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 62 毫秒
1.
Fanani ML  Topham MK  Walsh JP  Epand RM 《Biochemistry》2004,43(46):14767-14777
Diacylglycerol kinase (DGK) isoforms alpha and zeta were extracted from transfected cells that overexpressed these enzymes. We determined the lipid dependence of the binding of these isoforms to liposomes. The modulation by lipid of the rate of phosphorylation of diacylglycerol by these enzymes was also measured. Incorporation of phosphatidylethanolamine into the liposomes resulted in an increased partitioning of both isoforms of DGK to the membrane as well as an increased catalytic rate. We demonstrate that the increased catalytic rate is a consequence of both increased portioning of the enzyme to the membrane and increased catalytic activity of the membrane-bound form. DGKalpha, a calcium-dependent isoform, can be activated in a calcium-independent fashion in the presence of phosphatidylethanolamine. Similar effects are observed with cholesterol. In contrast, sphingomyelin inhibits the activity of both isoforms of DGK. Our results demonstrate that the translocation to membranes and activity of DGKalpha and DGKzeta are modulated by the composition and properties of the membrane. The enzymes are activated by the presence of lipids that promote the formation of inverted phases. However, the promotion of negative curvature is not the sole factor contributing to the lipid effects on enzyme binding and activity. A truncated form of DGKalphalacking both the E-F hand and the recoverin homology domain is constitutively active and is not further activated by any of the lipids tested or by calcium. However, a truncated form lacking only the recoverin homology domain is partially activated by either calcium or certain lipids.  相似文献   

2.
The lipid cofactor requirement of Escherichia coli sn-1,2-diacylglycerol kinase was studied using a beta-octylglucoside mixed micellar assay (Walsh, J. P., and Bell, R. M. (1986) J. Biol. Chem. 261, 6239-6247). The enzyme was shown to have an absolute requirement for a lipid activator. sn-1,2-Dioleoylglycerol was both an activator and a substrate for the enzyme, 1,3-dioleoylglycerol was an activator but not a substrate, and sn-1,2-dioctanoylglycerol was a substrate but not an activator. Activation was observed with a large number of phospholipids, sulfolipids, neutral lipids, and detergents. Lipids with longer alkyl/acyl chains stimulated activity to a greater extent and at lower concentrations than their shorter chain homologs. Anionic lipids were the best activators, and neutral lipids were somewhat less effective. Cationic lipids were poor activators. Lipid activation was cooperative in all cases, with Hill coefficients ranging from 2.9 to 4.7. Lipid activators stabilized the enzyme against inactivation induced by diacylglycerols. The effectiveness of several lipids in stabilizing the enzyme correlated with their effectiveness as kinetic activators, suggesting a common mechanism. Kinetic analyses also suggested that a lipid cofactor-induced conformational change occurs as a part of the activation process. beta-Octylglucoside was shown not to function as a lipid cofactor for diacylglycerol kinase. The requirement for detergent in the assay was related, instead, to the need to disperse and deliver water-insoluble substrates and cofactors to the enzyme. beta-Octylglucoside also provided an inert matrix to which lipid substrates and cofactors could be added, enabling study of their concentration dependencies.  相似文献   

3.
Enzyme preparations with variable phospholipid contents were obtained by removing lipids from sarcoplasmic reticulum with deoxycholate. Preparations containing from 90 to 37 phospholipids per enzyme showed normal values of both Ca2+-ATPase activity and steady-state phosphoenzyme levels. Fractions containing 37 to 23 phospholipids per enzyme had a reduced ATPase activity but normal phosphoenzyme levels, showing that in this range of lipid content the ATPase reaction is inhibited in a reaction step subsequent to phosphoenzyme formation but prior to phosphoenzyme decomposition. Delipidation below 23 lipids per enzyme caused a marked reduction of the amount of phosphoenzyme formed, so that although both reactions require lipids, fewer lipids are required for phosphoenzyme formation than for decomposition. The effect of lipid removal could be completely reversed by readdition of lipids to fractions containing more than 11 lipids per enzyme. It is proposed that phosphoenzyme formation requires full occupancy of a boundary domain of 23 lipids per enzyme, and that the selective inhibition of phosphoenzyme decomposition at higher lipid contents is caused by a decrease in the rotational mobility of the enzyme.  相似文献   

4.
Lin Q  Higgs HN  Glomset JA 《Biochemistry》2000,39(31):9335-9344
We previously purified a cytosolic phospholipase A1 that could catalyze the preferential hydrolysis of phosphatidic acid in mixed-micelle assays. Here we studied the enzyme's interactions with unilamellar lipid membranes and examined effects of the lipids on enzyme binding, stability, and catalysis. A major finding was that membrane lipids could influence the stability, activity, and specificity of the enzyme under conditions where enzyme binding to the membranes was likely to be saturated. Thus, the enzyme was unstable at 37 degrees C in the absence of membranes but bound to membranes that contained anionic phosphoglycerides and could be stabilized by these membranes in the presence of albumin. The overall activity of the bound enzyme toward membrane phosphoglycerides, assayed in the presence of albumin, increased when phosphatidylethanolamine was substituted for phosphatidylcholine. Furthermore, the enzyme's catalytic preference for phosphatidic acid increased when cholesterol and diacylglycerol were included in the membranes, sn-1-stearoyl-2-arachidonoylphosphatidylethanolamine was substituted for sn-1-palmitoyl-2-oleoylphosphatidylethanolamine, and the concentration of phosphatidic acid was increased from 0 to 10 mol % of the total membrane phosphoglycerides. Finally, changes in the relative contents of phosphatidylcholine and phosphatidylserine in the membranes influenced the enzyme's catalytic preference for different molecular species of phosphatidic acid. These results provide the first available information about the enzyme's ability to interact with membranes and identify conditions that yield high enzyme activity toward membrane-associated phosphatidic acid.  相似文献   

5.
The effects of fatty acids and phospholipids on the interaction of the full-length GTPase activating protein (GAP) as well as its isolated C-terminal domain and the Ha-ras proto-oncogene product p21 were studied by various methods, viz. GTPase activity measurements, fluorescence titrations and gel permeation chromatography. It is shown that all fatty acids and acidic phospholipids tested, provided the critical micellar concentration and the critical micellar temperature are reached, inhibit the GAP stimulated p21 GTPase activity. This is interpreted to mean that it is not the molecular structure of acidic lipid molecules per se but rather their physical state of aggregation which is responsible for the inhibitory effect of lipids on the GTPase activity. The relative inhibitory potency of various lipids was measured under defined conditions with mixed Triton X-100 micelles to follow the order: unsaturated fatty acids greater than saturated acids approximately phosphatidic acids greater than or equal to phosphatidylinositol phosphates much greater than phosphatidylinositol and phosphatidylserine. GTPase experiments with varying concentrations of p21 and constant concentrations of GAP and lipids indicate that the binding of GAP by the lipid micelles is responsible for the inhibition, a finding which was confirmed by fluorescence titrations and gel filtrations which show that the C-terminal domain of GAP is bound by lipid micelles.  相似文献   

6.
The CTP:phosphocholine cytidylyltransferase (CCT) governs the rate of phosphatidylcholine (PtdCho) biosynthesis, and its activity is governed by interaction with membrane lipids. The carboxy-terminus was dissected to delineate the minimum sequences required for lipid responsiveness. The helical domain is recognized as a site of lipid interaction, and all three tandem alpha-helical repeats from residues 257 through 290 were found to be required for regulation of enzymatic activity by this domain. Truncation of the carboxy-terminus to remove one or more of the alpha-helical repeats yielded catalytically compromised proteins that were not responsive to lipids but retained sufficient activity to accelerate PtdCho biosynthesis when overexpressed in vivo. The role of the helical region in lipid-activation was tested further by excising residues 257 through 309 to yield a protein that retained a 57-residue carboxy terminal domain fused to the catalytic core. This construct tested the hypothesis that the helical region inhibits activity in the absence of lipid rather than activates the enzyme in the presence of lipid. This hypothesis predicts constitutive activity for CCTalpha[Delta257-309]; however, this protein was tightly regulated by lipid with activities comparable to the full-length CCTalpha, in both the absence and presence of lipid. Activation of CCTalpha[Delta257-309] was dependent exclusively on anionic lipids, whereas full-length CCTalpha responded to either anionic or neutral lipids. Phosphatidic acid delivered in Triton X-100 micelles was the preferred activator of the second lipid-activation domain. These data demonstrate that CCTalpha can be regulated by lipids by two independent domains: (i) the three amphipathic alpha-helical repeats that interact with both neutral and anionic lipid mixtures and (ii) the last 57 residues that interact with anionic lipids. The results show that both domains are inhibitory in the absence of lipid and activating in the presence of lipid. Removal of both domains results in a nonresponsive, dysregulated enzyme with reduced activity. The data also demonstrate for the first time that the 57-residue carboxy-terminal domain in CCTalpha participates in lipid-mediated regulation and is sufficient for maximum activation of enzyme activity.  相似文献   

7.
Transfer of alkaline phosphatase (AP) directly from Escherichia coli cells into reverse micellar solutions (RMS) was studied by varying the water content, pH, and ionic strength of RMS. Prior to the mass transfer studies, the optimum conditions for the activity of alkaline phosphatase in reverse micellar solutions were determined. The maximum enzyme activity could be detected at higher pH and water content, indicating the ionization of p-nitrophenol to be crucial in the enzyme activity. The transfer of AP from E. coli followed a two-step process with most of the recovery being achieved in the first 3-10 min beyond which it slowed. A model suggesting separate locations for the enzyme in the cell wall has been proposed.  相似文献   

8.
增强UV-B辐射对麻花艽叶片的抗氧化酶的影响   总被引:12,自引:0,他引:12  
以青藏高原的特有植物麻花艽为材料,研究麻花艽叶片抗氧化酶系统对增强UV-B辐射的响应。结果表明:在UV-B处理初期,麻花艽叶片SOD、POD的酶活性都能增加,但随着处理时间的延长,SOD、POD的活性呈现下降趋势。麻花艽叶片CAT的酶活性在UVB处理后下降明显,但作为清除叶绿体中H2O2的关键酶AP的酶活性表现为明显地增加趋势,说明在对麻花艽叶片增强UV-B辐射反应中AP起有着重要作用。MDA的含量随UV-B处理时间延长而增加,表明UV-B降低了细胞内活性氧自由基的清除能力,膜脂过氧化作用加剧,导致了对麻花艽叶片的伤害效应。  相似文献   

9.
A comparative study of the catalytic activity of alpha-chymotrypsin and the spin label rotation frequency in the alpha-chymotrypsin active center of reverse micellar systems solvated by H2O-organic mixtures was carried out. It was found that the decrease in the label rotation frequency resulting from the substitution of water in the micellar inner cavity by glycerol, 2.3-butanediol and dimethylsulfoxide (up to 95%) caused a marked increase (20-fold in the case of 2.3-butanediol) of the enzyme catalytic activity. The experimental results are quantitatively interpreted in terms of a simple kinetic scheme postulating the existence of the enzyme in two interconvertible forms differing in the conformational (intramolecular) mobility, i.e., the resting one predominantly existing in aqueous solution, and the tense one whose proportion rises with an increase in the concentration of the water-miscible organic solvent in the reverse micellar system. The value of kcat (2.4 s-1) for the tense form of the enzyme exceeded by more than 25 times the catalytic activity of chymotrypsin in aqueous solution (0.09 s-1) for the resting form.  相似文献   

10.
The influence of micelle hydration degree (w0) and AOT concentration on fluorescence, circular dichroism (CD), catalytic activity, and stability of catalase in Aerosol OT (AOT) reversed micelles in heptane was investigated. The quantitative parameters--differential fluorescence of catalase (DeltaI), protein molar ellipticity ([theta]lambda), initial rate of catalytic reaction, catalase efficiency (kcat/Km), and rate constant of enzyme inactivation (kin, sec-1)--decreased with increasing AOT concentration in micellar systems, reflecting the interaction of solubilized catalase with the AOT micellar aggregates in heptane. The dependences of all these parameters on increasing hydration degree of micelles (w0) were characterized by the appearance of maxima at w0 of 8, 15-18, and 26-30. These maxima are suggested to reflect three different states of catalase in the micellar system, distinguished by their conformations and catalytic activity, which is determined by the micellar microenvironment of the enzyme.  相似文献   

11.
The effects of various lipids on calmodulin interaction with Ca-dependent phosphodiesterase were investigated. Palmitic, myristic and stearic acids increased the enzyme activity; the degree of the enzyme activation by calmodulin was decreased thereby. Oleic acid produced a weak activating effect on phosphodiesterase but completely blocked calmodulin action. The effects of the fatty acids under study were reversible, the activation constant was equal to 10(-4)-5 X 10(-4) M. In the presence of Ca2+ phosphoinositides and fatty acids changed the fluorescence intensity of dansyl-labelled calmodulin; in the absence of Ca2+ the lipids did not affect protein fluorescence. The lipids had no influence on the protein affinity for Ca2+. During chromatography of phosphodiesterase on calmodulin-Sepharose the enzyme was eluted from the column both in the presence of EGTA and palmitic acid. It was concluded that fatty acids prevent the formation of the calmodulin - phosphodiesterase complex. This effects may both be due to the lipid binding to the enzyme and to calmodulin.  相似文献   

12.
6-Phosphofructo-1-kinase (PFK-1), a major regulatory enzyme in the glycolysis pathway, is a cytoplasmic enzyme with complicated allosteric kinetics. Here we investigate the effects of lipids on the activity of PFK from Bacillus stearothermophilus (BsPFK), to determine whether BsPFK shares any of the membrane binding or lipid binding properties reported for some mammalian PFKs. Our results show that large unilamellar vesicles (LUVs) composed of either the phospholipid 1,2-dioleoyl-sn-glycero-3-phosphocholine (DOPC) or of 1:1 (mole ratio) DOPC and the fatty acid, oleic acid (OA), cause a three-fold increase in Vmax, depending on the lipid concentration and vesicle composition, but no change in Km. Further studies show lipids do not reverse the allosteric inhibitory effects of phosphoenolpyruvate (PEP) on BsPFK. SDS/PAGE studies do not show significant binding of the BsPFK tetramer to the surface of the phospholipid vesicles, suggesting that modulation of catalytic activity is due to binding of lipid monomers. By simulating the kinetics of BsPFK interaction with vesicles and lipid monomers we conclude that the change in BsPFK catalytic activity with respect to lipid concentration is consistent with monomer abstraction from vesicles rather than direct uptake of lipid monomers from solution.  相似文献   

13.
Bovine thyroid peroxidase (TPO), an enzyme requiring lipids for demonstrating catalytic activity, was incorporated in liposomes made of pure phospholipids. The enzyme did not show high differences in activity when bilayer thickness was changed, but dipalmitoyl phosphatidyl choline (DPPC) seemed to be more appropiate for activity. The perturbation caused on lipid fluidity by enzyme incorporation was studied by differential scanning calorimetry (DSC) and fluorescence polarization of the apolar probe 1,6-diphenyl-1,3,5-hexatriene (DPH). The complexes of TPO with dimyristoyl phosphatidyl choline (DMPC), DPPC, and distearoyl phosphatidyl choline (DSPC) bilayers showed transition temperatures (Tc) which were lower than the characteristic ones shown by liposomes with the respective phospholipids alone. The microsomal fraction from which TPO was extracted was in the fluid state at 37°C, the temperature at which thyroid peroxidase works ‘in vivo’. Since the effect of the protein in lowering the transition temperature of the phospholipids was so low, the contribution of phospholipids containing unsaturated fatty acids has to be essential for obtaining a fluid bilayer at body temperature.  相似文献   

14.
The misfolding and self-assembly of proteins into amyloid fibrils, which occur in several debilitating and age-related diseases, are affected by common components of amyloid deposits, notably lipids and lipid complexes. Previously, the effects of phospholipids on amyloid fibril formation by apolipoprotein (apo) C-II have been examined, where low concentrations of micellar phospholipids and lipid bilayers induce a new, straight rod-like morphology for apoC-II fibrils. This fibril appearance is distinct from the twisted-ribbon morphology observed when apoC-II fibrils are formed in the absence of lipids. We used total internal reflection fluorescence microscopy (TIRFM) to visualize the described polymorphism of apoC-II amyloid fibrils. The spontaneous assembly of apoC-II into either twisted-ribbon fibrils in the absence of lipids or into fibrils of straight rod-like morphology when lipids are present was captured by TIRFM. The latter was found to be better suited for visualization using TIRFM. The difference between seeding of apoC-II straight fibrils on microscopic quartz slide and in test tube suggested a role for the effects of incubation surface on fibril formation. Seed-dependent growth of apoC-II straight fibrils was probed further by using a dual-labelling construct, giving insights into the straight fibril growth pattern.  相似文献   

15.
The results of long-time research of antioxidative activity (AOA) of lipids from tissues of different species and lines of laboratory rodents are generalized. The classification of lipids according to their ability to inhibit the thermal autooxidation of methyl oleate is proposed. The participation of lipids in low-temperature autooxidation reactions at the initiation and chain propagation stages was proved by means of the model proposed. In addition to the lipid antioxidative activity, the initial quantity of peroxides in lipids due to the extent of their unsaturation and lipid antiperoxide activity are proposed for the estimation of the kinetic characteristics of lipids. The dependence of effects on the rate of radical initiation in the system is shown to be caused by the influence of physicochemical properties of lipids on the interrelation coordination and balance of biochemical functions in biological objects differing in the intensity of oxidation processes.  相似文献   

16.
Several seven-carbon fatty acyl lecithins with varied acyl chain branching have been synthesized and characterized as potential phospholipase A2 substrates. Micellar bis(4,4-dimethylpentanoyl) phosphatidylcholine, bis(5-methylhexanoyl)phosphatidylcholine, bis(3-methylhexanoyl)phosphatidylcholine, and bis(2-methylhexanoyl)phosphatidylcholine are poor substrates for phospholipase A2 (Naja naja naja). These branched lecithins also inhibit the hydrolysis of diheptanoylphosphatidylcholine by the enzyme with Ki values comparable to or smaller than the apparent Km of the linear compound. The terminally branched lecithins are excellent substrates for another surface-active hydrolytic enzyme, phospholipase C from Bacillus cereus. When only one acyl chain bears a methyl group, the hybrid lecithins 1-heptanoyl-2-(2-methylhexanoyl)phosphatidylcholine and 1-(3-methylhexanoyl)-2-heptanoylphosphatidylcholine are substrates comparable to diheptanoylphosphatidylcholine. Analysis of micellar structure and dynamics by 1H and 13C NMR spectroscopy, quasi-elastic light scattering, and comparison of critical micellar concentrations indicates little significant difference in the conformation and dynamics of these seven-carbon fatty acyl lecithin micelles, even when the methyl groups are adjacent to the carbonyls. Phospholipase A2 UV difference spectra induced by phospholipid binding imply different enzyme conformations or aggregation states caused by linear-chain and asymmetric-chain lipids compared to bis(methylhexanoyl)phosphatidylcholines. The differences in hydrolytic activity of phospholipase A2 against the branched-chain micellar lecithins can then be attributed to an enzyme-lipid interaction at the active site. The species with both fatty acyl chains branched bind to phospholipase A2 but are not turned over rapidly. Since poor enzymatic activity only occurs for lecithins with both chains methylated, the interaction of both chains with the enzyme must be important for catalytic efficiency.  相似文献   

17.
1. A direct method for determining the binding of triated water to lipids is described. The experimental conditions were practically identical to those previously employed (1974) in the determination of the cofactor activities of a series of oleyl-lipids in reactivation of the C55-isoprenoid alcohol phosphokinase apoprotein. 2. Active cofactor lipids (dioleyl lecithin, sodium oleate, 1-monoolein, 1-monomyristin)bound between 2.3 and 5.3 nmol 3H2O per nmol lipid, whereas less than 0.14 nmol 3H2O were bound per nmol of the inactive lipids (1,2- and 1,3-diolein, triolein, oleyl alcohol, methyl oleate, cholesteryl oleate). 3. When exposed to 3H2O vapour, the active lipids adsorbed between 1 and 2 nmol 3H2O per nmol lipid, whereas the inactive lipids adsorbed less than 0.1 nmol 3H2O per nmol lipid. 4. The active lipid cofactor, egg lecithin, bound more than twice as much 3H2O as egg phosphatidylethanolamine which was devoid of cofactor activity in the absence of detergent. 5. Appropriately hydrated lipid polar groups are concluded to be required for an alignment with polar amino acid side chains of the enzyme apoprotein in the formation of a mixed micellar lipoprotein complex. The enzyme reaction might occur at the resulting lipoprotein/water interface.  相似文献   

18.
Phospholipase C catalyzed hydrolysis of dimyristoyl phosphatidylcholine (DMPC) in phospholipid-bile salt mixed micelles was studied with particular attention on the relationship between interfacial enzyme activity and the physicochemical properties of substrate aggregates. Steady state kinetics is observed and it is argued that conditions for steady state exist because the enzyme encounters a steady supply of substrate by hopping between micelles at a rate faster than the chemical reaction rate. An existing kinetic model is reformulated to a more usable form. This presents a new approach to treating the kinetic data and allows extraction of the kinetic parameters of the model from the activity dependence on micellar lipid substrate surface concentration. The kinetic parameters were found to depend on the physicochemical properties of substrate aggregates, but remain constant over a range of lipid and bile salt concentrations. The substrate aggregates were characterized by time-resolved fluorescence quenching (TRFQ). The activity values and the micelle sizes group into two sets: (i) larger micelles for bile salt/lipid 5 with lower activity and longer steady state ( approximately 10 min). At least two sets of parameters, for bile salt/lipid 5, characterize the kinetics. Higher enzyme-micelle dissociation constant and lower catalytic rate are found for the group of smaller micelles. An explanation supporting our finding is that as micelles become smaller the overlap area for enzyme-micelle binding decreases, leading to weaker binding. Consequently the enzyme dissociation constant increases. Extension of the present approach to other phospholipases and substrates to establish its generality and correlation between micelle size and the catalytic rate are areas for future investigations.  相似文献   

19.
Taneva S  Johnson JE  Cornell RB 《Biochemistry》2003,42(40):11768-11776
CTP:phosphocholine cytidylyltranferase (CCT) regulates phosphatidylcholine (PC) biosynthesis. Its activity is controlled by reversible interactions with membrane lipids, mediated by an internal segment referred to as domain M. Although domain M peptides adopt an amphipathic alpha-helical structure when membrane bound, the structure of this domain in the context of the whole enzyme in the lipid-free and lipid-bound state is unknown. Here we derive lipid-induced secondary structural changes in CCTalpha using circular dichroism and three deconvolution programs. The analysis of two fragments, CCT236 (CCT1-236, housing the catalytic domain) and a synthetic domain M peptide (CCT237-293) aided the assignment of structural change to specific domains. To carry out this study, we developed a micellar lipid activating system that would avoid generation of CCT-induced lipid vesicle aggregates that interfere with the CD analysis. Lysophosphatidylcholine/phosphatidylglycerol (LPC/PG) mixed micelles supported full activation of CCT and caused an increase in the alpha-helix content of full-length CCT from 25 to 41%, at the expense of all other conformations. LPC/PG also induced an increase in alpha-helix content of the domain M peptide from 7 to 85% at the expense of all other conformers. This lipid system did not significantly affect the secondary structure of CCT236, nor did it affect the proteolytic fragmentation pattern of this region within full-length CCT, suggesting that the region containing the catalytic domain changes very little upon membrane activation of CCT. Our data suggest that lipids trigger a conformational switch in domain M from a mixed structure to an alpha-helix, thus creating a hydrophobic face for membrane insertion. Our results negate the idea that domain M is entirely helical in both the soluble and membrane-bound forms of CCT.  相似文献   

20.
The cholesterol oxidase-catalyzed oxidation of cholesterol in native low density (LDL) and high density lipoproteins (HDL3) as well as in monolayers prepared from surface lipids of these particles, has been examined. The objective of the study was to compare the oxidizability of cholesterol, and to examine the effects of lipid packing on oxidation rates. When [3H]cholesterol-labeled lipoproteins were exposed to cholesterol oxidase (Streptomyces sp.), it was observed that LDL [3H]cholesterol was oxidized much faster than HDL3 [3H]cholesterol. This was true both at equal cholesterol concentration per enzyme unit, and at equal amounts of lipoprotein particles per enzyme unit. About 95% of lipoprotein [3H]cholesterol was available for oxidation. The complete degradation of lipoprotein sphingomyelin by sphingomyelinase (Staphylococcus aureus) resulted in a 10-fold increase in the rate of LDL [3H]cholesterol oxidation, whereas the effects on rates of HDL3 [3H]cholesterol oxidation were less dramatic. A monolayer study with LDL surface lipids indicated that degradation of sphingomyelin loosened the lipid packing, because the ceramide formed occupied a smaller surface area than the parent sphingomyelin, and since the condensing effect of cholesterol on sphingomyelin packing was lost. The effects of sphingomyelin degradation on lipid packing in monolayers of HDL3-derived surface lipids were difficult to determine from monolayer experiments. Based on the finding that cholesterol oxidases are surface pressure-sensitive with regard to their catalytic activity, these were used to estimate the surface pressure of intact LDL and HDL3. The cut-off surface pressure of a Brevibacterium enzyme was 25 mN/m and 20 mN/m in monolayers of LDL and HDL3-derived surface lipids, respectively.(ABSTRACT TRUNCATED AT 250 WORDS)  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