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 46 毫秒
1.
1. Age-related changes in the specific activity of palmitoyl-CoA synthetase, sn-glycerol 3-phosphate acyltransferase (EC 2.3.1.15) and the esterification of [3H]palmitate into endogenous lipid in the microsomal fraction from rabbit brain have been determined throughout development. 2. The increased specific activity of sn-glycerol 3-phosphate acyltransferase at the onset of myelination (rising in parallel with other lipogenic enzymes) is consistent with a direct role of the acyltransferase in promoting the accumulation of cerebral lipid. In adult brain microsomes, although the specific activity was low, the total activity was only 20% lower than during active myelination. 3. Palmitoyl-CoA, synthesized by the palmitoyl-CoA synthetase in the microsomal membrane, was the preferred substrate for the esterification of sn-glycerol 3-phosphate. There was no evidence for a pool of palmitoyl-CoA formed from palmitate. 4. The esterification of [3H]palmitate into membrane-bound lipid remained high throughout development and may be part of an acyl-exchange cycle via lysophospholipids. [3H]palmitate was incorporated into both neutral lipids and phospholipids, while phosphatidic acid was the major product of sn-[1(3)-3H]-glycerol-3-phosphate esterification. 5. The microsomal fraction contained a pool of unesterified fatty acid, which was activated and esterified into sn-glycerol 3-phosphate.  相似文献   

2.
Rat liver peroxisomes oxidized palmitate in the presence of ATP, CoA and NAD+, and the rate of palmitate oxidation exceeded that of palmitoyl-CoA oxidation. Acyl-CoA synthetase [acid: CoA ligase (AMP-forming); EC 6.2.1.3] was found in peroxisomes. The substrate specificity of the peroxisomal synthetase towards fatty acids with various carbon chain lengths was similar to that of the microsomal enzyme. The peroxisomal synthetase activity toward palmitate (40--100 nmol/min per mg protein) was higher than the rate of palmitate oxidation by the peroxisomal system (0.7--1.7 nmol/min per mg protein). The data show that peroxisomes activate long chain fatty acids and oxidize their acyl-CoA derivatives.  相似文献   

3.
An enzymatic basis for the formation of pulmonary surfactant lipids in rat has been presented. The free fatty acid pools in lung and liver consisted mainly of palmitic, stearic, oleic, and arachidonic acids with relatively less polyunsaturated fatty acids in lung than in liver. The acyl chain specificities of the acyl-CoA synthetase systems in lung and liver microsomes were similar in that most of fatty acids found in the free fatty acid pools were effectively activated by both systems. The acyl-CoA pools had compositions significantly different from those of the free fatty acid pools in lung and liver with relatively more stearate and less polyunsaturated fatty acids. The lung acyl-CoA pool contained mainly palmitate (29%), stearate (31%), and oleate (22%) with very little polyunsaturated acyl-CoAs to compete for esterification. The use of an equimolar mixture of palmitoyl-CoA and arachidonoyl-CoA to acylate the endogenous monoacyl-glycerophosphocholine isomers in the lung microsomes yielded both the 2-palmitate and 2-arachidonate diacyl forms, whereas the major products formed by liver microsomes were the 2-arachidonate and 1-palmitate forms. These results indicate that the 1-acyl isomer is the major monoacyl-glycerophosphocholine species serving as substrate in lung microsomes, whereas both 1-acyl and 2-acyl isomers are present in liver microsomes. Thus, the enrichment of saturated and oligoenoic acids in the acyl-CoA pool combined with the predominance of the 1-acyl isomer in the acyl acceptor pool and the relatively higher selectivity for palmitoyl-CoA by the 1-acyl-GPC acyltransferase activity of lung constitute an important basis for attributing some of the formation of pulmonary surfactant lipids in rats to acyltransferase action.  相似文献   

4.
We have examined the mechanism by which extracellular free fatty acids regulate fatty acid biosynthesis in Ehrlich ascites tumor cells. De novo biosynthesis in intact cells was inhibited by stearate greater than oleate greater than palmitate greater than linoleate. The amount of citrate and long chain acyl-CoA in the cells was not changed appreciably by the addition of free fatty acids to the incubation medium, indicating than free fatty acids do not regulate fatty acid biosynthesis by changing the total intracellular content of these metabolites. By measuring the incorporation of labeled free fatty acids into acyl-CoA, however, it was determined that the fatty acid composition of the acyl-CoA poolwas changed dramatically to reflect the composition of the exogenous free fatty acids. The relative inhibitory effects of different free fatty acids appear to depend on the ability of their acyl-CoA derivatives to regulate acyl-CoA carboxylase activity. The acyl-CoA concentration needed to produce 50% inhibition of purified Ehrlich cell carboxylase was found to be 0.68 mum for stearoyl-CoA, 1.6 mum for oleoyl-CoA, 2.2 mum for palmitoyl-CoA, 23 mum for myristoyl-CoA, 30 mum for lauroyl-CoA, and 37 mum for linoleoyl-CoA. In contrast to their effects on de novo synthesis, all of the free fatty acids added except stearate stimulated chain elongation in intact cells. Microsomal chain elongation, the major system for elongation in Ehrlich cells, also was regulated by the composition of the cellular acyl-CoA pool. Lauroyl-CoA, myristoyl-CoA, and palmitoyl-CoA were good substrates for elongation by isolated microsomes; oleoyl-CoA, and linoleoyl-CoA were intermediate; and stearoyl-CoA was a very poor substrate. We conclude that free fatty acids regulate fatty acid biosynthesis by changing the composition of the cellular acyl-CoA pool. These changes control the rate of malonyl-CoA production and, because of the acyl-CoA substrate specificity of the microsomal elongation system, modulate the amount of malonyl-CoA used for chain elongation.  相似文献   

5.
1. The CoA and carnitine esters of 2-bromopalmitate are extremely powerful and specific inhibitors of mitochondrial fatty acid oxidation. 2. 2-Bromopalmitoyl-CoA, added as such or formed from 2-bromopalmitate, inhibits the carnitine-dependent oxidation of palmitate or palmitoyl-CoA, but not the oxidation of palmitoylcarnitine, by intact liver mitochondria. 3. 2-Bromopalmitoylcarnitine inhibits the oxidation of palmitoylcarnitine as well as that of palmitate or palmitoyl-CoA. It has no effect on succinate oxidation, but inhibits that of pyruvate, 2-oxoglutarate or hexanoate; however, the oxidation of these substrates (but not of palmitate, palmitoyl-CoA or palmitoyl-carnitine) is restored by carnitine. 4. In damaged mitochondria, added 2-bromopalmitoyl-CoA does inhibit palmitoylcarnitine oxidation; pyruvate oxidation is unaffected by the inhibitor alone, but is impaired if palmitoylcarnitine is subsequently added. 5. The findings have been interpreted as follows. 2-Bromopalmitoyl-CoA inactivates (in a carnitine-dependent manner) a pool of carnitine palmitoyltransferase which is accessible to external acyl-CoA. This results in inhibition of palmitate or palmitoyl-CoA oxidation. A second pool of carnitine palmitoyltransferase, inaccessible to added acyl-CoA in intact mitochondria, can generate bromopalmitoyl-CoA within the matrix from external 2-bromopalmitoylcarnitine; this reaction is reversible. Such internal 2-bromopalmitoyl-CoA inactivates long-chain beta-oxidation (as does added 2-bromopalmitoyl-CoA if the mitochondria are damaged) and its formation also sequesters intramitochondrial CoA. Since this CoA is shared by pyruvate and 2-oxoglutarate dehydrogenases, the oxidation of their substrates is depressed by 2-bromopalmitoylcarnitine, unless free carnitine is available to act as a ;sink' for long-chain acyl groups. 6. These effects are compared with those reported for other inhibitors of fatty acid oxidation.  相似文献   

6.
The substrate specificity of fatty-acyl-CoA ligase in liver microsomes has been studied in a system in which fatty acids are present initially as complexes with unilamellar vesicles of phosphatidylcholine. The latter were prepared by cosonication of phospholipids and different fatty acids. As compared with previous studies of the enzyme the activity of acyl-CoA ligase is several-fold higher for assays carried out with fatty acid substrates added as components of a bilayer. This was true for all fatty acids studied. Also as compared with data reported previously in the literature there was a systematic relationship between the structure of fatty acids, activity at Vmax for synthesis of acyl-CoA and avidity of binding to the ligase. Activity at Vmax was greatest for lauric acid and decreased with increasing chain length. The apparent avidity of enzyme for fatty acids was greatest for octanoic acid and decreased as chain length increased.  相似文献   

7.
Very long chain fatty acids (lignoceric acid) are oxidized in peroxisomes and pathognomonic amounts of these fatty acids accumulate in X-adrenoleukodystrophy (X-ALD) due to a defect in their oxidation. However, in cellular homogenates from X-ALD cells, lignoceric acid is oxidized at a rate of 38% of control cells. Therefore, to identify the source of this residual activity we raised antibody to palmitoyl-CoA ligase and examined its effect on the activation and oxidation of palmitic and lignoceric acids in isolated peroxisomes from control and X-ALD fibroblasts. The normalization of peroxisomal lignoceric acid oxidation in the presence of exogenously added acyl-CoA ligases and along with the complete inhibition of activation and oxidation of palmitic and lignoceric acids in peroxisomes from X-ALD by antibody to palmitoyl-CoA ligase provides direct evidence that lignoceroyl-CoA ligase is deficient in X-ALD and demonstrates that the residual activity for the oxidation of lignoceric acid was derived from the activation of lignoceric acid by peroxisomal palmitoyl-CoA ligase. This antibody inhibited the activation and oxidation of palmitic acid but had little effect on these activities for lignoceric acid in peroxisomes from control cells. Furthermore, these data provide evidence that peroxisomal palmitoyl-CoA and lignoceroyl-CoA ligases are two different enzymes.  相似文献   

8.
Investigations on the cholic acid CoA ligase activity of rat liver microsomes were made possible by the development of a rapid, sensitive radiochemical assay based on the conversion of [3H]choloyl-CoA. More than 70% of the rat liver cholic acid CoA ligase activity was associated with the microsomal subcellular fraction. The dependencies of cholic acid CoA ligase activity on pH, ATP, CoA, Triton WR-1339, acetone, ethanol, magnesium, and salts were investigated. The hypothesis that the long chain fatty acid CoA ligase activity and the cholic acid CoA ligase activity are catalyzed by a single microsomal enzyme was investigated. The ATP, CoA, and cholic (palmitic) acid kinetics neither supported nor negated the hypothesis. Cholic acid was not an inhibitor of the fatty acid CoA ligase and palmitic acid was not a competitive inhibitor of the cholic acid CoA ligase. The cholic acid CoA ligase activity utilized dATP as a substrate more effectively than did the fatty acid CoA ligase activity. The cholic acid and fatty acid CoA ligase activities appeared to have different pH dependencies, differed in thermolability at 41 degrees, and were differentially inactivated by phospholipase C. Moreover, fatty acid CoA ligase activity was present in microsomal fractions from all rat organs tested while cholic acid CoA ligase activity was detected only in liver microsomes. The data suggest that separate microsomal enzymes are responsible for the cholic acid and the fatty acid CoA ligase activities in liver.  相似文献   

9.
Acyl-CoA ligases from rat brain microsomes: an immunochemical study   总被引:1,自引:0,他引:1  
Acyl-CoA ligase activities, solubilized from rat brain microsomes, were fractionated into three different peaks by hydroxyapatite chromatography. Based on physical and chemical properties, we suggested that peak A (pamitoyl-CoA ligase) and peak C (lignoceroyl-CoA ligase) were two different enzymes (A. Bhushan, R. P. Singh, and I. Singh (1986) Arch. Biochem. Biophys. 246, 374-380). We raised antibodies against purified liver microsomal palmitoyl-CoA ligase (EC 6.2.1.3) and examined the effect of this antibody on acyl-CoA ligase activities for palmitic, arachidonic and lignoceric acids in microsomal enzyme extract and different acyl-CoA ligase peaks from the hydroxyapatite column. In an enzyme activity assay system in microsomal extract, the antisera inhibited the palmitoyl-CoA ligase activity but had very little effect on the acyl-CoA ligase activities for arachidonic and lignoceric acids. This antisera inhibited the acyl-CoA ligase activities for these three fatty acids in peak A and had no effect on these activities in peak B or peak C. Western blot analysis demonstrated that antibody to liver microsomal palmitoyl-CoA ligase cross-reacted with only peak A (palmitoyl-CoA ligase), but not with peak B or peak C. This immunochemical study demonstrates that palmitoyl-CoA ligase does not share immunological determinants with acyl-CoA ligases in peaks B or C, thus demonstrating that palmitoyl-CoA ligase (peak A) is different from the arachidonoyl-CoA and lignoceroyl-CoA ligase activities in peaks B or C.  相似文献   

10.
Palmitic acid solubilized with Triton WR-1339 was converted to palmitoyl-CoA by microsomal membranes but lignoceric acid solubilized with Triton WR-1339 was not an effective substrate even though the detergent dispersed the same amount of these fatty acids and was also not inhibitory to the enzyme [I. Singh, R. P. Singh, A. Bhushan, and A. K. Singh (1985) Arch. Biochem. Biophys. 236, 418-426]. This observation suggested that palmitoyl-CoA and lignoceroyl-CoA may be synthesized by two different enzymes. We have solubilized the acyl-CoA ligase activities for palmitic and lignoceric acid of rat brain microsomal membranes with Triton X-100 and resolved them into three separate peaks (fractions) by hydroxylapatite chromatography. Fraction A (palmitoyl-CoA ligase) had high specific activity for palmitic acid and Fraction C (lignoceroyl-CoA ligase) for lignoceric acid. Specific activity of palmitoyl-CoA ligase for palmitic acid was six times higher than in Fraction C and specific activity of lignoceroyl-CoA ligase for lignoceric acid was four times higher than in Fraction A. At higher concentrations of Triton X-100 (0.5%), lignoceroyl-CoA ligase loses activity whereas palmitoyl-CoA ligase does not. Lignoceroyl-CoA ligase lost 60% of activity at 0.6% Triton X-100. Palmitoyl-CoA ligase (T1/2 of 4.5 min) is more stable at 40 degrees C than lignoceroyl-CoA ligase (T1/2 of 1.5 min). The pH optimum of palmitoyl-CoA ligase was 7.7 and that of lignoceroyl-CoA ligase was 8.4. Similar to our results with intact membranes, palmitic acid solubilized with Triton WR-1339 was converted to palmitoyl-CoA by palmitoyl-CoA ligase whereas lignoceric acid when solubilized with Triton WR-1339 was not able to act as substrate for lignoceroyl-CoA ligase. Since solubilized enzyme activities for synthesis of palmitoyl-CoA and lignoceroyl-CoA from microsomal membranes can be resolved into different fractions by column chromatography and demonstrate different properties, we suggest that in microsomal membranes palmitoyl-CoA and lignoceroyl-CoA are synthesized by two different enzymes.  相似文献   

11.
A novel radiochemical assay for long-chain fatty acid:CoASH ligase activity (AMP) (EC 6.2.1.3) has been developed based on the conversion of [3H]CoASH to long-chain fatty acyl CoA. Fatty acyl [3H]CoA was quantitatively retained on Millipore filters upon filtration of the acidified reaction mixture under conditions where the [3H]CoASH was not retained. The assay was developed using microsomes derived from isolated fat cells as the source of fatty acid:CoASH ligase activity. The assay performed at 25 degrees C for 10 min was linear with added microsomal protein up to 7 mug. The assay was linear with time up to 24 min when 1 mug of protein was employed. Fatty acid:CoASH ligase activity was strongly dependent on ATP and magnesium, was stimulated by Triton WR-1339, and was two- to fivefold dependent on added fatty acid. The filter assay is easier than existing assays based on incorporation of labeled fatty acid and is equally sensitive.  相似文献   

12.
The acyl-CoA ligases convert free fatty acids to acyl-CoA derivatives, and these enzymes have been shown to be present in mitochondria, peroxisomes, and endoplasmic reticulum. Because their activity is obligatory for fatty acid metabolism, it is important to identify their substrate specificities and subcellular distributions to further understand the cellular regulation of these pathways. To define the role of the enzymes and organelles involved in the metabolism of very long chain (VLC) fatty acids, we studied human genetic cell mutants impaired for the metabolism of these molecules. Fibroblast cell lines were derived from patients with X-linked adrenoleukodystrophy (X-ALD) and Zellweger's cerebro-hepato-renal syndrome (CHRS). While peroxisomes are present and morphologically normal in X-ALD, they are either greatly reduced in number or absent in CHRS. Palmitoyl-CoA ligase is known to be present in mitochondria, peroxisomes, and endoplasmic reticulum (microsomes). We found enzyme-dependent formation of lignoceroyl-CoA in these same organelles (specific activities were 0.32 +/- 0.12, 0.86 +/- 0.12, and 0.78 +/- 0.07 nmol/h per mg protein, respectively). However, lignoceroyl-CoA synthesis was inhibited by an antibody to palmitoyl-CoA ligase in isolated mitochondria while it was not inhibited in peroxisomes or endoplasmic reticulum (ER). This suggests that palmitoyl-CoA ligase and lignoceroyl-CoA are different enzymes and that mitochondria lack lignoceroyl-CoA ligase. This conclusion is further supported by data showing that oxidation of lignoceric acid was found almost exclusively in peroxisomes (0.17 nmol/h per mg protein) but was largely absent from mitochondria and the finding that monolayers of CHRS fibroblasts lacking peroxisomes showed a pronounced deficiency in lignoceric acid oxidation in situ (1.8% of control). In spite of the observation that lignoceroyl-CoA ligase activity is present on the cytoplasmic surface of ER, our data indicate that lignoceroyl-CoA synthesized by ER is not available for oxidation in mitochondria. This organelle plays no physiological role in the beta-oxidation of VLC fatty acids. Furthermore, the normal peroxisomal oxidation of lignoceroyl-CoA but deficient oxidation of lignoceric acid in X-ALD cells indicates that cellular VLC fatty acid oxidation is dependent on peroxisomal lignoceroyl-CoA ligase. These studies allow us to propose a model for the subcellular localization of various acyl-CoA ligases and to describe how these enzymes control cellular fatty acid metabolism.  相似文献   

13.
A. Van Tol  W. C. Hü  lsmann 《BBA》1969,189(3):342-353
1. The distribution of palmitoyl-CoA:carnitine palmitoyltransferase has been studied in subcellular fractions of rat liver. By using two different estimations for the enzyme activity and by differential centrifugation and linear sucrose density gradient centrifugation, the enzyme is shown to be localized both in mitochondria and microsomes.

2. The mitochondrial palmitoyl-CoA: carnitine palmitoyltransferase is localized in the inner membrane plus matrix fraction.

3. During palmitate oxidation by isolated mitochondria, in the presence of a physiological concentration of carnitine, palmitoylcarnitine accumulates. From this and experiments with sonicated mitochondria, it is concluded that the capacities of long-chain fatty acid activation and of palmitoyl-CoA:carnitine palmitoyltransferase in vitro by far exceed the capacity of fatty acid oxidation.  相似文献   


14.
O Lazo  M Contreras  I Singh 《Biochemistry》1990,29(16):3981-3986
We found that peroxisomal lignoceroyl-CoA ligase, like palmitoyl-CoA ligase, is present in the peroxisomal membrane whereas the peroxisomal beta-oxidation enzyme system is localized in the matrix. To further define the role of peroxisomal acyl-CoA ligases (membrane component) in providing acyl-CoA for peroxisomal beta-oxidation, we examined the transverse topographical localization of enzymatic sites of palmitoyl-CoA and lignoceroyl-CoA ligases in the peroxisomal membranes. The disruption of peroxisomes by various techniques resulted in the release of a "latent" pool of lignoceroyl-CoA ligase activity while palmitoyl-CoA ligase activity remained the same. Proteolytic enzyme treatment inhibited palmitoyl-CoA ligase activity in intact peroxisomes but had no effect on lignoceroyl-CoA ligase activity. Lignoceroyl-CoA ligase activity was inhibited only if peroxisomes were disrupted with detergent before trypsin treatment. Antibodies to palmitoyl-CoA ligase and to peroxisomal membrane proteins (PMP) inhibited palmitoyl-CoA ligase in intact peroxisomes, and no pool of "latent" activity appeared when antibody-treated peroxisomes were disrupted with detergent. On the other hand, disruption of PMP antibody-treated peroxisomes with detergent resulted in the appearance of a "latent" pool of lignoceroyl-CoA ligase activity. These results demonstrate that the enzymatic site of palmitoyl-CoA ligase is on the cytoplasmic surface whereas that for lignoceroyl-CoA ligase is on the luminal surface of peroxisomal membranes. This implies that palmitoyl-CoA is synthesized on the cytoplasmic surface and is then transferred to the matrix through the peroxisomal membrane for beta-oxidation in the matrix.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

15.
Abstract: In vivo rates of palmitate incorporation into brain phospholipids were measured in awake rats following programmed intravenous infusion of unesterified [9,10-3H]palmitate to maintain constant plasma specific activity. Animals were killed after 2–10 min of infusion by microwave irradiation and analyzed for tracer distribution in brain phospholipid and phospholipid precursor, i.e., brain unesterified palmitate and palmitoyl-CoA, pools. [9,10-3H]Palmitate incorporation into brain phospholipids was linear with time and rapid, with >50% of brain tracer in choline-containing glycerophospholipids at 2 min of infusion. However, tracer specific activity in brain phospholipid precursor pools was low and averaged only 1.6–1.8% of plasma unesterified palmitate specific activity. Correction for brain palmitoyl-CoA specific activity increased the calculated rate of palmitate incorporation into brain phospholipids (0.52 nmol/s/g) by ∼60-fold. The results suggest that palmitate incorporation and turnover in brain phospholipids are far more rapid than generally assumed and that this rapid turnover dilutes tracer specific activity in brain palmitoyl-CoA pool owing to release and recycling of unlabeled fatty acid from phospholipid breakdown.  相似文献   

16.
The different topology of palmitoyl-CoA ligase (on the cytoplasmic surface) and of lignoceroyl-CoA ligase (on the luminal surface) in peroxisomal membranes suggests that these fatty acids may be transported in different form through the peroxisomal membrane (Lazo, O., Contreras, M., and Singh, I. (1990) Biochemistry 29, 3981-3986), and this differential transport may account for deficient oxidation of lignoceric acid in X-adrenoleukodystrophy (X-ALD) (Singh, I., Moser, A. B., Goldfisher, S., and Moser, H. W. (1984) Proc. Natl. Acad. Sci. U.S.A. 81, 4203-4207). To define the transport mechanism for these fatty acids through the peroxisomal membrane and its possible implication to lignoceric acid metabolism in X-ALD, we examined cofactors and energy requirements for the transport of palmitic and lignoceric acids in isolated peroxisomes from rat liver and peroxisomes isolated from X-ALD and control fibroblasts. The similar rates of transport of palmitoyl-CoA (87.6 +/- 6.3 nmol/h/mg protein) and palmitic acid in the fatty acid activating conditions (83.4 +/- 5.1 nmol/h/mg protein) and lack of transport of palmitic acid (4% of palmitoyl-CoA transport) when ATP and/or CoASH were removed or substituted by alpha,beta-methyleneadenosine-5'-triphosphate (AMPCPOP) and/or desulfoCoA-agarose from assay medium clearly demonstrate that transport of palmitic acid requires prior synthesis of palmitoyl-CoA by palmitoyl-CoA ligase on the cytoplasmic surface of peroxisomes. The 10-fold higher rate of transport of lignoceric acid (5.3 +/- 0.6 nmol/h/mg protein) as compared with lignoceroyl-CoA (0.41 +/- 0.11 nmol/h/mg protein) and lack of inhibition of transport of lignoceric acid when ATP and/or CoASH were removed or substituted with AMPCPOP or desulfoCoA-agarose suggest that lignoceric acid is transported through the peroxisomal membrane as such. Moreover, the lack of effect of removal of ATP or substitution with AMPOPCP (a nonhydrolyzable substrate) demonstrates that the translocation of palmitoyl-CoA and lignoceric acid across peroxisomal membrane does not require energy. The transport, activation, and oxidation of palmitic acid are normal in peroxisomes from X-ALD. The deficient lignoceroyl-CoA ligase (13% of control) and oxidation of lignoceric acid (10% of control) as compared with normal transport of lignoceric acid into peroxisomes from X-ALD clearly demonstrates that pathogenomonic accumulation of very long chain fatty acids (greater than C22) in X-ALD is due to the deficiency of peroxisomal lignoceroyl-CoA ligase activity.  相似文献   

17.
Fatty acid CoA ligase (AMP) (EC 6.2.1.3) specific activity was increased approximately 2-fold in microsomes prepared from isolated rat fat cells incubated with 400 microunits of insulin/ml (2.9 nM) for 45 to 60 min compared to paired controls using an assay based on the conversion of [3H]oleic acid to [3H]oleoyl-CoA. Similar insulin-dependent increases in microsomal fatty acid CoA ligase specific activities were observed using an assay based on the conversion of [3H]CoA to fatty acyl-[3H]CoA. Fatty acid CoA ligase activity was predominately (about 80%) associated with the microsomal fraction. The insulin-dependent increase in microsomal fatty acid CoA ligase specific activity was maximal in 2 to 5 min at 400 microunits/ml. At 10 min, 80 to 100 microunits of insulin/ml caused a maximal increase in fatty acid CoA ligase specific activity. Similar apparent Km values for ATP, CoA, and fatty acid were observed for fatty acid CoA ligase activity in microsomal preparations from control and insulin-exposed cells. These data suggest that fatty acid CoA ligase activity is regulated in adipose tissue by insulin. Such regulation may serve to promote the capture of fatty acid and thereby, triacylglycerol synthesis in adipose tissue.  相似文献   

18.
Long chain fatty acyl coenzyme A ligase (EC 6.2.1.3) purified from rat liver mitochondria has been characterized with respect to several kinetic parameters. Many of the kinetic properties of the mitochondrial enzyme are similar to those of the purified microsomal enzyme with respect to palmitoyl-CoA formation, but there are distinct differences. The fatty acid and nucleotide specificities of the mitochondrial enzyme are similar to those of the microsomal enzyme, as are the apparent Km values for ATP and coenzyme A. On the other hand, the mitochondrial enzyme differs from the microsomal enzyme in that it has a lower pH optimum, is different in molecular weight, and does not show simple saturation kinetics with palmitate as substrate. Of particular interest is the evidence presented which indicates that the mitochondrial long chain fatty acyl-CoA ligase, unlike short and medium chain ligases, does not utilize an acyladenylate as an intermediate in the formation of fatty acyl-CoA.  相似文献   

19.
When the localization of mycolic acid biosynthetic activity was examined with Bacterionema matruchotii cells disrupted by the ultrasonic vibration method, activity was detected only in the cell wall fraction, not in the inner membrane nor in the 78,000g supernatant. Either the supernatant or sugar was absolutely required for the incorporation of [14C]palmitate into mycolic acids. Among sugars examined, glucose was most effective, with maltose being second. Unexpectedly, trehalose was inert. As to substrate, the present system utilized free palmitic acid rather than palmitoyl-CoA. The reaction products from palmitate and glucose were glucose mycolate and trehalose monomycolate, in which the label from [14C]palmitate or [14C]glucose was incorporated. Glucose palmitate was also formed. Addition of trehalose resulted in a shift from glucose mycolate to trehalose monomycolate. These data clearly indicate that sugars play an important role in the synthesis of mycolic acids from free fatty acids.  相似文献   

20.
The present study was conducted to examine esterification of retinol by testicular microsomes. The microsomes were isolated from rat testes and were incubated under varying assay conditions with [3H]retinol. [3H]Retinylpalmitate was identified by reversed-phase high-performance liquid chromatography as an esterified product. The rate of esterification was increased by the addition of a fatty acyl-CoA. Coenzyme A esters of oleic, palmitic and stearic acids were equally effective substrates for retinol esterification. A 17-fold increase was observed in the presence of palmitoyl-CoA when microsomes had been pretreated with hydroxylamine, a reagent that reacts with coenzyme A thioesters to form hydroxamic acids. The esterifying activity was stimulated by the addition of dithiothreitol (4 mM) and fatty acid-free bovine serum albumin (1 mg/ml). The optimal concentrations for retinol and palmitoyl-CoA were 40 microM and 30-40 microM, respectively. The enzyme activity was inhibited by p-hydroxymercuribenzoate, sodium taurocholate and 5,5'-dithiobis-(2-nitrobenzoic acid), but not by EDTA. The enzyme activity was highest in microsomes (36%). However, some activity was present in mitochondria (29%). These results clearly show the presence of a fatty acyl-CoA: retinol acyltransferase that catalyzes the esterification of retinol in rat testes.  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