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.
This article presents two cases with preserved myocardial 201Tl uptake and absent uptake of two kinds of radioiodinated fatty acids: iodine-123-labeled 15-(p-iodophenyl)-3-(R,S)-methylpentadecanoic acid (BMIPP) and iodine-123-labeled 15-(p-iodophenyl)-9-(R,S)-methylpentadecanoic acid (9MPA). Although coronary angiography showed no stenotic lesion and left ventriculography revealed no wall motion abnormality, no myocardial uptake of BMIPP and 9MPA was observed in the first case. In the second case, no myocardial accumulation was recognized even in the initial phase of dynamic SPECT acquired soon after the injection of 9MPA. The results suggest that the non-visualized myocardium was not specific for BMIPP imaging and that rather than the early back diffusion of the tracers from the myocardium, abnormality of the myocardial cell membrane was a possible mechanism accounting for the phenomenon.  相似文献   

2.
In the final reaction of peroxisomal alpha-oxidation of 3-methyl-branched fatty acids a 2-hydroxy-3-methylacyl-CoA intermediate is cleaved to formyl-CoA and a hitherto unidentified product. The release of formyl-CoA suggests that the unidentified product may be a fatty aldehyde. When purified rat liver peroxisomes were incubated with 2-hydroxy-3-methylhexadecanoyl-CoA 2-methylpentadecanal was indeed formed. The production rates of formyl-CoA (measured as formate) and of the aldehyde were in the same range. While the production of formate remained unaltered in the presence of NAD+, the amount of 2-methylpentadecanal was decreased, which was accompanied by the formation of 2-methylpentadecanoic acid. These data indicate that (1) during alpha-oxidation the 2-hydroxy-3-methylacyl-CoA is cleaved to a 2-methyl-branched aldehyde and formyl-CoA and (2) liver peroxisomes are capable of converting this aldehyde to a 2-methyl-branched fatty acid.  相似文献   

3.
Little information is available regarding the determinants of systolic contractile function of the hypertrophied right ventricle (RV). The purpose of this study was to clarify the relationship between myocardial metabolism and contractile function in the hypertrophied RV due to pulmonary hypertension (PH). METHODS: Iodine-123-labeled 15-(p-iodophenyl)-3-(R,S)-methylpentadecanoic acid (BMIPP) and 99mTc-sestamibi (MIBI) SPECT were performed to calculate the RV-to-left ventricle (LV) tracer uptake ratio (RV/LV) in 21 patients with PH (6 with primary PH and 15 with chronic thromboembolic PH). The patients also underwent electron-beam CT to assess RV ejection function (RVEF) and percentage systolic wall thickening (%SWT) and right heart catheterization to measure mean pulmonary arterial pressure (mPAP). RESULTS: There were significant positive correlations between mPAP and MIBI-RV/LV (r = 0.89, p < 0.001) and between mPAP and BMIPP-RV/LV (r = 0.86, p < 0.001). However, 8 patients showed lower BMIPP-RV/LV than MIBI-RV/LV, indicating the impairment of myocardial fatty acid uptake in the RV. These patients had lower RVEF and %SWT compared to those with normal myocardial fatty acid uptake (RVEF = 28% +/- 10% compared to 40% +/- 9% and %SWT = 33% +/- 27% compared to 74% +/- 30%, respectively; p < 0.05 for both comparisons). Although mPAP did not differ between the groups, the RVEF-mPAP and %SWT-mPAP regression lines drawn from the patients with impaired myocardial fatty acid uptake were located below the lines from the patients with normal myocardial fatty acid uptake, suggesting disproportionately decreased RV myocardial contractility for a given mPAP in patients with impaired myocardial fatty acid uptake. The patients with the impaired fatty acid uptake in the RV had a significantly higher death rate (log-rank test, p < 0.05). CONCLUSION: The results from this preliminary study suggest that myocardial fatty acid uptake is impaired in the failing hypertrophied RV due to PH.  相似文献   

4.
To study the sensitivity of two fatty acid tracers to changes in beta-oxidation, the myocardial retention kinetics of 125I-iodine-15-(p-iodophenyl)-3(R,S)-methylpentadecanoic acid (BMIPP) and 14-18F-fluoro-6-thia-heptadecanoic acid (FTHA) were compared in states of oxygen deprivation due to ischemia and hypoxia. METHODS: Nineteen swine were studied by extracorporeal perfusion of the three coronary arteries. Fatty acid beta-oxidation rates were determined by infusion of tritiated palmitate into the left anterior descending artery (LAD) and by measurement of labeled water production in the LAD perfusion bed. After a baseline period of 30 min, animals were divided into three groups and subjected to a 50-min intervention period. For the control group, there was no change in perfusion; for the ischemia group, there was a 60% decrease in LAD perfusion; and for the hypoxia group, the perfusion rate was unchanged, but venous blood was used as the LAD perfusate. Continuous infusion of FTHA and BMIPP into the LAD started 10 min into the intervention period and continued until the end of the intervention period. Retention rates of the two tracers were compared between the LAD and circumflex perfusion beds. RESULTS: No difference in beta-oxidation rate occurred from the baseline to the intervention period in the control group. A 50% reduction in beta-oxidation occurred in the ischemia group, and an 80% reduction occurred in the hypoxia group. No difference in retention of BMIPP or FTHA occurred in the control group. In the ischemia group, reduction in retention of both tracers occurred. However, in the hypoxia group, FTHA uptake was unchanged, whereas BMIPP retention increased compared to the circumflex arterial bed. CONCLUSION: Decreased retention of both BMIPP and FTHA occurred with ischemia, despite the known differences in metabolism of the two tracers. This difference in metabolism was further highlighted in the setting of hypoxia with increased BMIPP uptake. Thus, these results suggest that uptake of both FTHA and BMIPP tracks reduction of fatty acid utilization in myocardial ischemia but fails in tracking reduction of fatty acid oxidation during hypoxia.  相似文献   

5.
To improve the image quality in simultaneous dual-isotope single-photon emission tomography (SPET) with iodine-123 labelled 15-(p-iodophenyl)-3-methylpentadecanoic acid (BMIPP) and thallium-201, we applied the triple-energy window method (TEW) for correction of the cross-talk and scatter artifact. Seventy-one patients with coronary artery disease were included. 201Tl cross-talk into the 123I acquisition window (group 1, n = 30) and 123I cross-talk into the 201Tl window (group 2, n = 41) were studied. In group 1, 123I images were first obtained (single-isotope images), followed by 201Tl injection and SPET acquisition using dual-isotope windows (dual-isotope images). In group 2, the order was reversed. The dual-isotope SPET images with and without TEW were compared with the single-isotope images. Qualitative evaluation was performed by scoring the segmental defect pattern. Detectability of the mismatched fatty acid metabolism on dual-isotope SPET was evaluated by receiver operating characteristic (ROC) curve analysis. Segmental defect pattern agreement between dual and corrected single images was significantly improved by TEW correction (P<0.01). The agreement was particularly improved in segments with absence of uptake. There was no significant difference between TEW-corrected dual-isotope SPET and corresponding single-isotope SPET with regard to either % defect count or background activity. Mismatched fatty acid metabolism depicted by dual-isotope SPET predicted abnormal wall motion more accurately with TEW than without TEW. With TEW, a practical method for scatter and cross-talk correction in clinical settings, simultaneous dual 123I-BMIPP/201Tl SPET is feasible for the assessment of myocardial perfusion/metabolism mismatch.  相似文献   

6.
Radioiodinated fatty acid analogues modified by methyl-substitution are used for single photon emission tomography (SPET) imaging of the heart. The effect of mono- and dimethyl-substitution on heart-to-organ ratios was investigated in humans to evaluate their relative merits for SPET image quality. Planar total body scans were performed in fasting patients with coronary artery disease, but without heart failure, 1 h after administration of 111 MBq 15-(p-[I-123]-iodophenyl)-3-(R,S)-methylpentadecanoic acid (BMIPP, n = 7) or 111 MBq 15-(p-[I-123]-iodophenyl)-3,3-dimethylpentadecanoic acid (DMIPP, n = 4). Because these branched fatty acids are used for cardiac imaging, we focused on heart-to-organ (heart/organ) ratios by comparing small regions of interest in heart, liver, lung, muscle and bladder. Both tracers showed good visualization of the heart. DMIPP showed a relatively high liver uptake: the heart/liver ratios for DMIPP and BMIPP were 0.39 +/- 0.05 and 1.00 +/- 0.12, respectively (P < 0.0001). Increased lung activity was found for BMIPP, with a heart/lung ratio of 1.63 +/- 0.17 versus 2.32 +/- 0.28 for DMIPP (P < 0.001). In contrast to DMIPP, BMIPP also showed increased activity in the bladder. In conclusion, BMIPP and DMIPP show different distribution patterns. Despite the more favourable heart/lung ratios for DMIPP, the high liver uptake affects cardiac SPET image quality and therefore BMIPP appears to provide superior cardiac SPET image quality in humans.  相似文献   

7.
The nitrite-oxidizing autotroph, Nitrobacter was studied with respect to fatty acid composition and lipids. One fatty acid, shown to be cis-11-actadecenoic acid (cis-vaccenic) accounted for almost 96% of the total fatty acids of the extractable lipids of Nitrobacter agilis, Nitrobacter winogradskyi and each of several isolates from Minnesota and Moroccan soils studied. The cis-vaccenic acid was high in all organisms, ranging from 85 to 95% when grown at 27degreesC in the log growth phase, the other major acid was plamitic (16 : 1). All converted ces-vaccenic acid to a 19-carbon cyclopropanecarboxylic acid upon entering the stationary growth phase. The 11-carbon cyclopropanecarboxylic acid was not degraded when stationary phase cells were reinoculated into fresh medium. In N. agilis the levels of cis-vaccenic acid ranged from 86.8% when grown at 33degreesC to 95.6% when grown at 19degreesC. Addition of acetate or casein hydrolyzate to the inorganic medium had virtually no effect on the fatty acid composition of N. agilis, while propionate effected both qualitative and quantitative changes. In all organisms phosphatidylcholine made up a large portion of the extractable lipids. The distribution was phosphatidylcholine, 54%; phosphatidylethanolamine, 23%; phosphatidylglycerol, 10%; and neutral lipids, 11% for N. agilis.  相似文献   

8.
It is recognized that iodine-123-labelled 15-(p-iodophenyl)-3-(R,S)-methylpentadecanoic acid (123I-BMIPP) 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 of 123I-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 123I-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 of 123I-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 of 123I-BMIPP from the myocardium mainly reflects the turnover rate of the triglyceride pool in the cytosol. The BMIPP metabolism, i.e. initial alpha-oxidation followed by subsequent cycles of beta-oxidation, was confirmed in vivo. The participation of the mitochondria in the metabolism was also proven.  相似文献   

9.
Effects of fatty acids on translocation of the gamma- and epsilon-subspecies of protein kinase C (PKC) in living cells were investigated using their proteins fused with green fluorescent protein (GFP). gamma-PKC-GFP and epsilon-PKC-GFP predominated in the cytoplasm, but only a small amount of gamma-PKC-GFP was found in the nucleus. Except at a high concentration of linoleic acid, all the fatty acids examined induced the translocation of gamma-PKC-GFP from the cytoplasm to the plasma membrane within 30 s with a return to the cytoplasm in 3 min, but they had no effect on gamma-PKC-GFP in the nucleus. Arachidonic and linoleic acids induced slow translocation of epsilon-PKC-GFP from the cytoplasm to the perinuclear region, whereas the other fatty acids (except for palmitic acid) induced rapid translocation to the plasma membrane. The target site of the slower translocation of epsilon-PKC-GFP by arachidonic acid was identified as the Golgi network. The critical concentration of fatty acid that induced translocation varied among the 11 fatty acids tested. In general, a higher concentration was required to induce the translocation of epsilon-PKC-GFP than that of gamma-PKC-GFP, the exceptions being tridecanoic acid, linoleic acid, and arachidonic acid. Furthermore, arachidonic acid and the diacylglycerol analogue (DiC8) had synergistic effects on the translocation of gamma-PKC-GFP. Simultaneous application of arachidonic acid (25 MicroM) and DiC8 (10 microM) elicited a slow, irreversible translocation of gamma-PKC- GFP from the cytoplasm to the plasma membrane after rapid, reversible translocation, but a single application of arachidonic acid or DiC8 at the same concentration induced no translocation. These findings confirm the involvement of fatty acids in the translocation of gamma- and epsilon-PKC, and they also indicate that each subspecies has a specific targeting mechanism that depends on the extracellular signals and that a combination of intracellular activators alters the target site of PKCs.  相似文献   

10.
The substrate binding site of polyenoic fatty acid isomerase (PFI) has been investigated using a series of alternate substrates and by examination of the pH dependence on the kinetic parameters of PFI with selected substrates. The pH dependence profile of PFI with EPA [(5Z,8Z,11Z,14Z,17Z)-eicosapentaenoic acid] shows the enzyme to be catalytically active over a wide pH range, with activity being optimal below pH 6.0. Analysis of the kinetic parameters of DHA [(4Z,7Z,10Z,13Z,16Z,19Z)-docosahexen oic acid]; adrenic acid [(7Z,10Z,13Z,16Z)-docosatetraenoic acid]; EPA; arachidonic acid [(5Z,8Z,11Z,14Z)-eicosatetraenoic acid]; anandamide (arachidonyl-N-ethanolamide); and eicosatrienoic acid [(5Z,8Z,11Z)-eicosatrienoic acid] demonstrates that substrates possessing omega-3 olefins (DHA and EPA) have the lowest K(m) values (1.9 and 9.6 microM, respectively). EPA and arachidonic acid showed the highest V(max) values (6.0 and 2.8 micromol min(-1) mg(-1), respectively). The twenty carbon omega-9 fatty acid eicosatrienoic acid showed a relatively large K(m) and had a V(max) approximately 20-fold less than EPA. Anandamide, a substrate analog lacking an ionizable carboxylate, showed a K(m) similar to the other omega-6 fatty acids (arachidonic acid and adrenic acid); however, the V(max) was approximately 5-fold lower than arachidonic acid and 8-fold lower than EPA. Moreover, anandamide demonstrated no pH dependency on its kinetic parameters over a range where EPA showed a 27-fold decrease in V/K(m). NMR spectroscopy was used to determine the structure of the product from reaction of PFI with DHA. These data showed the compound to be (4Z,7Z,9E,11E,16Z,19Z)-docosahexenoi c acid. Reaction of PFI with dihomo-gamma-linolenic acid resulted in the development of two products, one with the characteristic chromophore of a conjugated triene, the other with a chromophore characteristic of a conjugated diene. Analysis of the products from these reactions of PFI, in conjunction with the kinetic parameters from the alternate substrates, provides compelling evidence that the enzyme preferentially orients the substrate in the catalytic site with respect to the methyl terminus.  相似文献   

11.
omega-Imidazolyl carboxylic acids (C10-C12) have been used as probes of the active site and catalytic mechanism of the fatty acid hydroxylase P-450 BM3 from Bacillus megaterium. These compounds are the most potent inhibitors of P-450 BM3 yet reported. All are mixed inhibitors, increasing the Km and decreasing the kcat for laurate oxidation. All ligate the P-450 BM3 ferric heme iron, inducing a type II shift in the Soret absorbance band from 419 to 424 nm. Binding to the ferrous form is much weaker. 10-(Imidazolyl)decanoic acid was the best inhibitor (Kic = 0.9 microM, Kiu = 5.7 microM), while 12-(imidazolyl)dodecanoic acid (Kic = 1.35 microM, Kiu = 6.9 microM) was superior to 11-(imidazolyl)undecanoic acid (Kic = 7.5 microM, Kiu = 16 microM). Dissociation constants for binding to oxidized P-450 BM3 heme iron were determined spectrophotometrically as 8 microM (C12 azole) and 27 microM (C11 azole). The binding of 10-(imidazolyl)decanoic acid was too tight for an absolute Kd to be determined spectrophotometrically, but this value is <0.2 microM. The binding of different fatty acids to the enzyme was found to have distinct effects on the Kd for the azoles. Laurate induced tighter binding (Kd for the C12 azole lowered to 4.7 microM), while arachidonate weakened the affinity (Kd increased to 23 microM). Arachidonate diminished the affinity for the C10 azole sufficiently that a Kd could be determined by spectrophotometric titration (11 microM). Affinity for the C12 azole was decreased in active-site-mutants R47G (R47 tethers the fatty acid carboxylate group) and F87Y but increased in mutant F87G-indicating an important role for this residue in determining heme accessibility. The C10 azole binds much more weakly to the spin-state-insensitive F87Y (32. 2 microM), suggesting that the inhibitors may bind preferentially to different conformers of P-450 BM3. NADP+ binding in the reductase also tightened affinity of these inhibitors for P-450 BM3 (Kd for the C12 azole decreased to 2.7 microM), but this effect was not observed for FMN-deficient mutant W574D, suggesting that the interdomain effect of NADP+ on inhibitor binding was mediated via flavin mononucleotide. Resonance Raman spectroscopy indicates that the inhibitors form low-spin complexes with P-450 BM3 and that their binding induces movements of the heme vinyls relative to the ring.  相似文献   

12.
To clarify the role of the sterol regulatory element-binding protein-2 (SREBP-2), we established cell lines in which human SREBP-2(1-481) could be induced by isopropyl-beta-D-thiogalactopyranoside (IPTG). The range of IPTG-induced changes in SREBP-2(1-481) levels in '23-11' cells, one of these cell lines, was almost the same as that of sterol-induced changes in the levels of mature SREBP-2, indicating that IPTG was able to regulate the expression of SREBP-2(1-481) within the normal physiological range in this cell line. Sterols regulate the expression of the LDL receptor, HMG-CoA reductase, squalene synthase and fatty acid synthase in 23-11 cells as they also do in the parental cell line HeLa S3. IPTG increased mRNA levels of the LDL receptor and HMG-CoA reductase but not squalene synthase both in the presence or absence of excess sterols. Fatty acid synthase mRNA was increased 2 h after the IPTG addition in the absence of excess sterol (10% FBS), but was slightly increased 6 h after the IPTG addition in the presence of excess sterols. In the absence of excess sterols, both SREBP-2(1-481) and endogenous mature SREBP-2 exist in the nucleus. This suggests that an increased amount of SREBP-2 over the normal physiological range is required for the regulation of fatty acid synthase. IPTG increased both the surface binding of 125I-LDL and cholesterol biosynthesis from [14C]acetate significantly in a similar time course. In contrast, fatty acid biosynthesis from [14C]acetate was almost unchanged by IPTG during the same incubation period. These results suggest that physiological amounts of SREBP-2 play a key role in the regulation of cholesterol but not fatty acid metabolism.  相似文献   

13.
DNA encoding a single unit of the DvA-1 polyprotein of the parasitic nematode Dictyocaulus viviparus was isolated and the polypeptide ("rDvA-1L") expressed in Escherichia coli, to give a protein showing high binding affinity for fatty acids and retinoids. Fluorescent fatty acid probes show substantial changes in emission spectrum in the presence of rDvA-1L, which can be reversed by fatty acids (oleic, palmitic, stearic, arachidonic) and retinoids, but not by tryptophan, squalene, or cholesterol. Moreover, changes in intrinsic fluorescence of retinol or retinoic acid confirm a retinoid binding activity. Fluorescence titration experiments indicate stoichiometric binding to a single protein site per monomer unit with affinities (Kd) in the range 3 x 10(-8) M for 11-((5-dimethylaminonaphthalene-1-sulfonyl)amino)undecanoic acid, and by competition, 5 x 10(-8) M for oleic acid. The extreme blue shift of bound fluorescent fatty acid suggests an unusually low polarity for the protein binding site. The emission spectrum of the single tryptophan of rDvA-1L indicates that it is deeply buried in a nonpolar environment, and its spectrum is unaffected by ligand binding. Far UV circular dichroism of rDvA-1L reveals a high alpha-helix content (53%). Differential scanning calorimetry studies indicate that rDvA-1L is highly stable (T(m) approximately 98 degrees C), refolding efficiently following thermal denaturation. DvA-1 therefore represents an example of a new class of lipid binding protein, and is the first product of a polyprotein with this activity to be described.  相似文献   

14.
Recent studies from our laboratory have shown that methyl palmoxirate (MEP), an inhibitor of mitochondrial beta-oxidation of long chain fatty acids, can be used to increase incorporation of radiolabeled palmitic acid into brain lipids and reduce beta-oxidation of the fatty acid. Thus, MEP allows the use of carbon labeled palmitate for studying brain lipid metabolism in animals and humans by quantitative autoradiography or positron emission tomography (PET). As it is essential to pretreat human subjects with an acute dose of MEP prior to intravenous injection of [1-11C]palmitate for PET scanning, this study was undertaken to determine the plasma elimination half-life of MEP in rats and human subjects and to provide insight about the drug's absorption and metabolism. A gas chromatographic method was developed to measure MEP in body fluids. Following oral administration of MEP to rats (2.5 and 10 mg/kg) and to humans, the unmetabolized drug could not be detected in plasma or urine (sensitivity of detection was 1 ng). However, when MEP was injected intravenously (10 mg/kg) in rats, a peak initial concentration could be measured in plasma (7.7 microg/mL), the clearance of the drug from plasma was rapid (t1/2 = 0.6 min), which indicates that MEP readily enters tissue lipid pools or is metabolized like long-chain fatty acids. As no adverse experience occured in the 11 human subjects studied, oral administration of a single dose of MEP was safe under the conditions of this study and may be used to increase the incorporation of positron labeled palmitic acid for studying brain lipid metabolism in vivo by PET.  相似文献   

15.
PHIPA 3-10 [13-(4'-iodophenyl)-3-(p-phenylene)tridecanoic acid] is a p-phenylene-bridged, radioiodinated omega-phenyl fatty acid that has recently been developed to study coronary artery disease or cardiomyopathies. Here, we demonstrate that PHIPA 3-10 exhibits the characteristics of a long-chain fatty acid, including its ability to be efficiently taken up by myocytes and to function as a substrate for beta-oxidation before it is trapped. METHODS: Myocardial metabolism of carrier-added and carrier-free 131I-PHIPA 3-10 preparations were investigated in rats in vivo and in isolated Langendorff rat hearts. Heart extracts were analyzed by high-performance liquid chromatography, negative-ion electrospray mass spectrometry and investigation of intracellular distribution using density-gradient centrifugation. RESULTS: A single, rapidly formed metabolite was found in the heart extract and also, surprisingly, in the hydrolyzed lipids. The total amount of metabolite increased from 43% to 51% between 15 and 60 min postinjection. By high-performance liquid chromatography comparison with synthetic potential catabolites, the metabolite was assigned the name PHIPA 1-10 [11-(4'-iodophenyl)-1-(p-phenylene)undecanoic acid] and was the product of one beta-oxidation cycle. Additional proof was obtained from the mass spectrometric analysis of the metabolite formed in vivo. The formation of this metabolite could be suppressed by Etomoxir, a carnitine palmitoyl transferase I inhibitor, indicating beta-oxidation of 131I-PHIPA 3-10 in mitochondria. Final evidence for the involvement of mitochondria in the degradation of 131I-PHIPA 3-10 was obtained by density-gradient centrifugation of homogenized rat heart tissue. The position of the labeled free PHIPA 3-10 and free metabolite peaked within the fraction containing mainly mitochondria. CONCLUSION: In spite of its bulky structure, 131I-PHIPA 3-10 is extracted by the myocardium in a manner similar to the extraction of the unmodified fatty acid analog, IPPA. The retention of PHIPA 3-10 in heart muscle results from the presence of the p-phenylene group, which prevents more than one beta-oxidation cycle. Intracellular free PHIPA 3-10 and free PHIPA 1-10 are present in the mitochondria, whereas most of the esterified metabolite was found in the cytosolic lipid pool. Hence, the rapid appearance of PHIPA 1-10 in the lipid pool must be accounted for by mitochondrial leakage or by an unknown in-out transport system.  相似文献   

16.
In Mycobacterium phlei, fatty acid unsaturation increased with decreasing temperature. The 10-hexadecenoic acid content increased as the temperature was reduced from 35 degrees C to 26-20 degrees C. At lower temperatures tuberculostearic acid decreased while oleic and linoleic acids increased, the latter being found in M. phlei for the first time. Concomitantly palmitic acid content decreased, and the 6- and 9-hexadecenoic acids increased slightly on reducing the temperature from 35 to 10 degrees C. Thus, down to 26-20 degrees C palmitic acid was mainly replaced by 10-hexadecenoic acid. From this range down to 10 degrees C, palmitic and tuberculostearic acids were replaced by oleic and linoleic acids. Consequently, fatty acid branching decreased and mean chain length increased, as the temperature was reduced. These observations support the view that regulation of membrane fatty acid composition is part of microbial temperature adaptation, and that the mechanism behind the responses might be more complex than generally believed.  相似文献   

17.
This study reports the effects of a novel polyunsaturated 3-thia fatty acid, methyl 3-thiaoctadeca-6,9,12,15-tetraenoate on serum lipids and key enzymes in hepatic fatty acid metabolism compared to a saturated 3-thia fatty acid, tetradecylthioacetic acid. Palmitic acid treated rats served as controls. Fatty acids were administered by gavage in daily doses of 150 mg/kg body weight for 10 days. The aim of the present study was: (a) To investigate the effect of a polyunsaturated 3-thia fatty acid ester, methyl 3-thiaoctadeca-6,9,12,15-tetraenoate on plasma lipids in normolipidemic rats: (b) to verify whether the lipid-lowering effect could be consistent with enhanced fatty acid oxidation: and (c) to study whether decreased activity of esterifying enzymes and diversion to phospholipid synthesis is a concerted mechanism in limiting the availability of free fatty acid as a substrate for hepatic triglyceride formation. Repeated administration of the polyunsaturated 3-thia fatty acid ester for 10 days resulted in a reduction of plasma triglycerides (40%), cholesterol (33%) and phospholipids (20%) compared to controls. Administration of polyunsaturated and saturated 3-thia fatty acids (daily doses of 150 mg/kg body weight) reduced levels of lipids to a similar extent and followed about the same time-course. Both mitochondrial and peroxisomal fatty acid oxidation increased (1.4-fold- and 4.2-fold, respectively) and significantly increased activities of carnitine palmitoyltransferase (CPT) (1.6-fold), 2,4-dienoyl-CoA reductase (1.2-fold) and fatty acyl-CoA oxidase (3.0-fold) were observed in polyunsaturated 3-thia fatty acid treated animals. This was accompanied by increased CPT-II mRNA (1.7-fold). 2,4-dienoyl-CoA reductase mRNA (2.9-fold) and fatty acyl-CoA oxidase mRNA (1.7-fold). Compared to controls, the hepatic triglyceride biosynthesis was retarded as indicated by a decrease in liver triglyceride content (40%). The activities of glycerophosphate acyltransferase, acyl-CoA: 1,2-diacylglycerol acyltransferase and CTP:phosphocholine cytidylyltransferase were increased. The cholesterol lowering effect was accompanied by a reduction in HMG-CoA reductase activity (80%) and acyl-CoA:cholesterol acyltransferase activity (33%). In hepatocytes treated with methyl 3-thiaoctadeca-6,9,12,15-tetraenoate, fatty acid oxidation was increased 1.8-fold compared to controls. The results suggest that treatment with methyl 3-thiaoctadeca-6,9,12,15-tetraenoate reduces plasma triglycerides by a decrease in the availability of fatty acid substrate for triglyceride biosynthesis via enhanced fatty acid oxidation, most likely attributed to the mitochondrial fatty acid oxidation. It is hypothesized that decreased phosphatidate phosphohydrolase activity may be an additive mechanism which contribute whereby 3-thia fatty acids reduce triglyceride formation in the liver. The cholesterol-lowering effect of the polyunsaturated 3-thia fatty acid ester may be due to changes in cholesterol/cholesterol ester synthesis as 60% of this acid was observed in the hepatic cholesterol ester fraction.  相似文献   

18.
Five (20S)-10-hydroxycamptothecin derivatives carrying the long-chain fatty acid esters were prepared for the development of a new class of prodrug-type agents. In vitro experiments using three kinds of purified carboxylesterase isozymes from the liver microsomes of rat, pig, and human demonstrated that these derivatives were efficiently metabolized by enzymes compared with CPT-11.  相似文献   

19.
Five compounds, which inhibited the amidolytic activity of soluble tissue factor/activated factor VII complex (sTF/VIIa), were isolated from two traditional Chinese medicinal plants commonly used in the treatment of cardiovascular and cerebrovascular diseases. The active compounds were found to be linolenic, linoleic, and oleic acids from roots of Salvia miltiorrhiza; and two anacardic acids, 6-(8'Z-pentadecenyl)- and 6-(10'Z-heptadecenyl)-salicylic acids, from leaves of Ginkgo biloba. The IC50 values were in the range 30-80 micromol/L. Palmitic acid, isolated from roots of Salvia miltiorrhiza, and 2-[(3',7',11',15'-tetramethyl)-2'E,6'E,10'E, 14'E-hexadecatetraenyl]-1,4-hydroquinone, isolated from the marine sponge Adocia viola, did not inhibit sTF/VIIa. Further expansion of the structure-activity relationship to include anacardic acids, 6-(8'Z,11'Z-heptadecadienyl)- and 6-(8'Z, 11'Z, 14'Z-heptadecatrienyl)-salicylic acids from leaves of Anacardium spondias, and other fatty acids demonstrated that at least one cis double bond was essential for inhibitory activity, and that fatty acids containing two or three cis double bonds were optimal. Evidence from preincubation studies implied that these fatty acids may exert their effect by binding to VIIa and consequently preventing binding of sTF to VIIa.  相似文献   

20.
PURPOSE: The effects of embryonic development on lipid composition in the retina were studied in 7, 11, 15, and 18-day-old chick embryos and newly hatched chicks. METHODS: The proportions of phospholipids, free and esterified cholesterol, diacylglycerides, and free fatty acids were determined using the Iatroscan TLC/FID procedure. Gas chromatography and mass spectrometry were used to determine the fatty acid composition. RESULTS: The major phospholipid species were phosphatidylethanolamine, phosphatidylcholine, phosphatidylserine, phosphatidylinositol, lysophosphatidylcholine, and sphingomyelin. Concentrations of the analyzed components have been related to the chronology of concrete stages of retinal development. The fatty acid composition of the total lipids, (n-6):(n-3) and saturated: unsaturated fatty acid ratios, and other parameters are reported. The proportions of total saturated and total monounsaturated fatty acids decreased very little from day 7 to hatching, whereas total polyunsaturated fatty acids nearly doubled over the same period. The increase in C18:2(n-6) from day 11 onwards was not followed by a similar increase in C20:4(n-6), hence the C20:4 to C18:2 ratio decreased with age. CONCLUSIONS: The cholesterol:phospholipid ratio decreased from day 7 to day 15 and increased from day 15 to hatching. High proportions of esterified cholesterol, very probably originating in the retinal pigment epithelium, were also recorded. Total saturated and monounsaturated fatty acids decreased, while polyunsaturated fatty acids increased during the period of initial retinal growth.  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