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.
The metabolism of two new 3-methyl-branched iodovinyl fatty acids in rat hearts was evaluated by determining the subcellular and lipid pool distribution of these radiolabeled analogues after intravenous injection. Methyl branching had been introduced into the straight chain analogue, 19-iodo-18-nonadecenoic acid (IVN), to produce the monomethyl analogue, 19-iodo-3-(R,S)-methyl-18-nonadecenoic acid (BMIVN) and the dimethyl derivative, 19-iodo-3,3-dimethyl-18-nonadecenoic acid (DMIVN) in the hope of inhibiting oxidation. Since the presence of 3-methyl branching results in delayed myocardial clearance in rats, differences were sought in the lipid and subcellular distribution of these branched analogues that might correlate with the prolonged retention and reflect differences in metabolism. Hearts of rats injected intravenously with the radiolabeled fatty acids were removed and homogenized and the homogenates partitioned between the chloroform-methanol (organic) fraction and the aqueous fraction. Comparison of the distribution of radioactivity between the organic and aqueous fractions showed that most of the DMIVN and BMIVN activity was in the organic fraction with IVN activity initially divided equally between the two fractions. Identification of the lipid components of these organic fractions showed that there was slow incorporation of DMIVN into the triglyceride and polar lipid fractions with a slow loss from the free fatty acid fraction. With the straight chain IVN analogue which shows rapid washout from rat hearts, there was loss of activity from all 3 lipid components during the 60 min. The monomethyl branched BMIVN analogue demonstrated predominant storage in the polar lipid fraction with some incorporation into triglycerides. Subcellular distribution studies of the three analogues also showed differences that correlated with the observed differences in heart retention properties. With the unbranched IVN analogue, radioactivity was found primarily in the cytoplasmic fraction 30 min after injection, whereas the branched analogues demonstrated a much higher association with the microsomal and mitochondrial fractions of the heart. In rats fed prior to injection, these differences in the subcellular distribution profiles were minimized. The lipid and subcellular distribution patterns reported here for the methyl branched analogues as compared to those of the straight chain iodovinyl fatty acid may provide some understanding as to the mechanisms of retention in rat myocardium.Research supported by the Office of Health and Environmental Research, U.S. Department of Energy, under contract DE-AC0 5-840 R21400 with Martin Marietta Energy Systems, Inc.  相似文献   

2.
3.
4.
The intracellular and subcellular distribution of 16-(123I)-iodo-9-hexadecenoic acid were studied in isolated rat hearts, perfused with or without glucose. At various time intervals after injection, cardiac lipids were extracted and the activity was determined for all fractions and all lipid classes. The total cardiac activity was maximal within 1 min postinjection and most of the activity was in the aqueous phase. The presence of glucose in the perfusion medium induced an increase of total cardiac and organic fraction activities. In the latter fraction, activity was very low for FFA, but high for triglycerides (TG), and especially polar lipids. The presence of an exogenous substrate, led to a more active esterification of fatty acids. Coronary effluent analysis showed, in the hydrophilic phase, a lower activity spike in the presence than in the absence of glucose. In the mitochondrial fraction most activity occurred in the organic phase, especially as polar lipids. In the nonmitochondrial fraction, activity was much higher in the aqueous phase. At 90 s postinjection of 1-14C-palmitic acid, over 80% of the myocardial activity was found in the hydrophilic fraction, which indicates, as for the iodo-fatty acid (IFA), an immediate and important oxidation, especially without glucose. These data seem to prove that IFA is taken up by the myocardial cell, subsequently enters the mitochondria and, without an early deiodination, is oxidized with iodide release. Changes in IFA metabolism, consecutive to modifications of glucose concentration in the perfusion medium can be observed by external detection of the myocardial activity curve. Omega-Iodinated fatty acids do not undergo a nonspecific deiodination and are therefore well suited for an external study of myocardial metabolism.  相似文献   

5.
The methyl-branched fatty acids, if radioiodine labelled in alpha position, are potentially adapted to a selective study of FA myocardial uptake. To determine the position and the number of methyl radicals that are necessary to obtain a maximal uptake and a minimal degradation, we measured time-activity evolution of isolated and perfused rat hearts after an injection of iodinated fatty acids which are mono- or dimethylated in alpha or beta position. Except for dimethyl fatty acid, the uptake is similar for all fatty acids studied to that of the straight chain analogue; beta mono- or dimethyl fatty acids seem best adapted to a study of the uptake because alpha monomethyl fatty acids undergo a metabolic degradation and alpha mono- and dimethyl fatty acids induce ventricular fibrillations.  相似文献   

6.
Interpretation of dynamic and integrated myocardial tomograms requires elucidation of the biochemical fate of the tracer and characterization of its tissue distribution and rate of efflux. The fate of [1-11C] and [1-14C]palmitate was studied in 13 open-chest dogs during control or ischemic extracorporeal perfusion of the left circumflex coronary artery. Residue detection of myocardial radioactivity, and radio-biochemical analyses of sequential transmural biopsies and arterial and coronary venous effluent were performed for 30 min after intracoronary bolus administration of tracer. In control hearts, 10.3% of initially extracted tracer was retained in tissue (2.9% in triglyceride, 3.5% in phospholipid, and 3.9% in other lipid and aqueous fractions), 73.7% was oxidized, and 16.1% back-diffused unaltered. With ischemia (pump flow 10% of normal), 28.1% was retained (18% in triglyceride, 6.0% in phospholipid, and 4.1% in other lipid and aqueous fractions), 27.2% was oxidized, and 44.4% back diffused (p less than 0.05 compared to control). Throughout the 30-min study interval, triglyceride, diglyceride, and nonesterified fatty acid comprised a significantly greater fraction of initially extracted radioactivity in ischemic than in control hearts. Thus, during ischemia externally detected clearance rates cannot be used as a direct measure of fatty acid metabolism because of marked influences on efflux of nonmetabolized radiolabeled palmitate and the distribution of tracer retained in tissue. Quantitative measurements of specific metabolic processes by tomography will require development and validation of tracers confined to individual metabolic pathways or pools.  相似文献   

7.
Subcellular distribution studies of two aryl branched-chain and one aryl straight-chain iodinated fatty acids were carried out as part of a continuing effort to determine if such acids are metabolized by beta-oxidation. For the omega-iodoaryl fatty acids, a change in subcellular radioactivity location was observed which was chain-length dependent. Chain lengths of 15 carbons, straight and branched, were largely found in the nuclear-membrane fraction, whereas a chain length of 8 carbons was largely located in the cytosol. No unequivocal evidence for metabolic trapping in the mitochondria was observed for omega-iodoaryl branched-chain fatty acids using the centrifugation technique employed in this study.  相似文献   

8.
Metabolism of methyl-branched iodo palmitic acids in cultured hepatocytes   总被引:1,自引:0,他引:1  
The metabolic fate of methyl-branched iodo fatty acids was studied in primary culture of rat hepatocytes. We compared 16-iodo-2-R,S-methyl palmitic acid (2-Me), which can be beta oxidized, with 16-iodo-3-R,S-methyl palmitic acid (3-Me) which can be beta oxidized only after an initial alpha oxydation and with 16-iodo-2,2-dimethyl palmitic acid (2,2-Me2) and 16-iodo-3,3-dimethyl palmitic acid (3,3-Me2) which cannot be beta oxidized at all. The normal fate of natural fatty acids was given by comparative experiments with [1-14C] palmitic acid. Monomethyl-branched iodo fatty acids were taken up in the same range as palmitic acid but more than dimethyl-branched iodo fatty acids. After a 15-h incubation, acido-soluble products (ASP) accounted for 75% of the radioactivity taken up as 16-iodo-2-methyl palmitic acid, 50% as other methyl-branched iodo fatty acids and only 30% as palmitic acid, which indicated that all the methyl-branched iodo fatty acids underwent a strong deiodination process. Fatty acids were esterified in the following order: palmitic acid greater than 16-iodo-3-R,S-methyl palmitic acid greater than 16-iodo-2-R,S-methyl palmitic acid greater than 16-iodo-2,2-dimethyl palmitic acid greater than 16-iodo-3,3-dimethyl palmitic acid. Cultured hepatocytes, labelled for 3 h with the various fatty acids and reincubated for 12 h without fatty acid, secreted large amounts of free dimethyl-branched iodo fatty acids as compared to the monomethyl ones and palmitic acid.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

9.
It is recognized that iodine-123-labelled 15-(p-iodophenyl)-3-(R,S)-methylpentadecanoic acid (123IBMIPP) slowly washes out of the myocardium. The mechanism for the washout was investigated in normal rat hearts by analyses of the subcellular distribution and lipid classes based on the BMIPP metabolism. Rat hearts were excised at 1–120 min after intravenous injection of123I-BMIPP. After counting the radioactivity, the hearts were digested with Nagarse and homogenized, and then fractionated into the cytosolic, mitochondrial, microsomal and crude nuclear fractions by centrifugations. The radioactivity of each fraction was counted, and the lipid classes were analysed by radio-thin-layer chromatographic and high-performance liquid chromatographic methods. The heart uptake of 1231-BMIPP was maximal at 5 min (6.81%±0.36% ID/g), and 41% of the radioactivity disappeared within 120 min. The myocardial radioactivity was immediately distributed into the cytosolic, mitochondrial, microsomal and crude nuclear fractions. The distribution (%) of each fraction was almost identical from 5 min through 120 min. The cytosolic fraction was always the major site of radioactivity deposition (60%), and the time-activity curve of the cytosolic fraction paralleled that of the whole heart throughout the 120-min study period. In the cytosolic fraction, most of the radioactivity was incorporated into the triglyceride class, and the rest was present in the free fatty acid, phospholipid (phosphatidylcholine) and diglyceride classes. In the mitochondrial fraction, the radioactivity was mostly incorporated into the phospholipid class (phosphatidylethanolamine), followed by free fatty acids. The final metabolite of123I-BMIPP,123I-p-iodophenylacetic acid (123I-PIPA), initially appeared in the mitochondrial fraction as early as 1 min, and subsequently in the cytosolic fraction at 5 min. Another intermediary metabolite,123I-p-iodophenyldodecanoic acid (123I-PIPC12), was found only in the mitochondrial fraction after 5 min. In conclusion, the slow washout kinetics of123I-BMIPP from the myocardium mainly reflects the turnover rate of the triglyceride pool in the cytosol. The BMIPP metabolism, i.e. initial -oxidation followed by subsequent cycles of -oxidation, was confirmed in vivo. The participation of the mitochondria in the metabolism was also proven.  相似文献   

10.

The effects of 3-methyl substitution on the heart retention and metabolism of 3-R,S-methyl-(BMIPP) and 3,3-dimethyl-(DMIPP) analogues of 15-(p-iodophenyl)-pentadecanoic acid (IPP) were studied in rats. Methyl substitution considerably increased the myocardial half-time values in fasted rats: IPP, 5–10 min; BMIPP, 30–45 min; DMIPP, 6–7 h. Because of the observed differences in the relative myocardial uptake and retention of these agents, an evaluation of the subcellular distribution profiles and the distribution of radioactivity within various lipid pools extracted from cell components was performed. Studies with DMIPP in food-deprived rats have shown high levels of the free fatty acid and only slow conversion to triglycerides. These data are in contrast to the rapid clearance of the straight chain IPP analogue and rapid incorporation into triglycerides, and suggest that the prolonged myocardial retention observed with DMIPP in vivo may result from inhibition of β oxidation. Subcellular distribution studies have shown predominant association of DMIPP and BMIPP with the mitochondrial and microsomal fractions, while IPP was primarily found in the cytoplasm. Because of the unique “trapping” properties and the high heart: blood ratios, [123I]DMIPP should be useful for evaluation of aberrations in regional myocardial uptake.

  相似文献   

11.
The metabolic fate of methyl-branched iodo fatty acids was studied in primary culture of rat hepatocytes. We compared 16-iodo-2-R,S-methyl palmitic acid (2-Me), which can be oxidized, with 16-iodo-3-R,S-methyl palmitic acid (3-Me) which can be oxidized only after an initial oxydation and with 16-iodo-2,2-dimethyl palmitic acid (2,2-Me2) and 16-iodo-3,3-dimethyl palmitic acid (3,3-Me2) which cannot be oxidized at all. The normal fate of natural fatty acids was given by comparative experiments with [1-14C] palmitic acid. Monomethyl-branched iodo fatty acids were taken up in the same range as palmitic acid but more than dimethyl-branched iodo fatty acids. After a 15-h incubation, acido-soluble products (ASP) accounted for 75% of the radioactivity taken up as 16-iodo-2-methyl palmitic acid, 50% as other methyl-branched iodo fatty acids and only 30% as palmitic acid, which indicated that all the methyl-branched iodo fatty acids underwent a strong deiodination process. Fatty acids were esterified in the following order: palmitic acid >16-iodo-3-R,S-methyl palmitic acid>16-iodo-2-R,S-methyl palmitic acid>16-iodo-2,2-dimethyl palmitic acid>16-iodo-3,3-dimethyl palmitic acid. Cultured hepatocytes, labelled for 3 h with the various fatty acids and reincubated for 12 h without fatty acid, secreted large amounts of free dimethylbranched iodo fatty acids as compared to the monomethyl ones and palmitic acid. Only hepatocytes prelabelled with 16-[125I]iodo-2,2-dimethyl palmitic acid exhibited an appreciable secretion of labeled triglycerides, but at a lower rate than with [1-14C] palmitic acid. Conversely, the 16-iodo-monomethyl palmitic acids remained chiefly in hepatocyte triglycerides. Minute amounts of 16-iodo-methyl-branched-palmitic acids were found in hepatocyte or secred phospholipids as compared with palmitic acid. This metabolic fate of methyl-branched iodo palmitic acids argues against their utilization as imaging probes to monitor in vivo the synthesis and the secretion of triglycerides by the liver.  相似文献   

12.
Myocardial metabolism of radioiodinated methyl-branched fatty acids   总被引:3,自引:0,他引:3  
Methylated fatty acids labeled with radioactive iodine have been proposed as a means of studying regional myocardial uptake of fatty acids in man. To investigate the methylated fatty acid that is best adapted for an assessment of uptake, we have studied the influence of the number and the position of the methyl groups of IFA intracellular metabolism; 16-iodo-2-methyl-hexadecanoic (mono-alpha), 16-iodo-2,2-methyl hexadecanoic (di-alpha), 16-iodo-3-methyl-hexadecanoic (mono-beta), and 16-iodo-3,3-methyl-hexadecanoic (di-beta) acids were injected into the coronary arteries of isolated rat hearts. Intracellular analysis shows that the degradation of mono-alpha was always lower than that of IHA and the storage was always much higher. The differences between mono-beta and IHA were similar to those observed with mono-alpha, but were much more pronounced. With the two dimethylated IFAs there was an inhibition of both oxidation and esterification which led to an accumulation of free FAs in myocardial cells. In conclusion, mono-beta, di-alpha, and di-beta are potentially suitable for studying the cellular uptake of IFA since all of them, and particularly the dimethylated IFAs, have a low oxidation rate.  相似文献   

13.
Twenty-eight dogs with acute anterior myocardial infarcts due to proximal occlusion of the left anterior descending coronary artery (LAD) were studied at various periods following the occlusion to determine: (a) the time course and location of abnormal lipid accumulation after infarction, (b) the degree of muscle-cell injury associated with increased lipid deposition, and (c) whether uptake of fatty acid from the circulating fat pool contributes to lipid accumulation in certain myocardial regions. The findings show that myocardial lipid accumulation begins as early as 6 hr after proximal LAD occulsion. The increased lipid deposition occurs as nonmembrane-bound lipid droplets in muscle cells with and without ultrastructural evidence of irreversible injury. Analysis of tissue uptake of intravenoulsy injected [14C] oleic acid conjugated with albumin revealed relatively selective concentration of label in the peripheral and border regions of the infarct, but occasionally even the central subendocardial portion of the infarct concentrated the fatty acid. Thin-layer chromotography showed that most of the label was associated with the triglyceride fraction when the radiolabeled fatty acid was injected 6 or 24 hr after LAD occlusion. These myocardial cellular and topographical alterations will have to be considered when labeled fatty acids are used for imaging acute myocardial infarcts and/or if attempts are made to identify myocardial fat-laden cells scintigraphically.  相似文献   

14.
IntroductionOur group has synthesized technetium-labeled fatty acids (FA) that are extracted into the myocardium and sequestered due to heart-type fatty acid binding protein (H-FABP) binding. In this article, we further address the detailed subcellular distribution and potential myocardial metabolism of [99mTc]“4+1” FA.MethodsExperiments were conducted using isolated hearts of Wistar rats, as well as of wild-type and H-FABP?/? mice. Myocardium samples underwent subcellular fractionation [subsarcolemmal mitochondria (SM), intermyofibrillar mitochondria (IM), cytosol with microsomes, and nuclei and crude membranes] and analysis by thin-layer chromatography and high-performance liquid chromatography.ResultsThe largest fraction of tissue radioactivity was associated with cytosol [79.69±8.88% of infused dose]. About 9.07±0.95% and 3.43±1.38% of the infused dose were associated with SM and IM fractions, respectively. In the rat heart, etomoxir, an inhibitor of carnitin-palmitoyl transferase I, did not significantly decrease radioactivity associated with mitochondrial fractions, whereas myocardial extraction of [123I]-labeled 15-(p-iodophenyl)-pentadecanoic acid (13.26% vs. 49.49% in controls) and the radioactivity associated with the SM and IM fractions were blunted. The percentage of the infused dose in the mitochondrial and crude fractions increased with the number of NH-amide groups of the FA derivative. Absence of H-FABP significantly decreased radioactivity count in the cytosolic fraction (P<.001). No metabolic product of [99mTc]“4+1” FA could be detected in any isolated heart.ConclusionsMyocardial [99mTc]“4+1” FA extraction reflects binding to H-FABP and membrane structures (including the mitochondrial membrane). However, the compounds do not undergo mitochondrial metabolism because they do not reach the mitochondrial matrix.  相似文献   

15.
Biodistribution studies with the radioiodinated 3(R)- and 3(S)-isomers of 15-(p-iodophenyl)-3-methylpentadecanoic acid (BMIPP) in rats have shown that 3(R)-BMIPP has 20%-25% higher heart uptake than 3(S)-BMIPP (15-180 min). In contrast, the 3(S)-isomer has slightly higher liver uptake, and uptake in other tissues examined is similar. METHODS: To evaluate the possible differences in metabolic fate of the two isomers, a mixture of [125I]-3(R)/[131I]-3(S)-BMIPP was administered to fasted female Fisher rats. Groups of rats (3 per group) were killed 15, 60 and 180 min after administration. Urine and feces were collected from a fourth group (n = 3) over 7 d. Samples of blood, heart, liver, lungs, kidney and urine were Folch extracted. The distributions of 125I and 131I in the organic (lipid), aqueous and pellet samples were determined. The lipid samples as well as the organic fractions from base-hydrolyzed triglyceride (TG) fractions and acid-hydrolyzed urine samples were then analyzed by thin-layer chromatography (TLC) and high-performance liquid chromatography (HPLC). RESULTS: The relative distributions of 125I and 131I in the lipid, aqueous and pellet samples were similar for both isomers. Distribution of 125I and 131I in the various components of the lipid extracts observed by TLC (hexane:ether:HOAc, 70:30:1) was also similar, with principal incorporation into the free fatty acid (FFA) and TG pools. HPLC analyses (C18) of the FFA fraction showed similar 125I and 131I profiles, corresponding to BMIPP, and the alpha-methyl-C14 (14-(p-iodophenyl)-3-(R,S)-methyltetradecanoic acid) and C12, C10 and C6 carbon chain-length catabolites. By TLC, radioactive components of 125I and 131I in the urine had the same TLC mobility as hippuric acid. HPLC analyses (C18) of acid-hydrolyzed urine gave a single 125I/131I component with the same relative retention time as 2-(p-iodophenyl)acetic acid, which is the final alpha/beta-oxidative BMIPP catabolite. Unexpectedly, HPLC of lipids from base-hydrolyzed TG from the heart tissue showed 125I/131I components with the same retention times as shorter-chain fatty acids, similar to the FFA fraction, with only low levels of activity detected in BMIPP. CONCLUSION: These results show that 3(R)-BMIPP and 3(S)-BMIPP are metabolized similarly in rat tissues and that higher myocardial extraction observed for 3(R)-BMIPP may reflect differences in the relative membrane transport of the two isomers.  相似文献   

16.
Time courses of radioactivity (residue curves) were obtained following bolus injection into working rat hearts of two 125I-labeled long chain fatty acids: 16-iodohexadecanoic acid (IHDA) and 15-p-iodophenylpentadecanoic acid (IPPA). Residue curves were analyzed in terms of a rapid vascular washout component, an early tissue clearance component, and a very slow late component. For IHDA and IPPA in control hearts, early myocardial clearance kinetics were rate limited by the diffusion of catabolites. Sensitivity of the kinetics to impaired fatty acid oxidation was examination by pretreatment of animals with 2[5(4-chlorophenyl)pentyl]oxirane-2-carboxylate (POCA). Decreased fatty acid oxidation was indicated in IHDA and IPPA residue curves by a decrease in the relative size of the early clearance component. Analysis of radiolabeled species in coronary effluent and heart homogenates showed that back diffusion of IPPA was slower than that of IHDA; this discrepancy was most apparent in POCA hearts. In vitro binding assays suggested higher tissue:albumin relative affinity for IPPA than for IHDA. Thus, IPPA early clearance kinetics were more closely related to the clearance of labeled catabolite(s) and were therefore more sensitive to the oxidation rate of long chain fatty acids.  相似文献   

17.
The aim of the present study was to demonstrate that it is possible to estimate the intracellular metabolism of a fatty acid labelled with iodine using external radioactivity measurements. 123I-16-iodo-9-hexadecenoic acid (IHA) was injected close to the coronary arteries of isolated rat hearts perfused according to the Langendorff technique. The time course of the cardiac radioactivity was measured using an INa crystal coupled to an analyser. The obtained curves were analysed using a four-compartment mathematical model, with the compartments corresponding to the vascular-IHA (O), intramyocardial free-IHA (1), esterified-IHA (2) and iodide (3) pools. Curve analysis using this model demonstrated that, as compared to substrate-free perfusion, the presence of glucose (11 mM) increased IHA storage and decreased its oxidation. These changes were enhanced by the presence of insulin. A comparison of these results with measurements of the radioactivity levels within the various cellular fractions validated our proposed mathematical model. Thus, using only a mathematical analysis of a cardiac time-activity curve, it is possible to obtain quantitative information about IHA distribution in the different intracellular metabolic pathways. This technique is potentially useful for the study of metabolic effects of ischaemia or anoxia, as well as for the study of the influence of various substrates or drugs on IHA metabolism in isolated rat hearts.  相似文献   

18.
These studies evaluated the kinetics of tracer uptake and washout after step-function labeling with 14C-palmitate. Washout and uptake function curve analysis for total radioactivity (TR) was derived according to the expressions: TR = Fx integral of 0 infinity C(t) x dt and TR = Fx integral of 0 infinity (Css - C(t)) x dt, respectively, with Vc = TR/Ca, where F = coronary flow; Css = steady-state concentration; C(t) = concentration with respect to time; Ca = arterial concentration; and Vc = distribution volumes within the fatty acid pathway. The only radioactive metabolites in venous effluent were fatty acids and 14CO2. The estimated Vc of fatty acids was small (1.2-1.7 ml/g dry wt or 0.4-0.5 mumol/g dry wt) and compatible with labeled substrate trapped in the blood volume. The Vc of 14CO2 was much larger (11.4-15.8 ml/g dry wt or 3.6-4.2 mumol/g dry wt) and correlated with counts contained in the aqueous soluble and fatty acid fractions in tissue. The counts in tissue were distributed between the aqueous soluble fraction (40%), which was rapidly depleted during washout, and a lipid fraction (60%) (triacylglycerols and phospholipids), which was resistant to washout. Distributions in tissue radioactivity between the aqueous soluble and lipid fractions support the notion of a dual pathway in fatty acid oxidation, one arm of which passes through the resident pool of triacylglycerols, which has a long time constant. The presence of this pool may impart an error in estimating fatty acid oxidation by external labeling techniques.  相似文献   

19.
Time courses of radioactivity (residue curves) were obtained following bolus injection into working rat hearts of two 125I-labeled long chain fatty acids: 16-iodohexadecanoic acid (IHDA) and 15-p-iodophenylpentadecanoic acid (IPPA). Residue curves were analyzed in terms of a rapid vascular washout component, an early tissue clearance component, and a very slow late component. For IHDA and IPPA in control hearts, early myocardial clearance kinetics were rate limited by the diffusion of catabolites. Sensitivity of the kinetics to impaired fatty acid oxidation was examination by pretreatment of animals with 2[5(4-chlorophenyl)pentyl]oxirane-2-carboxylate (POCA). Decreased fatty acid oxidation was indicated in IHDA and IPPA residue curves by a decrease in the relative size of the early clearance component. Analysis of radiolabeled species in coronary effluent and heart homogenates showed that back diffusion of IPPA was slower than that of IHDA; this discrepancy was most apparent in POCA hearts. In vitro binding assays suggested higher tissue: albumin relative affinity for IPPA than for IHDA. Thus, IPPA early clearance kinetics were more closely related to the clearance of labeled catabolite(s) and were therefore more sensitive to the oxidation rate of long chain fatty acids.  相似文献   

20.
The lipid pools of the heart (i.e. the triglyceride and phospholipid pool) participate in the free fatty acid metabolism. The degree of involvement, for instance will be determined by the available substrate in the blood. Scintigraphy with 17-iodo-123 heptadecanoic acid was performed to study free fatty acid kinetics in the normal human myocardium during control and glucose infusion (n = 9). In both situations the derived time-activity curves, measured during a period of 75 min, obeyed a monoexponential plus a constant curve fitting [A(t) = A(o)exp(-tln2/T1/2)+C]. During glucose infusion the half-time values did not change but the lipid storage increased in favour of the oxidation pool. In the three protocols used, hypoglycaemic responses were observed, therefore these protocols cannot be advocated.  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