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.
H+-ATPase was solubilized from the tonoplast of mung bean (Vignaradiata L.) hypocotyls and purified by fast protein liquid chromatographyon a Mono Q ion-exchange column. The purified ATPase hardlycontained any phospholipid, but it did contain 10 to 15 moleculesof sterol and 25 to 30 molecules of glycolipid per ATPase molecule,and it had little activity without exogenously added phospholipids.Each individual polar head group, acylglyceride and fatty acidthat constituted a phospholipid was incapable by itself of activatingthe ATPase. Sterols and cerebroside had little activating effect.Maximal activation of ATPase was noted with asolectin or variousmolecular species of phosphatidylcholine (PC) at 0.005% to 0.01%(w/v). The activation by the various molecular species of PCwas dependent on the length and degree of unsaturation of fattyacyl chains. PC with two saturated and long fatty acyl chainsof more than 18 carbon atoms failed entirely to activate theATPase. PC, PS and PG with 1-palmitoyl (16:0)-2-oleoyl(18:1)fatty acyl chains all activated ATPase to nearly the same extentas asolectin, but the activation by PE and PA with the samefatty acyl composition was 52% and 15% of that by asolectin,respectively. The molecular species of PC with phase-transitiontemperatures below 50C activated ATPase, as determined at 38C.The dependence on temperature of the activation by the molecularspecies of PC indicated that the activation of the ATPase beganclose to the temperature of the phase transition of the PC added.These data indicate that phospholipids in the liquid-crystallinephase are essential for the catalytic activity of the ATPase. (Received June 4, 1992; Accepted January 18, 1993)  相似文献   

2.
Polyunsaturated phospholipids are common in biological membranes and affect the lateral structure of bilayers. We have examined how saturated sphingomyelin (SM; palmitoyl and stearoyl SM (PSM and SSM, respectively)) and phosphatidylcholine (PC; dipalmitoyl PC and 1-palmitoyl-2-stearoyl PC (DPPC and PSPC, respectively)) segregate laterally to form ordered gel phases in increasingly unsaturated PC bilayers (sn-1: 16:0 and sn-2: 18:1...22:6; or sn-1 and sn-2: 18:1…22:6). The formation of gel phases was determined from the lifetime analysis of trans-parinaric acid. Using calorimetry, we also determined gel phase formation by PSM and DPPC in unsaturated PC mixed bilayers. Comparing PSM with DPPC, we observed that PSM formed a gel phase with less order than DPPC at comparable bilayer concentrations. The same was true when SSM was compared with PSPC. Furthermore, we observed that at equal saturated phospholipid concentration, the gel phases formed were less ordered in unsaturated PCs having 16:0 in sn-1, as compared to PCs having unsaturated acyl chains in both sn-1 and sn-2. The gel phases formed by the saturated phospholipids in unsaturated PC bilayers did not appear to achieve properties similar to pure saturated phospholipid bilayers, suggesting that complete lateral phase separation did not occur. Based on scanning calorimetry analysis, the melting of the gel phases formed by PSM and DPPC in unsaturated PC mixed bilayers (at 45 mol % saturated phospholipid) had low cooperativity and hence most likely were of mixed composition, in good agreement with trans-parinaric acid lifetime data. We conclude that both interfacial properties of the saturated phospholipids and their chain length, as well as the presence of 16:0 in sn-1 of the unsaturated PCs and the total number of cis unsaturations and acyl chain length (18 to 22) of the unsaturated PCs, all affected the formation of gel phases enriched in saturated phospholipids, under the conditions used.  相似文献   

3.
The main fatty acids at the sn-1 position of phospholipids (PLs) are saturated or monounsaturated fatty acids such as palmitic acid (C16:0), stearic acid (C18:0), and oleic acid (C18:1) and are constantly replaced, like unsaturated fatty acids at the sn-2 position. However, little is known about the molecular mechanism underlying the replacement of fatty acids at the sn-1 position, i.e., the sn-1 remodeling. Previously, we established a method to evaluate the incorporation of fatty acids into the sn-1 position of lysophospholipids (lyso-PLs). Here, we used this method to identify the enzymes capable of incorporating fatty acids into the sn-1 position of lyso-PLs (sn-1 lysophospholipid acyltransferase [LPLAT]). Screenings using siRNA knockdown and recombinant proteins for 14 LPLATs identified LPLAT7/lysophosphatidylglycerol acyltransferase 1 (LPGAT1) as a candidate. In vitro, we found LPLAT7 mainly incorporated several fatty acids into the sn-1 position of lysophosphatidylcholine (LPC) and lysophosphatidylethanolamine (LPE), with weak activities toward other lyso-PLs. Interestingly, however, only C18:0-containing phosphatidylcholine (PC) and phosphatidylethanolamine (PE) were specifically reduced in the LPLAT7-mutant cells and tissues from knockout mice, with a concomitant increase in the level of C16:0- and C18:1-containing PC and PE. Consistent with this, the incorporation of deuterium-labeled C18:0 into PLs dramatically decreased in the mutant cells, while deuterium-labeled C16:0 and C18:1 showed the opposite dynamic. Identifying LPLAT7 as an sn-1 LPLAT facilitates understanding the biological significance of sn-1 fatty acid remodeling of PLs. We also propose to use the new nomenclature, LPLAT7, for LPGAT1 since the newly assigned enzymatic activities are quite different from the LPGAT1s previously reported.  相似文献   

4.
The fatty acid composition of phospholipids in thoracic muscles of Pyrrhocoris apterus was related to acclimatization temperature and diapause. Two unsaturated fatty acids, linoleic (18:2n-6) and oleic (18:1n-9), and two saturated, palmitic (16:0) and stearic (18:0), dominated at all temperatures. In contrast to most other reports, the proportion of unsaturated fatty acids did not increase with decreasing temperature; there was a positive correlation between the unsaturation ratio and temperature in total phospholipids (r=0.67). The most prominent response to cold acclimatization was an increase in the proportion of 16:0 fatty acid and a corresponding decrease in the proportion of fatty acids with 18 carbons. The negative correlation between the proportion of 16:0 and temperature was stronger in phospholipids with phosphatidylethanolamine (PE) head group (r=−0.85) than in phospholipids with phosphatidylcholine (PC) head group (r=−0.58). Changes in fatty acid profiles associated with photoperiodic induction of diapause had the same trend as changes related to cold acclimatization. Similar to most other reports, the proportion of PE increased, while the proportion of PC decreased with decreasing temperature. In contrast to a general rule, the PE-phospholipids were less unsaturated than PC-phospholipids.  相似文献   

5.
The reactions leading to triacylglycerol (TAG) synthesis in oilseeds have been well characterized. However, quantitative analyses of acyl group and glycerol backbone fluxes that comprise extraplastidic phospholipid and TAG synthesis, including acyl editing and phosphatidylcholine-diacylglycerol interconversion, are lacking. To investigate these fluxes, we rapidly labeled developing soybean (Glycine max) embryos with [14C]acetate and [14C]glycerol. Cultured intact embryos that mimic in planta growth were used. The initial kinetics of newly synthesized acyl chain and glycerol backbone incorporation into phosphatidylcholine (PC), 1,2-sn-diacylglycerol (DAG), and TAG were analyzed along with their initial labeled molecular species and positional distributions. Almost 60% of the newly synthesized fatty acids first enter glycerolipids through PC acyl editing, largely at the sn-2 position. This flux, mostly of oleate, was over three times the flux of nascent [14C]fatty acids incorporated into the sn-1 and sn-2 positions of DAG through glycerol-3-phosphate acylation. Furthermore, the total flux for PC acyl editing, which includes both nascent and preexisting fatty acids, was estimated to be 1.5 to 5 times the flux of fatty acid synthesis. Thus, recycled acyl groups (16:0, 18:1, 18:2, and 18:3) in the acyl-coenzyme A pool provide most of the acyl chains for de novo glycerol-3-phosphate acylation. Our results also show kinetically distinct DAG pools. DAG used for TAG synthesis is mostly derived from PC, whereas de novo synthesized DAG is mostly used for PC synthesis. In addition, two kinetically distinct sn-3 acylations of DAG were observed, providing TAG molecular species enriched in saturated or polyunsaturated fatty acids.  相似文献   

6.
Glycerophospholipids (GPs) that differ in the relative position of the two fatty acyl chains on the glycerol backbone (i.e., sn-positional isomers) can have distinct physicochemical properties. The unambiguous assignment of acyl chain position to an individual GP represents a significant analytical challenge. Here we describe a workflow where phosphatidylcholines (PCs) are subjected to ESI for characterization by a combination of differential mobility spectrometry and MS (DMS-MS). When infused as a mixture, ions formed from silver adduction of each phospholipid isomer {e.g., [PC (16:0/18:1) + Ag]+ and [PC (18:1/16:0) + Ag]+} are transmitted through the DMS device at discrete compensation voltages. Varying their relative amounts allows facile and unambiguous assignment of the sn-positions of the fatty acyl chains for each isomer. Integration of the well-resolved ion populations provides a rapid method (< 3 min) for relative quantification of these lipid isomers. The DMS-MS results show excellent agreement with established, but time-consuming, enzymatic approaches and also provide superior accuracy to methods that rely on MS alone. The advantages of this DMS-MS method in identification and quantification of GP isomer populations is demonstrated by direct analysis of complex biological extracts without any prior fractionation.  相似文献   

7.
Plasmenyl phospholipids (1-alk-1′-enyl-2-acyl-3-glycerophospholipids, plasmalogens) are a structurally unique class of lipids that contain an α-unsaturated ether substituent at the sn-1 position of the glycerol backbone. Several studies have supported the hypothesis that plasmalogens may be antioxidant molecules that protect cells from oxidative stress. Because the molecular mechanisms responsible for the antioxidant properties of plasmenyl phospholipids are not fully understood, the oxidation of plasmalogens in natural mixtures of phospholipids was studied using electrospray tandem mass spectrometry. Glycerophosphoethanolamine (GPE) lipids from bovine brain were found to contain six major molecular species (16:0p/18:1-, 18:1p/18:1-, 18:0p/20:4-, 16:0p/20:4, 18:0a/20:4-, and 18:0a/22:6-GPE). Oxidation of GPE yielded lyso phospholipid products derived from plasmalogen species containing only monounsaturated sn-2 substituents and diacyl-GPE with oxidized polyunsaturated fatty acyl substituents at sn-2. The only plasmalogen species remaining intact following oxidation contained monounsaturated fatty acyl groups esterified at sn-2. The mechanism responsible for the rapid and specific destruction of plasmalogen GPE may likely involve unique reactivity imparted by a polyunsaturated fatty acyl group esterified at sn-2. This structural feature may play a central role determining the antioxidant properties ascribed to this class of phospholipids.  相似文献   

8.
The contents and composition of lipids in citrus leaves in relation to their general resistance to infection by strains of Xanthomonas campestris pv. citri (Xcc) were determined. The composition and contents of total polar lipids and phospholipids and the degree of fatty acid unsaturation were significantly different between resistant and susceptible species. Leaves from resistant plants had less phospholipids, but more free sterols than those from susceptible plants. The predominant fatty acids in the phospholipids were palmitic (16:0), linoleic (18:2) and α-linolenic acid (18:3). The degree of fatty acid unsaturation was higher in susceptible plants than in resistant plants. Major phospholipids in citrus leaves were phosphatidylchloline (PC), phosphatidylethanolamine (PE), phosphatidylglycerol (PG), and phosphatidylinositol (PI). β-Sitosterol, campesterol and lanosterol were major sterols in the leaves of citrus plants with resistant species having a higher ratio of free sterols to total phospholipids than susceptible species. Differences in lipid metabolism may contribute to differences in Xcc-resistance of citrus leaves.  相似文献   

9.
We report the first study of the effect of NaCl on the double-bond isomeric composition of fatty acids and theirsn-1/sn-2 positional distribution in the membrane phospholipids of a moderately halophilic eubacterium. The major phospholipids, phosphatidylethanolamine and phosphatidylglycerol, ofVibrio costicola grown in 1M or 3M NaCl both have ansn-1 saturated,sn-2 unsaturated distribution of fatty acids. There is a greater effect of salinity on the fatty acid composition of phosphatidylglycerol compared with phosphatidylethanolamine. The fatty acids in phosphatidylethanolamine of cultures grown in 1M compared with 3M NaCl have the same unsaturation index and average chain length, but different double-bond isomeric compositions. In comparison, the fatty acid composition of phosphatidylglycerol is more unsaturated, with a different double-bond isomeric distribution, and has a shorter average chain length in cultures grown in 3M compared with 1M NaCl. The pattern of fatty acid isomers of 16:1 and 18:1 shows thatV. costicola uses the anaerobic pathway of fatty acid biosynthesis. The presence of the isomers 16:1c11 and 18:1c13 in the phospholipids of cultures grown in 3M but not in 1M NaCl indicates that external salinity affects the specificity of fatty acid synthetase in this moderately halophilic bacterium.  相似文献   

10.
Root and hypocotyl plasma membrane H+-ATPases were partially purified from deoxycholate-solubilized fractions of microsomes in mung bean (Vigna radiata L.) plants in the presence of glycerol. Certain properties of the ATPases and the manner in which phospholipids affect their activity were compared. Root ATPase was similar to hypocotyl ATPase with respect to substrate specificity, salt stimulation, pH dependence, Km for ATP·Mg2+ and inhibitor sensitivity, except for inhibition by vanadate. Both purified ATPases required phospholipids for their activation. Optimum concentrations of exogenously added phospholipid mixture (asolectin) to hypocotyl and root ATPase mixture were 0.03% and 1.0%, respectively. Root ATPase activation did not decrease if more than 1.0% asolectin was added. Qualitatively, phosphatidylserine and phosphatidylcholine brought about greater ATPase activation than other phospholipids. The hypocotyl ATPase was activated by phosphatidylinositol, phosphatidylserine and phosphatidylglycerol to a greater extent than the root ATPase. Root, but not hypocotyl ATPase, was slightly inhibited by the addition of phosphatidylinositol, phosphatidylethanolamine, and phosphatidic acid. The hypocotyl plasma membrane contained phosphatidylinositol + phosphatidylserine, phosphatidylglycerol and phosphatidic acid, and unsaturated fatty acids in greater abundance than the root plasma membrane. The differential activation of the plasma membrane ATPases may arise from these differences.  相似文献   

11.
The interaction of dietary fats and carbohydrates on liver mitochondria were examined in male FBNF1 rats fed 20 different low-fat isocaloric diets. Animal growth rates and mitochondrial respiratory parameters were essentially unaffected, but mass spectrometry-based mitochondrial lipidomics profiling revealed increased levels of cardiolipins (CLs), a family of phospholipids essential for mitochondrial structure and function, in rats fed saturated or trans fat-based diets with a high glycemic index. These mitochondria showed elevated monolysocardiolipins (a CL precursor/product of CL degradation), elevated ratio of trans-phosphocholine (PC) (18:1/18:1) to cis-PC (18:1/18:1) (a marker of thiyl radical stress), and decreased ubiquinone Q9; the latter two of which imply a low-grade mitochondrial redox abnormality. Extended analysis demonstrated: i) dietary fats and, to a lesser extent, carbohydrates induce changes in the relative abundance of specific CL species; ii) fatty acid (FA) incorporation into mature CLs undergoes both positive (>400-fold) and negative (2.5-fold) regulation; and iii) dietary lipid abundance and incorporation of FAs into both the CL pool and specific mature tetra-acyl CLs are inversely related, suggesting previously unobserved compensatory regulation. This study reveals previously unobserved complexity/regulation of the central lipid in mitochondrial metabolism.  相似文献   

12.
Summary Ligninase activity of Phanerochaete chrysosporium INA-12 was increased when vegetable oils emulsified with sorbitan polyoxyethylene monooleate (Tween 80) were added to growth medium. Maximal enzyme yield was 22.0 nkat·ml-1 in olive oil cultures after 4 days incubation. P. chrysosporium INA-12 was also able to utilize tall oil fatty acids for ligninase synthesis. An extracellular lipase activity was detected during the primary phase of growth in culture containing vegetable oils. On the other hand, ligninase production was 1.5-fold enhanced when olive oil cultures were supplemented with soybean asolectin as a phospholipid source. In cultures supplied with olive oil plus asolectin, P. chrysosporium INA-12 mycelium exhibited a preferential enrichment of oleic acid (C18:1), phosphatidylcholine (PC) and lysophosphatidylcholine (LPC) as compared to lipid-free medium. PC and LPC enrichment was associated with an increased ratio of saturated versus unsaturated fatty acids of phospholipids.  相似文献   

13.
The seed oil of Thunbergia alata has an unusual fatty acid composition which consists of more than 80 % 16:1Δ6. This fatty acid is produced in the plastid by the action of a Δ6 palmitoyl (16:0)-ACP desaturase. To examine the biosynthesis of triacylglycerol (TAG) containing high concentrations of this unusual monoenoic fatty acid, endosperm dissected from developing T. alata seeds was labeled with [1-14C]-acetate. At early time points (5–15 min), the predominant labeled lipid was PC whereas at later time points (greater than 30 min) TAG became the major labeled lipid. Analysis of the acyl group composition of each lipid revealed that radiolabeled 16:1Δ6 was highest at early time points in PC while at later time points, it was found to be highest in TAG. Further analysis of the distribution of labeled acyl groups within PC indicated that 16:1Δ6 at the sn-2 position comprised the majority (55–78 %) of total labeled acyl groups whereas 16:1Δ6 at the sn-1 position constituted only a small fraction (12–15 %) of the total labeled acyl groups. In contrast, unlabeled PC contained lower amounts of 16:1Δ6 (16 %) at the sn-2 position. These results are consistent with previous studies suggesting a flux of novel monoenoic acids through PC during TAG biosynthesis, and furthermore imply a stereospecific flux through the sn-2 position of PC.  相似文献   

14.
Stereospecific analyses of glycerolipids from 7-, 14- and 21-day-old cultures of the yeast Lipomyces lipoferus revealed that each position of the glycerolipids had a unique distribution of fatty acids which changed to varying degrees with age, and that, in the triacylglycerols, age had a greater effect on fatty acid content at sn-3 that at sn-1 or sn-2. Age-related changes in unsaturation were, however, greater in the phospholipids than in the triacylglycerols. Among the major phospholipids of L. lipoferus (phosphatidylcholine, phosphatidylinositol and phosphatidylethanolamine), changes in the proportion of unsaturated to saturated fatty acids, and in the number of double bonds per mole, were greater at sn-2 than at sn-1, except for phosphatidylinositol between 14 and 21 days of age. The pattern of acylation of phosphatidylinositol between 14 and 21 days was thus different from that of phosphatidylcholine and phosphatidylethanolamine. Furthermore, at the three ages investigated, phosphatidylinositol had low and relatively constant levels of unsaturation compared with phosphatidylcholine and phosphatidylethanolamine. The net decrease in phospholipid double bonds per mole in aging cells of L. lipoferus, and previous data, suggest that aging in this yeast is accompanied by a decrease in membrane fluidity.  相似文献   

15.
Wada H  Murata N 《Plant physiology》1990,92(4):1062-1069
Changes in glycerolipid and fatty acid composition with a change in growth temperature were studied in the cyanobacterium, Synechocystis PCC6803. Under isothermal growth conditions, temperature did not significantly affect the composition of the various classes of lipids, but a decrease in temperature altered the degree of unsaturation of C18 acids at the sn-1 position, but not that of C16 acids at the sn-2 position of the glycerol moiety in each class of lipids. When the growth temperature was shifted from 38°C to 22°C, the desaturation of C18 acids, but not that of C16 acids, was stimulated. The desaturation of fatty acids occurred only in the light and was inhibited by chloramphenicol, rifampicin and 3-(3,4-dichlorophenyl)-1, 1-dimethylurea, but not by cerulenin, an inhibitor for fatty acid synthesis. These findings suggest that desaturase activities are induced after a shift from a higher to a lower temperature, and that the desaturation of fatty acids is connected with the reactions involved in photosynthetic electron transport.  相似文献   

16.
The fatty acid composition of phospholipids in the microsomesand the vanadate-sensitive H+-ATPase activity of the roots ofone-year-old Scots pine (Pinus sylvestris L.) seedlings werestudied during flushing in spring. The seedlings in hydroponiccultures were subjected to different root temperatures (5, 12or 20°C). The shoot was maintained at 20/15° C (day/night)during the 35 d experiment. After 35 d at 5° C, root growthwas totally inhibited and shoot growth partly inhibited. In roots grown at 5° C the fatty acid composition of themicrosomal phospholipids and the degree of fatty acid unsaturation(bond index) were unchanged, while in roots grown at 12 and20° C the fatty acid composition changed and bond indexdecreased. At those root temperatures, the most obvious changewas a decline in the proportion of linolenic acid (C18:3). Inthe new white roots grown either at 12°C or 20°C theproportion of C18:2 was higher and the proportion of C18:3 lowerthan in 1-year-old roots. Independently of root temperature,H+-ATPase activity, determined on a fresh weight basis, declinedto half of the original activity during the experiment. Thedecline in H+ -ATPase activity was most rapid during the firstweek. In the old roots the decline in H+-ATPase activity followedclosely the decline in amount of membrane protein. In new rootsH+-ATPase activity was high and increased with increasing roottemperature. These results suggest that in the roots of Scotspine seedlings, vanadate-sensitive H+-ATPase activity is dependenton age, while changes in the microsomal fatty acid compositionof phospholipids are regulated mainly by root temperature. Key words: Fatty acids of phospholipids, microsomes, H+-ATPase, root temperature, Scots pine  相似文献   

17.
Lipid composition of the membrane and rhodopsin packing density strongly modulate the early steps of the visual response of photoreceptor membranes. In this study, lipid-order and bovine rhodopsin function in proteoliposomes composed of the sn-1 chain perdeuterated lipids 14:0d27-14:1-PC, 16:0d31-16:1-PC, 18:0d35-18:1-PC, or 20:0d39-20:1-PC at rhodopsin/lipid molar ratios from 1:70 to 1:1000 (mol/mol) were investigated. Clear evidence for matching of hydrophobic regions on rhodopsin transmembrane helices and hydrophobic thickness of lipid bilayers was observed from 2H nuclear magnetic resonance order parameter measurements at low rhodopsin concentrations. Thin bilayers stretched to match the length of transmembrane helices observed as increase of sn-1 chain order, while thicker bilayers were compressed near the protein. A quantitative analysis of lipid-order parameter changes suggested that the protein adjusts its conformation to bilayer hydrophobic thickness as well, which confirmed our earlier circular-dichroism measurements. Changes in lipid order parameters upon rhodopsin incorporation vanished for bilayers with a hydrophobic thickness of 27 ± 1 Å, suggesting that this is the bilayer thickness at which rhodopsin packs in bilayers at the lowest membrane perturbation. The lipid-order parameter studies also indicated that a hydrophobic mismatch between rhodopsin and lipids triggers rhodopsin oligomerization with increasing rhodopsin concentrations. Both hydrophobic mismatch and rhodopsin oligomerization result in substantial shifts of the equilibrium between the photointermediates metarhodopsin I and metarhodopsin II; increasing bilayer thickness favors formation of metarhodopsin II while oligomerization favors metarhodopsin I. The results highlight the importance of hydrophobic matching for rhodopsin structure, oligomerization, and function.  相似文献   

18.
Over one hundred different phospholipid molecular species are known to be present in mammalian cells and tissues. Fatty acid remodeling systems for phospholipids including acyl-CoA:lysophospholipid acyltransferases, CoA-dependent and CoA-independent transacylation systems, are involved in the biosynthesis of these molecular species. Acyl-CoA:lysophospholipid acyltransferase system is involved in the synthesis of phospholipid molecular species containing sn-1 saturated and sn-2 unsaturated fatty acids. The CoA-dependent transacylation system catalyzes the transfer of fatty acids esterified in phospholipids to lysophospholipids in the presence of CoA without the generation of free fatty acids. The CoA-dependent transacylation reaction in the rat liver exhibits strict fatty acid specificity, i.e., three types of fatty acids (20:4, 18:2 and 18:0) are transferred. On the other hand, CoA-independent transacylase catalyzes the transfer of C20 and C22 polyunsaturated fatty acids from diacyl phospholipids to various lysophospholipids, especially ether-containing lysophospholipids, in the absence of any cofactors. CoA-independent transacylase is assumed to be involved in the accumulation of PUFA in ether-containing phospholipids. These enzymes are involved in not only the remodeling of fatty acids, but also the synthesis and degradation of some bioactive lipids and their precursors. In this review, recent progresses in acyltransferase research including the identification of the enzyme’s genes are described.  相似文献   

19.
Attempts at a better understanding of the cell membrane organization and functioning need to assess the physical properties which partly depend (i) on the positional distribution of the fatty acids in the membrane phospholipids (PLs) and (ii) on the way by which the PL molecular species are affected by exogenous fatty acids. To do that, the effects of essential (polyunsaturated) fatty acid (EFA) deficiency and enrichment were studied in the liver microsomes of piglets feeding on either an EFA-deficient diet or an EFA-enriched diet containing hydrogenated coconut oil or a mixture of soya + corn oils, respectively. After derivatization, the diacylated forms of choline and ethanolamine PLs were analyzed using a combination of Chromatographic techniques and fast-atom bombardmentmass spectrometry. The dinitrobenzoyl-diacylglycerol derivatives corresponding to the molecular species of phosphatidylcholine (PC) and phosphatidylethanolamine (PE) were identified. It appears that three factors brought about a marked apparent relative retention: the nature of (i) the base of the polar head, (ii) fatty acids at the sn−1 position and (iii) fatty acids at the sn−2 position. The highest apparent relative retentions were displayed by the 18:0–20:5(n−3)-PE and 16:0–22:6(n−3)-PE. It is noteworthy that the behavior of 20:3 n−9 — which is synthesized during the EFA-deficient diet by the same bioconversion system as 20:4 n−6 — was very similar to that of 20:4 n−6 during the formation of PC and PE molecular species and that the molecular species of PE containing 20:4(n−6) and 20:3(n−9), gathered together as metabolical homologues, were also apparently retained, particularly in association with 16:0. Present observations are consistent with some others showing retention or preferential distribution of EFA in PE and suggest that specific acyltransferase(s), ethanolamine phosphotransferase and methyltransferase would be mainly involved for PE and PC formation in liver endoplasmic reticulum. Fast-atom bombardment-mass spectrometry of intact phospholipids enables us to show that there is no very long chain dipolyunsaturated phospholipid in liver endoplasmic reticulum.  相似文献   

20.
《L' Année biologique》1999,38(1):17-26
Incorporation of 14C fatty acids in phospholipids of plasma membranes and sarcoplasmic reticulum of rat heart was studied. Mainly phosphatidylcholine (PC) and phosphatidylethanolamine (PE) were labelled. Our studies showed that the incorporation of unsaturated fatty acids (oleic and linoleic acid) was higher than for saturated fatty acids (palmitic and stearic acid). The range of uptake was between 0.2 and 2.2 nmol·mg−1 protein·h−1 and 0.5–7.4 nmol·μgatom−1 P1·h−1, respectively. Uptake of activity in individual phospholipids (measured after separation on TLC) was calculated as percentage of total activity. Incorporation in phosphatidylcholine was higher than in phosphatidylethanolamine. Phosphatidylcholine showed an increasing sequence for the following fatty acids: C18:0 < C16:0 < C18:0 < C18:2. However, phosphatidylethanolamine showed a decreasing sequence for the incorporation of the same fatty acids. Labelling of PC was always greater than for PE, except for stearic acid which was better incorporated into phosphatidylethanolamine. Uptake of the same fatty acid into phospholipids of sarcoplasmic reticulum was always higher than uptake into plasma membranes. As incorporation of fatty acids bound to albumin was studied in isolated membranes of rat heart, the addition of ATP and CoASH was an absolute requirement.  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