles decreases, the phase transition temperature decreases to temperatures observed for sonicated unilamellar vesicles. FTIR spectroscopy confirms that the incorporation of the short-chain lipid in the vesicle bilayer does not drastically alter the gauche bond conformation of the long-chain lipids (i.e., their transness in the gel state and the presence of multiple gauche bonds in the liquid-crystalline state).  相似文献   

3.
The activity of phospholipase A2 from cobra venom toward phospholipid in single-walled, sonicated vesicles was analyzed, particularly with respect to its activity toward the saturated phosphatidylcholines in the gel and liquid crystalline states. When egg phosphatidylcholine vesicles are used as substrate, the phospholipase has an apparent Km of 4.4 mM, an apparent Vmax of 100 mumol min-1 mg-1 of protein, and a pH optimum of 5.0 at 40 degrees C. The phospholipase hydrolyzed the gel state of dimyristoyl phosphatidylcholine vesicles and dipalmitoyl phosphatidylcholine vesicles at a rate 2 to 3 times greater than the liquid crystalline state, taking into account temperature effects on the enzymatic reaction itself. The results suggest that, toward sonicated vesicles, there is no specific enhancement of the rate when the both liquid crystalline and gel states are present together, as has been suggested to occur for multibilayers studied with other phospholipases. An apparent stimulation of activity as the reaction proceeded was observed above the phase transition temperature. This might be attributed to an increase in the phase transition temperature caused by free fatty acids so that, in the presence of reaction products, the enzyme is actually hydrolyzing gel state phospholipid which was found to be the preferred lipid state for phospholipase activity.  相似文献   

4.
K M Eum  G Riedy  K H Langley  M F Roberts 《Biochemistry》1989,28(20):8206-8213
Small unilamellar vesicles which form when gel-state long-chain phosphatidylcholines are mixed with micellar short-chain lecithins undergo an increase in size as the long-chain species melts to its liquid-crystalline form. Analysis of the vesicle population with quasi-elastic light scattering shows that the particle size increases from 90-A radius to greater than 5000-A radius. Resonance energy transfer experiments show total mixing of lipid probes with unlabeled vesicles only when the Tm of the long-chain phosphatidylcholine is exceeded. This implies that the large size change represents a fusion process. Aqueous compartments are also mixed during this transition. 31P NMR analysis of the vesicle mixtures above the phase transition shows a great degree of heterogeneity with large unilamellar particles coexisting with oligo- and multilamellar structures. Upon cooling the vesicles below the Tm, the original size distribution (e.g., small unilamellar vesicles) is obtained, as monitored by both quasi-elastic light scattering and 31P NMR spectroscopy. This temperature-induced fusion of unilamellar vesicles is concentration dependent and can be abolished at lower total phospholipid concentrations. It occurs over a wide range of long-chain to short-chain ratios and occurs with 1-palmitoyl-2-stearoylphosphatidylcholine and dimyristoylphosphatidylcholine as well. Characterization of this fusion event is used to understand the anomalous kinetics of water-soluble phospholipases toward these unusual vesicles.  相似文献   

5.
Apparent values of Km and Vmax have been measured for catalysis of hydrolysis of unsonicated egg lecithin liposomes, activated through addition of 0.4 M n-hexanol, by phospholipases A2 from bee and snake venoms and by phospholipase C from Clostridium welchii as a function of the concentration of three surfactants: hexadecylamine, hexadecyltrimethylammonium bromide, and dihexadecyl phosphate. For all three enzymes, values of Km and Vmax show little or no dependence on the concentration of these ionic surfactants, demonstrating that the liposomal surface charge is not a crucial factor in determining susceptibility to phospholipase-catalyzed hydrolysis.  相似文献   

6.
Several short-chain asymmetric lecithins with a total of 14 carbons in the acyl chains (ranging from 1-lauroyl-2-acetylphosphatidylcholine to 1-hexanoyl-2-octanoylphosphatidylcholine) have been synthesized and characterized. The specific activities of phospholipase A2 from cobra venom, phospholipase A2 from porcine pancreas, and phospholipase C from Bacillus cereus toward these lecithins as micelles have been determined. The results of these kinetic studies allow the definition of hydrophobic binding requirements in the active sites of these water-soluble phospholipases. For phospholipase C, with the exception of monomyristoylphosphatidylcholine, each of the asymmetric short-chain lecithins exhibits high activity, comparable to the 14-carbon symmetric short-chain species, diheptanoylphosphatidylcholine. Therefore, for phospholipase C, in addition to the acyl linkages, only a certain degree of hydrophobicity in the fatty acyl chains is requisite for substrate binding and appreciable hydrolysis; there is no chain specificity. The activity of phospholipase A2 from cobra venom toward the same asymmetric lecithins is quite different. As the sn-2 chain lengthens, activity is increased to a maximum for diheptanoyl-PC. Further increase in the number of carbons in the sn-2 chain has no effect on hydrolysis rates. For this enzyme, seven carbons in the sn-2 chain are necessary for optimal activity. In contrast, porcine pancreatic phospholipase A2 activity shows very little dependence on sn-2 chain length.  相似文献   

7.
J R Bian  M F Roberts 《Biochemistry》1990,29(34):7928-7935
Small bilayer particles form spontaneously from gel-state long-chain phospholipids such as dipalmitoylphosphatidylcholine and 0.2 mol fraction short-chain lecithins (e.g., diheptanoyl-phosphatidylcholine). When the particles are incubated at temperatures greater than the Tm of the long-chain phosphatidylcholine (PC), the particles rapidly fuse (from 90-A to greater than or equal to 5000-A radius); this transition is reversible. A possible explanation for this behavior involves patching or phase separation of the short-chain component within the gel-state particle and randomization of both lipid species above Tm. Differential scanning calorimetry, 1H T1 values of proteodiheptanoyl-PC in diheptanoyl-PC-d26/dipalmitoyl-PC-d62 matrices of varying deuterium content, solid-state 2H NMR spectroscopy as a function of temperature, and fluorescence pyrene excimer-to-monomer ratios as a function of mole fraction diheptanoyl-PC provide evidence that such phase separation must occur. These results are used to construct a phase diagram for the diheptanoyl-PC/dipalmitoyl-PC system, to propose detailed geometric models for the different lipid particles involved, and to understand phospholipase kinetics toward the different aggregates.  相似文献   

8.
A fluorescent assay for Clostridium perfringens phospholipase C is described using 1-palmitoyl-2-[6(pyren-1-yl)hexanoyl]-sn-glycero-3- phospho-N-(trinitrophenyl)aminoethanol (PPHTE) as the substrate. This method is based on the decrease of the quenching of pyrene monomer fluorescence when phospholipase C hydrolyzes PPHTE into pyrenediglyceride and phospho(trinitrophenyl)-aminoethanol. The hydrolysis of egg lecithin/PPHTE (25:1 molar ratio) substrate by C. perfringens phospholipase C was linear with time for at least 2 min. Optimal conditions for the hydrolysis by phospholipase C were 50 mM Tris-HCl pH 7.0-30 mM CaCl2/63 microM egg lecithin and 2.5 microM PPHTE. The Km and Vmax values for the hydrolysis of egg lecithin/PPHTE vesicles were 28 microM and 280 pmol min-1, respectively. The detection limit of the assay was 40 microU of C. perfringens phospholipase C. When diglyceride was included into egg lecithin/PPHTE vesicles up to 30 mol% the reaction velocity increased 13-fold. Higher molar proportions of diglyceride were inhibitory. When the hydrolysis of mixtures of different naturally occurring phospholipids and PPHTE was studied egg lecithin was found to be the best substrate. When dipalmitoylphospholipids with different polar head groups were used the reaction velocity decreased in the order egg lecithin greater than or equal to dipalmitoylphosphatidylserine greater than dipalmitoylphosphatidic acid greater than dipalmitoylphosphatidylcholine greater than dipalmitoylphosphatidylglycerol.  相似文献   

9.
The rates of the reaction products formation under simultaneous phospholipase D effect on phosphatidyl ethanolamine and phosphatidyl choline were studied. The hydrolysis of cephalin, unlike the phospholipase D effect on lecithin, does not require Ca2+ ions. Ca2+ does not affect the enzymatic degradation of lecithin and inhibits the reaction with cephalin in "inorganized" phospholipid emulsions. The hydrolysis of micellar phospholipids by phospholipase D (in the presence of the anionic detergent sodium dodecyl sulfate) is accelerated by Ca2+ ions for both substrates. The apparent Km value is equal to 1.5 mM and does not depend on the phospholipid type. In contrast, the value of kcat for lecithin is twice as high as that for cephalin. It was demonstrated that the phase state of the phospholipids and the chemical nature of the alcohol residue in the phospholipid molecule are essential for the substrate specificity of phospholipase D.  相似文献   

10.
R E Burrier  P Brecher 《Biochemistry》1984,23(22):5366-5371
Sonicated dispersions of egg yolk phosphatidylcholine and triolein as vesicles and microemulsions have been used as substrates for the assay of a purified acid lipase. Previous studies have also shown that triolein localized in the surface phase of emulsions is the preferred substrate. In this study, we examined enzyme activity following several surface modifications using both vesicles and microemulsions. When the acidic phospholipids phosphatidylserine and phosphatidic acid were incorporated into both vesicles and microemulsions at up to 10 mol % of the total phospholipid, a dose-dependent reduction in the apparent Km was observed. Using the vesicles as substrate, a dose-dependent decrease in Vmax was also observed. Agarose gel electrophoresis was used to verify suspected changes in net particle charge. Analogous inclusion of phosphatidylethanolamine, sphingomyelin, or cholesterol did not affect kinetic parameters. Addition of oleic acid to sonication mixtures produced vesicles with a decreased apparent Km and Vmax, but triolein hydrolysis in microemulsions was not significantly altered. Triolein-containing vesicles prepared by using dimyristoyl- or dipalmitoylphosphatidylcholine were hydrolyzed maximally at the gel liquid-crystalline transition temperatures of the appropriate phospholipid. Differential scanning calorimetry was used to verify the temperatures of transition in these vesicles. The results indicate that acid lipase activity is influenced by the charge or physical state of the surface phase of model substrates and suggest that degradation of core components of naturally occurring substrates such as lipoprotein may be influenced by chemical changes on the surface of these particles.  相似文献   

11.
1. The action of eight purified phospholipases on intact human erythrocytes has been investigated. Four enzymes, e.g. phospholipases A2 from pancreas and Crotalus adamanteus, phospholipase C from Bacillus cereus, and phospholipase D from cabbage produce neither haemolysis nor hydrolysis of phospholipids in intact cells. On the other hand, both phospholipases A2 from bee venom and Naja naja cause a non-haemolytic breakdown of more than 50% of the lecithin, while sphingomyelinase C from Staphylococcus aureus is able to produce a non-lytic degradation of more than 80% of the sphingomyelin. 2. Phospholipase C from Clostridium welchii appeared to be the only lipolytic enzyme tested, which produces haemolysis of human erythrocytes. Evidence is presented that the unique properties of the enzyme itself, rather than possible contaminations in the purified preparation, are responsible for the observed haemolytic effect. 3. With non-sealed ghosts, all phospholipases produce essentially complete breakdown of those phospholipids which can be considered as proper substrates for the enzymes involved. 4. Due to its absolute requirement for Ca2+, pancreatic phospholipase A2 can be trapped inside resealed ghosts in the presence of EDTA, without producing phospholipid breakdown during the resealing procedure. Subsequent addition of Ca2+ stimulates phospholipase A2 activity at the inside of the resealed cell, eventually leading to lysis. Before lysis occurs, however, 25% of the lecithin, half of the phosphatidylethanolamine and some 65% of the phosphatidylserine can be hydrolysed. This observation is explained in relation to an asymmetric phospholipid distribution in red cell membranes.  相似文献   

12.
This study describes the biochemical properties of the rat hepatic microsomal NADPH-specific short-chain enoyl CoA reductase and NAD(P)H-dependent long-chain enoyl CoA reductase. Of the substrates tested, crotonyl CoA and trans-2-hexenoyl CoA are reduced by the short-chain reductase only in the presence of NADPH. The trans-2-octenoyl CoA and trans-2-decenoyl CoA appear to undergo reduction to octanoate and decanoate, respectively, catalyzed by both enzymes; 64% conversion of the C8:1 is catalyzed by the short-chain reductase, while 36% conversion is catalyzed by the long-chain enzyme. For the C10:1 substrate, 45% is converted by the short-chain reductase, while 55% is reduced by the long-chain reductase. trans-2-Hexadecenoyl CoA is a substrate for the long-chain enoyl CoA reductase only. Reduction of C4 and C6 enoyl CoA's was unaffected by bovine serum albumin (BSA), whereas BSA markedly stimulated the conversion of C10 and C16 enoyl CoA's to their respective saturated product. Reduction rates as a function of microsomal protein concentration, incubation time, pH, and cofactors are reported including the apparent Km and Vmax for substrates and cofactors. In general, the apparent Km's for the substrates ranged from 19 to 125 microM. The apparent Vmax for the short-chain enoyl CoA reductase was greatest with trans-2-hexenoyl CoA, having a turnover of 65 nmol/min/mg microsomal protein, while the apparent Vmax for the long-chain enzyme was greatest with trans-2-hexadecenoyl CoA, having a turnover of 55 nmol/min/mg microsomal protein. With respect to electron input, NADPH-cytochrome P-450 reductase, either alone, mixed with phospholipid, or incorporated into phospholipid vesicles, possessed no enoyl CoA reductase activity. Cytochrome c did not affect the NADPH-dependent conversion of the trans-2-enoyl CoA. In addition, anti-NADPH-cytochrome P-450 reductase IgG did not inhibit the reduction of trans-2-hexadecenoyl CoA in hepatic microsomes. Finally, the NADPH-specific short-chain and NAD(P)H-dependent long-chain enoyl CoA reductases were solubilized and completely separated from NADPH-cytochrome P-450 reductase by employing DE-52 column chromatography. These studies demonstrate the noninvolvement of NADPH-cytochrome P-450 reductase in either the short-chain (13) or long-chain enoyl CoA reductase system. Thus, the role of NADPH-cytochrome P-450 reductase in the microsomal elongation of fatty acids appears to be at the level of the first reduction step.  相似文献   

13.
W Yuan  D M Quinn  P B Sigler  M H Gelb 《Biochemistry》1990,29(25):6082-6094
The action of the phospholipases A2 (PLA2s) from Naja naja naja, Naja naja atra, and Crotalus atrox venoms as well as the enzyme from porcine pancreas on a number of short-chain, water-soluble substrates was studied. The inhibition of these enzymes by short-chain phosphonate- and thiophosphonate-containing phospholipid analogues was also examined. The kinetic patterns observed for the action of the venom PLA2s on substrates containing phosphocholine head groups all deviated from a classical Michaelis-Menten-type behavior. With a substrate containing an anionic head group, the kinetic pattern observed was more normal. In contrast, Michaelis-Menten-type behavior was observed for the action of the porcine pancreatic PLA2 acting on all of the substrates studied. A short-chain phospholipid analogue in which the enzyme-susceptible ester was replaced with a phosphonate group was found to be a tight-binding inhibitor of the venom PLA2s with IC50 values that were some 10(4)-10(5)-fold lower than the concentration of substrate used in the assay. The degree of inhibition was found to depend dramatically on the stereochemical arrangement of substituents in the inhibitor which strongly suggests that the inhibitors are binding directly to the active site of the PLA2s. By comparison, the phosphonate analogue functioned as a poor inhibitor of the porcine pancreatic PLA2. Direct inhibitor binding studies indicated that the short-chain phosphonate inhibitor bound weakly to the venom enzymes in the absence of the short-chain substrates. Several other unusual features of the inhibition were also observed. The data are interpreted in terms of a model in which the enzyme and substrate form a lipid-protein aggregate at substrate concentrations below the critical micelle concentration (cmc). Possible reasons for the selective binding of the inhibitor to the enzyme-substrate microaggregate are discussed.  相似文献   

14.
The kinetics of the Ca2+-dependent, alkaline pH optimum, membrane-bound phospholipase A2 from the P388D1 macrophage-like cell line were studied using various phosphatidylcholine (PC) and phosphatidylethanolamine (PE) substrates. This enzyme exhibits "surface dilution kinetics" toward PC in Triton X-100 mixed micelles, and the "dual phospholipid model" was found to adequately describe its kinetic behavior. With substrate in the form of sonicated vesicles, the dual phospholipid model should give rise to Michaelis-Menten type kinetics. However, the hydrolysis of dipalmitoyl-PC, 1-palmitoyl-2-oleoyl-PC, and 1-stearoyl-2-arachidonoyl-PC vesicles exhibited two distinct activities. Below 10 microM, the data appeared to follow Michaelis-Menten behavior, while at higher concentrations, the data could best be fit to a Hill equation with a Hill coefficient of 2. These PCs had Vmax values for the low substrate concentration range of 0.2-0.6 nmol min-1 mg-1 and Km values of 1-2 microM. At the high substrate concentration range, the Vmax values were between 5 and 7 nmol min-1 mg-1. PC containing unsaturated fatty acids had an apparent Km, determined from the Hill equation, of about 15 microM, while the apparent Km of dipalmitoyl-PC was 0.6 microM. When 70% glycerol was included in the assays, a single Michaelis-Menten curve was obtained for both dipalmitoyl-PC and 1-stearoyl,2-arachidonoyl-PC. Possible explanations for these kinetic results include reconstitution of the membrane-bound phospholipase A2 in the phospholipid vesicle or the enzyme has tow distinct phospholipid binding function. The kinetics for both dipalmitoyl-PC and dipalmitoyl-PE hydrolysis in vesicles was very similar, indicating that the enzyme does not greatly prefer one of these head groups over the other. The enzyme also showed no preference for arachidonoyl containing phospholipid. Enzymatic activity toward PC containing saturated fatty acids was linear to about 15% hydrolysis while the hydrolysis of PC containing unsaturated fatty acids was linear to only about 5%. This loss of linearity was due to inhibition by released unsaturated fatty acids. Arachidonic acid was found to be a competitive inhibitor of dipalmitoyl PC hydrolysis with a K1 of 5 microM. This tight binding suggests a possible in vivo regulatory role for arachidonic acid. Three compounds of the arachidonic acid cascade, prostaglandin F2 alpha, 6-keto-prostaglandin F1 alpha, and thromboxane B2, showed no inhibition of enzymatic activity.  相似文献   

15.
On the substrate specificity of rat liver phospholipase A1   总被引:1,自引:0,他引:1  
The substrate specificity of purified phospholipase A1 was studied using mixed micelles of phospholipid and Triton X-100. The kinetic analysis employed determined Vmax, Ks (a dissociation constant for the phospholipase A1-mixed micelle complex), and Km (the Michaelis constant for the catalytic step which reflects the binding of the enzyme to the substrate in the interface). The order of Vmax values was phosphatidic acid greater than phosphatidylethanolamine greater than phosphatidylcholine greater than phosphatidylserine. The order of Ks values was phosphatidylcholine greater than phosphatidylethanolamine greater than phosphatidic acid greater than phosphatidylserine; the order of Km values was phosphatidic acid greater than phosphatidylethanolamine = phosphatidylserine greater than phosphatidylcholine. When present together, phosphatidylcholine inhibited the hydrolysis of phosphatidylethanolamine but phosphatidylethanolamine did not affect the hydrolysis of phosphatidylcholine. Sphingomyelin, phosphatidylcholine plasmalogen, and phosphatidylethanolamine plasmalogen had no effect on the hydrolysis of phosphatidylethanolamine. The effects of the reaction products, lysolipids and/or fatty acids, were also considered for their influence on phosphatidylethanolamine hydrolysis catalyzed by phospholipase A1. Free fatty acid was found to inhibit, whereas lysophospholipids stimulated hydrolysis of phosphatidylethanolamine. In a mixture of 1,2- and 1,3-diacylglycerides in mixed micelles, only the acyl chain at the sn-1 position of the 1,2 compound was hydrolyzed. Surface charge did not modulate the hydrolysis of phosphatidylcholine vesicles or mixed micelles. In conclusion, it is hypothesized that steric hindrance at position 3 of the glycerol regulates substrate binding in the active site and that an acyl group in position 1 is favored over a vinyl ether linkage for binding.  相似文献   

16.
The release of free arachidonic acid from membrane phospholipids is believed to be the rate-controlling step in the production of the prostaglandins, leukotrienes, and related metabolites in inflammatory cells such as the macrophage. We have previously identified several different phospholipases in the macrophage-like cell line P388D1 potentially capable of controlling arachidonic acid release. Among them, a membrane-bound, alkaline pH optimum, Ca2+-dependent phospholipase A2 is of particular interest because of the likelihood that the regulatory enzyme has these properties. This phospholipase A2 has now been solubilized from the membrane fraction with octyl glucoside and partially purified. The first two steps in this purification are butanol extractions that yield a lyophilized, stable preparation of phospholipase A2 lacking other phospholipase activities. This phospholipase A2 shows considerably more activity when assayed in the presence of glycerol, regardless of whether the substrate, dipalmitoylphosphatidylcholine, is in the form of sonicated vesicles or mixed micelles with the nonionic surfactant Triton X-100. Glycerol (70%) increases both the Vmax and the Km with both substrate forms, giving a Vmax of about 15 nmol min-1 mg-1 and an apparent Km of about 60 microM for vesicles and a Vmax of about 100 nmol min-1 mg-1 and an apparent Km of about 1 mM for mixed micelles. Vmax/Km is slightly greater for vesicles than for mixed micelles. The lyophilized preparation of the enzyme is routinely purified about 60-fold and is suitable for evaluating phospholipase A2 inhibitors such as manoalide analogues. Subsequent steps in the purification are acetonitrile extraction followed by high performance liquid chromatography on an Aquapore BU-300 column and a Superose 12 column. This yields a 2500-fold purification of the membrane-bound phospholipase A2 with a 25% recovery and a specific activity of about 800 nmol min-1 mg-1 toward 100 microM dipalmitoylphosphatidylcholine in mixed micelles. When this material was subjected to analysis on a Superose 12 sizing column, the molecular mass of the active fraction was approximately 18,000 daltons.  相似文献   

17.
The mechanism of action of hepatic triacylglycerol lipase (EC 3.1.1.3) was examined by comparing the hydrolysis of a water-soluble substrate, tributyrin, with that of triolein by hepatic triacylglycerol lipase purified from human post-heparin plasma. The hydrolyzing activities toward tributyrin and triolein were coeluted from heparin-Sepharose at an NaCl concentration of 0.7 M. The maximal velocity of hepatic triacylglycerol lipase (Vmax) for tributyrin was 17.9 mumol/mg protein per h and the Michaelis constant (Km) value was 0.12 mM, whereas the Vmax for triolein was 76 mumol/mg per h and the Km value was 2.5 mM. The hydrolyses of tributyrin and triolein by hepatic triacylglycerol lipase were inhibited to similar extends by procainamide, NaF, Zn2+, Cu2+, Mn2+, SDS and sodium deoxycholate. Triolein hydrolysis was inhibited by the addition of tributyrin. Triolein hydrolysis was also inhibited by the addition of dipalmitoylphosphaidylcholine vesicles. In contrast, the additions of triolein emulsified with Triton X-100 and dipalmitoylphosphatidylcholine vesicles enhanced the rate of tributyrin hydrolysis by hepatic triacylglycerol lipase. In the presence of dipalmitoylphosphatidylcholine, the Vmax and Km values of hepatic triacylglycerol lipase for tributyrin were 41 mumol/mg protein per h and 0.12 mM, respectively, indicating that the enhancement of hepatic triacylglycerol lipase activity for tributyrin by dipalmitoylphosphatidycholine vesicles was mainly due to increase in the Vmax. The enhancement of hepatic triacylglycerol lipase activity for tributyrin by phospholipid was not correlated with the amount of tributyrin associated with the phospholipid vesicles. On Bio-Gel A5m column chromatography, glycerol tri[1-14C]butyrate was not coeluted with triolein emulsion, and hepatic triacylglycerol lipase activity was associated with triolein emulsion even in the presence of 2 mM tributyrin. These results suggest that hepatic triacylglycerol lipase has a catalytic site for esterase activity and a separate site for lipid interface recognition, and that on binding to a lipid interface the conformation of the enzyme changes, resulting in enhancement of the esterase activity.  相似文献   

18.
We have investigated the action and substrate specificity of phospholipase A2 (EC 3.1.1.4) purified from cobra venom (Naja naja naja) toward intact and Triton-solubilized human erythrocytes, toward ghost membranes, and toward extracted ghost lipids in mixed micelles with Triton X-100. We have found that: (i) phospholipids in the outer surface of intact erythrocytes are extremely poor substrates for the phospholipase, (ii) phospholipids in ghost erythrocyte membranes and in Triton-solubilized erythrocytes are suitable substrates for the enzyme, (iii) in these latter systems which contain a mixture of lipids, phosphatidylethanolamine is preferentially hydrolyzed, whereas in model studies on individual phospholipid species in mixed micelles with Triton, phosphatidylcholine is the preferred substrate of the enzyme, and (iv) the preferential hydrolysis of phosphatidylethanolamine is also observed for extracted ghost lipid mixtures in mixed micelles. These results demonstrate a dependence of phospholipase A2 activity on the ghosting procedure and a dependence of substrate specificity on the presence of other lipids. The relevance of these findings to the interpretation of membrane lipid asymmetry studies utilizing phospholipases is considered in detail.  相似文献   

19.
Radiolabeled, autoclaved yeast were tested as a substrate for mammalian phospholipase A2 activity because the only other membranous substrate used for this purpose, autoclaved Escherichia coli, totally lacks a major mammalian phospholipid, phosphatidylcholine. Candida albicans were grown in the presence of [1-14C]oleate and then autoclaved. Sixty three percent of the incorporated label was in yeast phospholipid, and more than 95% of that was in the 2-acyl position. The distribution of label in the yeast phospholipids (phosphatidylcholine and -ethanolamine, -serine + -inositol, and phosphatidic acid corresponded closely to the chemical distribution of phosphorus in those phospholipids. Snake venom (Naja naja) and human synovial fluid phospholipase A2 hydrolyzed yeast phospholipid exclusively to release 14C-labeled fatty acid. When 50-60% of the yeast phospholipid was hydrolyzed, the radioactive fatty acids as determined by gas-liquid chromatographic analysis were predominantly oleate (45%) and linoleate (greater than 54%). Hydrolysis of yeast phospholipid by both enzymes was near-linear with protein and time under conditions of optimal pH (neutral-alkaline) and Ca2- (1-5 mM) previously reported for optimal hydrolysis of autoclaved E. coli phospholipid. N. naja phospholipase A2 showed less preference for phosphatidylethanolamine than -choline as liposomes or yeast phospholipid as compared to human synovial fluid phospholipase A2 which clearly preferred phosphatidylethanolamine to -choline as a liposome or yeast phospholipid. These results illustrate that radiolabeled phospholipids of autoclaved yeast, enriched in phosphatidylcholine, are readily hydrolyzed by snake venom and human nonpancreatic phospholipases A2 and may, therefore, be useful in the measurement of in vitro enzymatic activity.  相似文献   

20.
We examined the action of porcine pancreatic and bee-venom phospholipase A2 towards bilayers of phosphatidylcholine as a function of several physical characteristics of the lipid-water interface. 1. Unsonicated liposomes of dimyristoyl phosphatidylcholine are degraded by both phospholipases in the temperature region of the phase transition only (cf. Op den Kamp et al. (1974) Biochim. Biophys. Acta 345, 253--256 and Op den Kamp et al. (1975) Biochim. Biophys. Acta 406, 169--177). With sonicates the temperature range in which hydrolysis occurs is much wider. This discrepancy between liposomes and sonicates cannot be ascribed entirely to differences in available substrate surface. 2. Below the phase-transition temperature the phospholipases degrade dimyristoyl phosphatidylcholine single-bilayer vesicles with a strongly curved surface much more effectively than larger single-bilayer vesicles with a relatively low degree of curvature. 3. Vesicles composed of egg phosphatidylcholine can be degraded by pancreatic phospholipase A2 at 37 degrees C, provided that the substrate bilayer is strongly curved. The bee-venom enzyme shows a similar, but less pronounced, preference for small substrate vesicles. 4. In a limited temperature region just above the transition temperature of the substrate the action of both phospholipases initially proceeds with a gradually increasing velocity. This stimulation is presumably due to an increase of the transition temperature, effectuated by the products of the phospholipase action. 5. Structural defects in the substrate bilayer, introduced by sonication below the phase-transition temperature (cf. Lawaczeck et al. (1976) Biochim. Biophys. Acta 443, 313--330) facilitate the action of both phospholipases. The results lead to the general conclusion that structural irregularities in the packing of the substrate molecules facilitate the action of phospholipases A2 on phosphatidylcholine bilayers. Within the phase transition and with bilayers containing structural defects these irregularities represent boundaries between separate lipid domains. The stimulatory effect of strong bilayer curvature can be ascribed to an overall perturbation of the lipid packing as well as to a change in the phase-transition temperature.  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