首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
We propose a method for characterizing quantitatively the stereoselectivity of lipases during hydrolysis of triacylglycerols. Although it is of general applicability, we demonstrate it specifically for sn-1,3-regiospecific lipases. In this case the method generates a "stereoselectivity fingerprint" that consists of ratios of the specificity constants for the various reactions that produce and consume the 1,2-sn- and 2,3-sn-diacylglycerols. We use the method to determine the stereoselectivity fingerprint of several lipases during the hydrolysis of the prochiral substrate triolein. Our method opens up the possibility of correlating quantitative fingerprints with structural information, in the quest to elucidate the mechanisms underlying the stereoselectivity of lipases.  相似文献   

2.
In the present study, we propose a general and accessible method for the resolution of enantiomeric 1,2-sn- and 2,3-sn-diacylglycerols based on derivatization by isocyanates, which can be easily used routinely by biochemists to evaluate the stereopreferences of lipases in a time course of triacylglycerol (TAG) hydrolysis. Diacylglycerol (DAG) enantiomers were transformed into carbamates using achiral and commercially available reagents. Excellent separation and resolution factors were obtained for diacylglycerols present in lipolysis reaction mixtures. This analytical method was then applied to investigate the stereoselectivity of three model lipases (porcine pancreatic lipase, PPL; lipase from Rhizomucor miehei, MML; and recombinant dog gastric lipase, rDGL) in the time course of hydrolysis of prochiral triolein as a substrate. From the measurements of the diglyceride enantiomeric excess it was confirmed that PPL was not stereospecific (position sn-1 vs sn-3 of triolein), whereas MML and rDGL preferentially hydrolyzed the ester bond at position sn-1 and sn-3, respectively. The enantiomeric excess of DAGs was not constant with time, decreasing with the course of hydrolysis. This was due to the fact that DAGs can be products of the stereospecific hydrolysis of TAGs and substrates for stereospecific hydrolysis into monoacylglycerols.  相似文献   

3.
A simple method is described for the determination of molecular species of enantiomeric sn-1,2- and sn-2,3-diacylglycerols derived from natural triacylglycerols by Grignard degradation. The method is based on a preparative separation of the enantiomeric diacylglycerols as 3,5-dinitrophenylurethane (DNPU) derivatives by high performance liquid chromatography (HPLC) on a chiral column (25 cm x 4.6 mm ID) containing R-(+)-1-(1-naphthyl)ethylamine as a stationary phase. This is followed by polar capillary gas-liquid chromatography (GLC) of the trimethylsilyl (TMS) ether derivatives of the enantiomeric diacylglycerols derived from the DNPU derivatives using trichlorosilane, which does not cause acyl migration and racemization during the reaction. The cleavage is better than 94% complete. The method was standardized with synthetic sn-1,2- and sn-2,3-dipalmitoyl- and rac-1,2-dioleoylglycerols and was applied to the identification and quantitation of individual molecular species of enantiomeric diacylglycerols generated by Grignard degradation of the triacylglycerols from corn oil, cocoa butter, and lard.  相似文献   

4.
The 2-monoacylglycerol acyltransferase (EC 2.3.1.22, acylglycerol palmitoyl transferase) catalyzes the synthesis of X-1,2-diacylglycerols from 2-monoacylglycerol and acyl CoA with an apparently variable stereochemical specificity. A microassay for determining the ratio of sn-1,2- and sn-2,3-diacylglycerols formed by the acylation of radioactive 2-monoacylglycerol in intact cells or in cell-free systems in the presence of free fatty acids and cofactors has been developed. The diacylglycerols are isolated by thin-layer chromatography using nonradioactive racemic diacylglycerols as carriers. The enantiomer content is determined following a chemical synthesis of X-1,2-diacylphosphatidylcholines and a stereospecific stepwise release of the sn-1,2- and sn-2,3-diacylglycerols by phospholipase C. By using thin-layer chromatography for the isolation of the hydrolysis products, known samples ranging in enantiomer ratios from 0.05 to 20 and containing 5000 to 200,000 cpm can be assayed to within 1% of the major and within 10% of the minor enatiomer content. The method is applicable to the determination of the enantiomer content of X-1,2-diacylglycerols generated via other acyltransferases and via lipolysis of triacylglycerols and diacylglycerolphospholipids in other biological systems.  相似文献   

5.
The structure of mucosal triacylglycerols was studied in rat intestinal mucosa in vivo during the absorption of a low molecular weight fraction of butter oil and of the corresponding free fatty acids of medium and long chain length. The mucosal lipids were isolated by solvent extraction and the acylglycerol structures were determined by combined AgNO3- thin-layer chromatography and gas-liquid chromatography techniques and stereospecific analysis. Evidence was obtained for a rapid biosynthesis of triacylglycerols from diacylglycerols arising from the operation of both the monoacylglycerol and the phosphatidic acid biosynthetic pathways. Both sn-1,2- and sn-2,3-diacylglycerols appeared to be converted to triacylglycerols at significant rates, but a preferential utilization of sn-1,2-diacylglycerols could not be excluded. Endogenous dilution varied from a miniumum of 5% during triacylglycerol biosynthesis from monoacylglycerols to 15% during their synthesis from free fatty acids, and was characterized by a preferential placement of the endogenous acids in the sn-3 and 2 positions of the triacylglycerol molecules. Exogenous myristic acid was preferentially associated with the sn-3 position, and stearic acid became preferentially bound to the sn-1 position. The complexity of the triacylglycerol end products prevented an exact estimate of the contribution of the phosphatidic acid pathway, but the acylglycerol structures were compatible with a minimum of 20% of total triacylglycerol yield at all times.  相似文献   

6.
The biosynthesis of diacylglycerols was studied in rat intestinal mucosa during in vivo absorption of a low molecular weight fraction fraction of butter oil and of the corresponding medium and long chain fatty acids. The experimental fat solutions were given by stomach tube to the animals after a 24-h fast and mucosal scraping were collected 3 h later. The lipids were isolated and the acylclycerols determined by combined thin-layer chromatography gas-liquid chromatography techniques and stereospecific analyses. Free fatty acid feeding led mainly to sn-1,2-diacyl-glycerols, which contained exogenous and endogenous fatty acids. During triacylglycerol feeding, both sn-1,2-and sn-2,3-diacylglycerols were recovered in significant amounts from the intestinal mucosa. The composition of the sn-2,3-diacylglycerols corresponded to that with exogenous fatty acids but the sn-1,2-diacylglycerols clearly contained both exogenous and endogenous fatty acids. In all cases it was possible to isolate endogenous sn-1,2-diacylglycerols made up largely of species with linoleic and arachidonic acids in the 2 position and palmitic and stearic acids in the 1 position, which apparently were not converted to triacylglycerols. The in vivo reacylation of 2-monoacylglycerols via both sn-1,2- and sn-2,3-diacylglycerols is in agreement with similar findings in vitro with everted sacs of rat intestinal mucosa.  相似文献   

7.
In intact platelets, a permeable diacylglycerol having a 1,2-sn- but not 2,3-sn- configuration activated protein kinase C directly. In the presence of Ca2+-ionophore this diacylglycerol caused full activation of platelet release reaction. 1,3-Isomer was inactive. Among these isomers only 1,2-sn-diacylglycerol was converted rapidly to the corresponding phosphatidic acid in both intact and broken cell preparations. Thus, the diacylglycerol which functions in stimulus-response coupling possesses a 1,2-sn-glycerol backbone, and other isomers are not involved in the signal transduction through the protein kinase C pathway.  相似文献   

8.
A series of synthetic analogues of d-GlcN alpha 1-6-d-myo-inositol-1-HPO(4)-sn-1,2-dipalmitoylglycerol, consisting of 22 variants of the d-GlcN or lipid components, were tested in trypanosomal and human (HeLa) cell-free systems. The assays measured the abilities of the analogues to act as substrates or inhibitors of the enzymes of glycosylphosphatidylinositol biosynthesis downstream of GlcNAc-phosphatidylinositol (GlcNAc-PI) de-N-acetylase. One compound, 4-deoxy-d-GlcN alpha 1-6-d-myo-inositol-1-HPO(4)-sn-1,2-dipalmitoylglycerol, proved to be an inhibitor of both the trypanosomal and HeLa pathways, whereas 4-O-methyl-d-GlcN alpha 1-6-d-myo-inositol-1-HPO(4)-sn-1,2-dipalmitoylglycerol and the 4'-epimer, d-GalN-alpha1-6-d-myo-inositol-1-HPO(4)-sn-1,2-dipalmitoylglycerol, were neither substrates nor inhibitors. The results with other analogues showed that the 6-OH of the alpha-d-GlcN residue is not required for substrate recognition in the trypanosomal and human pathways, whereas the 3-OH group is essential for both. Parasite-specific recognition of the beta-linked analogue d-GlcN beta 1-6-d-myo-inositol-1-HPO(4)-sn-1,2-dipalmitoylglycerol is striking. This suggests that, like the GlcNAc-PI de-N-acetylase, the trypanosomal glycosylphosphatidylinositol alpha-mannosyltransferases, inositol acyltransferse and ethanolamine phosphate transferase, do not recognize the 2-, 3-, 4-, and 5-OH groups of the d-myo-inositol residue, whereas the human inositol acyltransferase and/or first alpha-mannosyltransferase recognizes one or more of these groups. All of the various lipid analogues tested served as substrates in both the trypanosomal and HeLa cell-free systems, suggesting that a precise lipid structure and stereochemistry are not essential for substrate recognition. However, an analogue containing a single C18:0 alkyl chain in place of sn-1,2-dipalmitoylglycerol proved to be a better substrate in the trypanosomal than in the HeLa cell-free system. These findings should have a bearing on the design of future generations of specific inhibitors of the trypanosomal glycosylphosphatidylinositol biosynthetic pathway.  相似文献   

9.
The aim of the present study was to investigate the applicability of a previously developed method for the analysis of triacylglycerol molecular species to the simultaneous determination of triacylglycerols, diacylglycerols and monoacylglycerols of human very-low-density lipoproteins (VLDL). Ten elderly women were recruited for the study. Blood was obtained in fasting conditions and VLDL were isolated by ultracentrifugation. Neutral lipids were separated by solid-phase extraction and were subsequently injected on a reversed-phase HPLC system, with an elution system composed of acetone in acetonitrile. The method allowed the separation of four monoacylglycerols, 18 diacylglycerols and 24 triacylglycerols, including the resolution of positional isomers of diacylglycerols. Monoacylglycerols were composed of oleic, linoleic, palmitic and stearic acids. The major diacylglycerols were 1,2-dilinoleoyl-glycerol and 1,3-dilinoleoyl-glycerol (14.24+/-1.02 and 17.93+/-1.42%, respectively). The main triacylglycerols quantified were dioleoyl-stearoyl-glycerol (OOS), oleoyl-dipalmitoyl-glycerol (OPP), trilinoleoyl-glycerol (LLL) and linoleoyl-distearoyl-glycerol (LSS), accounting for 11.25+/-2.15, 10.14+/-2.05, 9.35+/-2.30 and 8.56+/-1.56%, respectively. An inverse relationship between polarity and fatty acid disappearance from triacylglycerols (r(2)=0.82, P<0.05) and from diacylglycerols (r(2)=0.93, P<0.01) was discovered. In conclusion, the method allowed, for the first time, the easy, rapid and simultaneous determination in a single chromatogram of triacylglycerol, diacylglycerol and monoacylglycerol molecular species of human VLDL by reversed-phase HPLC.  相似文献   

10.
A bacterium that assimilates 2,3-dichloro-1-propanol was isolated from soil by enrichment culture. The strain was identified as Pseudomonas sp. by the taxonomic studies. The strain converted 2,3-dichloro-1-propanol to 3-chloro-1,2-propanediol, releasing chloride ion. The conversion was stereospecific because the residual 2,3-dichloro-1-propanol and formed 3-chloro-1,2-propanediol gave optical rotation. The resting cells converted various halohydrins to the dehalogenated alcohols, and cell-free extracts had strong epoxyhydrolase activity. These results indicated that the strain assimilated 2,3-dichloro-1-propanol via 3-chloro-1,2-propanediol, glycidol, and glycerol. The possibility to manufacture optically active 2,3-dichloro-1-propanol is discussed.  相似文献   

11.
Resting cells of ethene grown Mycobacterium 2W produced 1,2-epoxypropane stereospecifically from propene as revealed by optical rotation, 1H n.m.r. using a chiral shift reagent, and also by complexation gas chromatography involving a glass capillary column coated with an optically active metal chelate. The gas-liquid chromatography method allowed the rapid screening of 11 strains with regard to stereospecific formation of 1,2-epoxypropane, 1,2-epoxybutane and 1-chloro-2,3-epoxypropane. Bacteria grown on either ethene, propene or butadiene all predominantly produced the R form of 1,2-epoxypropane from propene and 1,2-epoxybutane from 1-butene while the strains tested for 1-chloro-2,3-epoxypropane production from 3-chloro-1-propene predominantly accumulated the S enantiomer.  相似文献   

12.
The polymorphic behavior and molecular packing in different polymorphic forms of stereospecific triacylglycerols, 1,2-dipalmitoyl-3-lauroyl-sn-glycerol (PP12) and 1,2-dipalmitoyl-3-myristoyl-sn-glycerol (PP14) were examined by X-ray diffraction, differential scanning calorimetry, infrared and Raman spectroscopy techniques. The molecular packing and the polymorphic behavior of the metastable forms of these two compounds are very similar. In both compounds the isotropic liquid, on quenching, crystallizes into a hexagonally packed alpha-phase. The long spacing periodicity of the alpha-phase indicates a tilted bilayer structure to compensate the voids created by the short acyl chains. Upon heating, the alpha-phase converts into an orthorhombic perpendicular (O perpendicular) beta'2-phase. The beta' 2-phase, on further heating, exothermally converts to beta' 1-phase with slightly different O perpendicular subcell. On incubation of PP12 near the melting temperature of beta' 1-phase, there is a slow (hours) conversion to a beta-phase with triclinic parallel (T//) packing. However the beta' 1-phase of PP14 is the most stable structure and the beta-phase is absent. The thermodynamic parameters of the O perpendicular packings of these compounds compared to those of the higher homologue, tripalmitoylglycerol, suggest that the O perpendicular subcell is more stable in PP12 and PP14. The X-ray diffraction long spacings indicate that all the polymorphic forms of these compounds pack in a bilayer structure. The vibrational spectra confirm the lateral chain packing and inter- and intra-molecular order/disorder in the various polymorphic forms. The Raman and infrared spectra further indicate perturbation in the carbonyl and the end methyl plane regions of the beta'-phases.  相似文献   

13.
The glycolipid transfer protein purified from pig brain facilitates the transfer of various glycosphingolipids and glyceroglycolipids (Yamada, K., Abe, A. and Sasaki, T. (1985) J. Biol. Chem. 260, 4615-4621). In this paper, the transfer of Man beta 1----4Glc beta 1-Cer and Man alpha 1----4Man beta 1-Cer isolated from a bivalve, Corbicula japonica, the transfer of 3-[Glc alpha 1-]-sn-1,2-diacylglycerol and 3-[Glc alpha 1----2Glc alpha 1-]-sn-1,2-diacylglycerol prepared from Streptococcus lactis, and the transfer of 3-[Glc beta 1-]-rac-1,2-dipalmitylglycerol have been investigated. The transfer of these lipids from liposomes to mitochondria was assayed by the decrease of these lipids in the donor liposomes. These lipids were determined by chromatographic isolation of the lipids, acid hydrolysis of the isolated lipids, and subsequent determination of glucose in the hydrolysate. The glycolipid transfer protein facilitated the transfer of ManGlcCer and ManManGlcCer. The transfer protein did not facilitate the transfer of Glc alpha-diacylglycerol or Glc alpha Glc alpha-diacylglycerol. However, the transfer of Glc beta-dipalmitylglycerol was facilitated by the protein. These results strongly suggest that the glycolipid transfer protein has the specificity to the presence of beta-linked glucose or galactose directly linked to either ceramide or diacylglycerol.  相似文献   

14.
The second step of glycosylphosphatidylinositol anchor biosynthesis in all eukaryotes is the conversion of D-GlcNAcalpha1-6-d-myo-inositol-1-HPO(4)-sn-1,2-diacylglycerol (GlcNAc-PI) to d-GlcNalpha1-6-d-myo-inositol-1-HPO(4)-sn-1,2-diacylglycerol by GlcNAc-PI de-N-acetylase. The genes encoding this activity are PIG-L and GPI12 in mammals and yeast, respectively. Fragments of putative GlcNAc-PI de-N-acetylase genes from Trypanosoma brucei and Leishmania major were identified in the respective genome project data bases. The full-length genes TbGPI12 and LmGPI12 were subsequently cloned, sequenced, and shown to complement a PIG-L-deficient Chinese hamster ovary cell line and restore surface expression of GPI-anchored proteins. A tetracycline-inducible bloodstream form T. brucei TbGPI12 conditional null mutant cell line was created and analyzed under nonpermissive conditions. TbGPI12 mRNA levels were reduced to undetectable levels within 8 h of tetracycline removal, and the cells died after 3-4 days. This demonstrates that TbGPI12 is an essential gene for the tsetse-transmitted parasite that causes Nagana in cattle and African sleeping sickness in humans. It also validates GlcNAc-PI de-N-acetylase as a potential drug target against these diseases. Washed parasite membranes were prepared from the conditional null mutant parasites after 48 h without tetracycline. These membranes were shown to be greatly reduced in GlcNAc-PI de-N-acetylase activity, but they retained their ability to make GlcNAc-PI and to process d-GlcNalpha1-6-d-myo-inositol-1-HPO(4)-sn-1,2-diacylglycerol to later glycosylphosphatidylinositol intermediates. These results suggest that the stabilities of other glycosylphosphatidylinositol pathway enzymes are not dependent on GlcNAc-PI de-N-acetylase levels.  相似文献   

15.
1. Previous experiments showed that fatty acids were incorporated into triacylglycerols by homogenates of Ceratitis capitata larvae far more efficiently than by pharate adult homogenates. This metabolic behaviour of both stages of development of the insect has been interpreted throughout the existence of a different acyltransferase activity. To obtain new data on the acyltransferase mechanism, a time-course of the stereospecific incorporation of labelled myristic, palmitic, oleic and linoleic acids into the sn-positions of triacylglycerols has been followed. 2. Studies on the stereospecific incorporation of labelled fatty acids confirmed previous results. Palmitic acid was mainly incorporated into sn-1 and sn-3 positions whereas position 2 exhibited a low incorporation. Myristic acid acylated sn-3 position at a higher rate than it acylated the other sn-positions. Oleic acid was more specifically distributed than palmitic acid and linoleic acid was more efficiently incorporated than the monounsaturated acid. All these data reflect substrate differences in the acyltransferase activity of larval homogenates. Pharate adult homogenates incorporated fatty acids very scarcely and mainly into positions (1 + 3). 3. Kinetics of incorporation of labelled fatty acids into the sn-positions points to a non-random distribution with respect to the major saturated and unsaturated fatty acids in triacylglycerols of larvae of Ceratitis capitata.  相似文献   

16.
Using the classical emulsified system and the monomolecular film technique, we compared several interfacial properties of turkey pancreatic lipase (TPL) and human pancreatic lipase (HPL). TPL, like HPL, presented the interfacial activation phenomenon when vinyl ester was used as substrate. In the absence of colipase and bile salts, using tributyrin emulsion or monomolecular films of dicaprin at low surface pressure, TPL, unlike HPL, hydrolyzes pure tributyrin emulsion as well as dicaprin films maintained at low surface pressures. TPL was also able to hydrolyze triolein emulsion in the absence of any additive and despite the accumulation of long-chain free fatty acids at the interface. The difference of behaviors between TPL and HPL can be explained by the penetration power of each enzyme. The enzyme that presents the maximal pi(c) (TPL) interacts more efficiently with interfaces, and it is not denaturated at high interfacial energy. Turkey pancreatic lipase is more active on rac-dicaprin than HPL; a maximal ratio of 9 was found between the catalytic activities of the two lipases measured at their surface pressure optima (20 mN m(-1)). A kinetic study on the surface pressure dependency, stereospecificity, and regioselectivity of TPL was performed using enantiopure diglyceride (1,2-sn-dicaprin and 2,3-sn-dicaprin) and a prochiral isomer (1,3-dicaprin) that were spread as monomolecular films at the air-water interface. At low surface pressure (15 mN m(-1)), TPL acts preferentially on primary carboxylic ester groups of the diglyceride isomers (1,3-dicaprin), but at high surface pressure (23 mN m(-1)), this enzyme prefers both adjacent ester groups of the diglyceride isomers (1,2-sn-dicaprin and 2,3-sn-dicaprin). HPL prefers adjacent ester groups of the diglyceride isomers (1,2-sn-dicaprin and 2,3-sn-dicaprin). Furthermore, TPL was found to be markedly stereospecific for the sn-1 position of the 1,2-sn-enantiomer of dicaprin at low surface pressure (15 mN m(-1)), while at high surface pressure (23 mN m(-1)), this lipase presents a stereopreference for the sn-3 position of the 2,3-sn-enantiomer of dicaprin. HPL is stereospecific for the sn-1 position of the 1,2-sn-enantiomer of dicaprin both at 15 and 23 mN m(-1).  相似文献   

17.
Male rats with thoracic duct cannulae were intubated with mustard-seed oil or the corresponding fatty acid methyl esters and the lymph was collected over 0-24 h. The chylomicron and very low density lipoprotein fractions were obtained by conventional ultracentrifugation. The triacylglycerols and glycerophospholipids were isolated and the positional distribution and molecular association of fatty acids were determined by stereospecific and chromatographic methods. The oleic, linoleic, and linolenic acids were recovered in the lymph in the proportion in which they occurred in the fat fed, while eicosenoic, erucic, and lignoceric acids were rejected to about the same extent by the two pathways of intestinal triacylglycerol biosynthesis. It is shown that the lymph triacylglycerols arising via the monoacylglycerol or the phosphatidic acid pathway possess structures that are closely similar to each other and to that of the original mustard-seed oil. It is proposed that this is a result of comparable fatty acid and positional specificity of the acyltransferases associated with the acylglycerol synthesis in the animal and plant tissues and the wide range of fatty acid chain lengths in the mustard-seed oil.  相似文献   

18.
The lipoteichoic acids from Bifidobacterium bifidum spp. pennsylvanicum were extracted from cytoplasmic membranes or from disintegrated bacteria with aqueous phenol and purified by gel chromatography. The lipoteichoic acid preparations contained phosphate, glycerol, galactose, glucose and fatty acids in a molar ratio of 1.0:1.0:1.3:1.2:0.3. Chemical analysis and NMR studies of the native preparations and of products from various acid and alkaline hydrolysis procedures gave evidence for the structure of two lipoteichoic acids. The lipid anchor appeared to be 3-O-(6'-(sn-glycero-1-phosphoryl)diacyl-beta-D-galactofuranosyl)-sn-1, 2-diacylglycerol. The polar part showed two structural features not previously described for lipoteichoic acids. A 1,2-(instead of the usual 1,3-) phosphodiester-linked sn-glycerol phosphate chain is only used substituted at the terminal glycerol unit with a linear polysaccharide, containing either beta(1----5)-linked D-galactofuranosyl groups or beta(1----6)-linked D-glucopyranosyl groups.  相似文献   

19.
A suitable approach which combines nuclear magnetic resonance (NMR) spectroscopy and molecular dynamics (MD) simulations have been used to study the structure and the dynamics of the glycosylphosphatidylinositol (GPI) anchor Manalphal-2Manalpha1-6Manalphal -4GlcNalpha1-6myo-inositol-1-OPO(3)-sn-1,2-dimyristoylglycerol (1) incorporated into dodecylphosphatidylcholine (DPC) micelles. The results have been compared to those previously obtained for the products obtainable from (1) after phospholipase cleavage, in aqueous solution. Relaxation and diffusion NMR experiments were used to establish the formation of stable aggregates and the insertion of (1) into the micelles. MD calculations were performed including explicit water, sodium and chloride ions and using the Particle Mesh Ewald approach for the evaluation of the electrostatic energy term. The MD predicted three dimensional structure and dynamics were substantiated by nuclear overhauser effect (NOE) measurements and relaxation data. The pseudopentasaccharide structure, which was not affected by incorporation of (1) into the micelle, showed a complex dynamic behaviour with a faster relative motion at the terminal mannopyranose unit and decreased mobility close to the micelle. This motion may be better described as an oscillation relative to the membrane rather than a folding event.  相似文献   

20.
A systematic study was undertaken to observe the effects of dietary (dioleoyl) triacyl-sn-glycerol structure on chylomicron composition and metabolism. First studied was a series of 1,2-dioleoyl-3-(saturated)acyl-sn-glycerols, where the fatty acid esterified at the 3-position was varied from 14 to 24 carbons. Next a series of 1,3-dioleoyl-2-acyl glycerols was studied, with various fatty acids esterified at the glycerol 2-position. These stereospecific triacyl-sn-glycerols were fed to donor rats and lymph chylomicrons were isolated, analyzed, and reinjected into recipient rats to study their disappearance from plasma and delivery to tissues. As shown by their compositions, chylomicrons obtained after feeding triacylglycerols containing all sn-3 fatty acid of chain length greater than 20 carbons were under-represented, possibly due to poorer digestion by lipases, or poorer absorption by the intestine. The 18-carbon saturated chain fatty acid (stearic acid) was equally well represented in chylomicrons whether in the 2- or 3-position of the fed triacylglycerol. The presence of increased amounts of long-chain saturated fatty acids in donor chylomicron triacylglycerols affected the metabolism of chylomicrons injected into the bloodstream of recipient rats. In particular the rate of removal of labeled cholesteryl esters, tracing removal of the partially degraded chylomicron remnants was slowed by the saturated chains, with palmitic acid and the 20-carbon fatty acid, arachidic acid, showing the most severe effects. There were clear differences in the removal from plasma of injected lymph chylomicrons derived from fed triacylglycerols containing stearic acid in either the 2- or 3-position, with evidence for remnants from the symmetrical triacylglycerols being less rapidly removed from the circulating blood. This effect was investigated further by injected model emulsions of chylomicrons, where the 2-position was substituted with saturated or transunsaturated acyl chains. Quantitation of removal from the blood stream of these model lipoproteins confirmed that a saturated or transunsaturated long chain fatty acid at the 2-position of the emulsion triacylglycerols slowed remnant removal from the blood. In some cases, with both lymph chylomicron and with emulsions, the lipolytic step mediated by lipoprotein lipase was also slowed.  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