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.
Chronic dietary administration of 3,3′,4,4′,5,5′-hexabromobiphenyl (HBB), 1 mg/kg diet, caused a decrease in retinol (20-fold) and retinyl esters (23-fold) in the livers of female rats, but resulted in a 6.4-fold increase in retinol and 7.4-fold increase in retinyl esters in the kidneys. Liver acyl-CoA: retinol acyltransferase and retinyl palmitate hydrolase activities were reduced while serum concentration of retinol was unaffected by HBB feeding. Metabolism of a physiological dose of [11-3H]retinyl acetate (10 μg), was examined in rats fed either vitamin A-adequate diet, or marginal amounts of vitamin A, or vitamin A-adequate diet containing HBB. A 13-fold greater amount of the administered vitamin A was found in kidneys of HBB-treated rats. In rats fed adequate or low amounts of vitamin A, kidney radioactivity was primarily in the retinol fraction, while in HBB-fed rats the radioactivity was associated mostly with retinyl esters. Fecal and urinary excretion of radioactivity was greatly increased in HBB-treated rats. Chronic HBB feeding results in a loss of ability of liver to store vitamin A, and severely alters the uptake and metabolism of vitamin A in the kidneys. We conclude that HBB causes major disturbances in the regulation of vitamin A metabolism.  相似文献   

2.
Retinyl esters represent an insoluble storage form of vitamin A and are substrates for the retinoid isomerase (Rpe65) in cells of the retinal pigment epithelium (RPE). The major retinyl-ester synthase in RPE cells is lecithin:retinol acyl-transferase (LRAT). A second palmitoyl coenzyme A-dependent retinyl-ester synthase activity has been observed in RPE homogenates but the protein responsible has not been identified. Here we show that diacylglycerol O-acyltransferase-1 (DGAT1) is expressed in multiple cells of the retina including RPE and Müller glial cells. DGAT1 catalyzes the synthesis of retinyl esters from multiple retinol isomers with similar catalytic efficiencies. Loss of DGAT1 in dgat1 -/- mice has no effect on retinal anatomy or the ultrastructure of photoreceptor outer-segments (OS) and RPE cells. Levels of visual chromophore in dgat1 -/- mice were also normal. However, the normal build-up of all-trans-retinyl esters (all-trans-RE’s) in the RPE during the first hour after a deep photobleach of visual pigments in the retina was not seen in dgat1 -/- mice. Further, total retinyl-ester synthase activity was reduced in both dgat1 -/- retina and RPE.  相似文献   

3.
By use of a new high-resolution high-pressure liquid chromatographic method for the separation of isomeric forms of retinol, retinal, retinyl ester and retinal oxime, various retinoids were analyzed in separated retinal pigment epithelial tissue or neural retinal tissue from fresh bleached bovine eyes after incubation in the dark at either 30 or 4°C for 90 min. 11-cis-Retinoids significantly increased during incubation at 30°C, relative to those at 4°C, in the retinal pigment epithelium, but not in the retina. The major forms of vitamin A in incubated retinal pigment epithelium and neural retina were retinyl esters (70%) and all-trans-retinol (69%), respectively. Thus, in keeping with observations on the isomerization of radioactive retinol in homogenates of eye tissues, the retinal pigment epithelium seems to be the primary site of 11-cis-retinoid formation from endogenous all-trans-retinoids in the bovine eye.  相似文献   

4.
A study was conducted on the incorporation of [11-3H]retinyl acetate into various retinyl esters in liver tissues of rats either vitamin A-sufficient, vitamin A-deficient or vitamin A-deficient and maintained on retinoic acid. Further, the metabolism of [11-3H]retinyl acetate to polar metabolites in liver tissues of these three groups of animals was investigated. Retinol metabolites were analyzed by high-performance liquid chromatography. In vitamin A-sufficient rat liver, the incorporation of radioactivity into retinyl palmitate and stearate was observed at 0.25 h after the injection of the label. The label was further detected in retinyl laurate, myristate, palmitoleate, linoleate, pentadecanoate and heptadecanoate 3 h after the injection. The specific radioactivities (dpm/nmol) of all retinyl esters increased with time. However, the rate of increase in the specific radioactivity of retinyl laurate was found to be significantly higher (66-fold) than that of retinyl palmitate 24 h after the injection of the label. 7 days after the injection of the label, the specific radioactivity between different retinyl esters were found to be similar, indicating that newly dosed labelled vitamin A had now mixed uniformly with the endogenous pool of vitamin A in the liver. The esterification of labelled retinol was not detected in liver tissues of vitamin A-deficient or retinoic acid-supplemented rats at any of the time point studied. Among the polar metabolites analyzed, the formation of [3H]retinoic acid from [3H]retinyl acetate was found only in vitamin A-deficient rat liver 24 h after the injection of the label. A new polar metabolite of retinol (RM) was detected in liver of the three groups of animals. The formation of 3H-labelled metabolite RM from [3H]retinyl acetate was not detected until 7 days after the injection of the label in the vitamin A-sufficient rat liver, suggesting that metabolite RM could be derived from a more stable pool of vitamin A.  相似文献   

5.
Chronic dietary administration of 3,3',4,4',5,5'-hexabromobiphenyl (HBB), 1 mg/kg diet, caused a decrease in retinol (20-fold) and retinyl esters (23-fold) in the livers of female rats, but resulted in a 6.4-fold increase in retinol and 7.4-fold increase in retinyl esters in the kidneys. Liver acyl-CoA:retinol acyltransferase and retinyl palmitate hydrolase activities were reduced while serum concentration of retinol was unaffected by HBB feeding. Metabolism of a physiological dose of [11-3H]retinyl acetate (10 micrograms), was examined in rats fed either vitamin A-adequate diet, or marginal amounts of vitamin A, or vitamin A-adequate diet containing HBB. A 13-fold greater amount of the administered vitamin A was found in kidneys of HBB-treated rats. In rats fed adequate or low amounts of vitamin A, kidney radioactivity was primarily in the retinol fraction, while in HBB-fed rats the radioactivity was associated mostly with retinyl esters. Fecal and urinary excretion of radioactivity was greatly increased in HBB-treated rats. Chronic HBB feeding results in a loss of ability of liver to store vitamin A, and severely alters the uptake and metabolism of vitamin A in the kidneys. We conclude that HBB causes major disturbances in the regulation of vitamin A metabolism.  相似文献   

6.
All-trans retinoyl fluoride was prepared by treating all-trans retinoic acid with diethylaminosulfurtrifluoride. The crystalline product, which was characterized by melting point, infrared, 1H-NMR, 19F-NMR and elementary analysis, showed λmax at 382 nm in hexane (ε = 4.98·104 M?1·cm?1) and at 392 nm in methanol (ε = 4.60·104 M?1·cm?1). Its biological activity in the rat growth assay, relative to all-trans retinyl acetate, was 22% ± 10%. Upon oral administration for 5 days to vitamin A-depleted rats, retinoyl fluoride (1020 μg) was rapidly metabolized to a polar metabolite fraction and, in the intestine, to an unstable retinol-like metabolite, purpotedly 15-fluororetinol. Upon administering intraperitoneally smaller doses (47–94 μg) of [11-3H]retinoyl fluoride, which was synthesized from [11-3H] retinoic acid, radioactive retinoic acid was noted in the liver and plasma but not in the intestine. As expected, a radioactive polar fraction appeared in the intestine and liver, but radioactive retinol, retinyl ester and some common oxidation products were not detected. Of the administered radioactivity, 72% was excreted in the urine, and only 4% was found in the feces over a 7-day period. Hydrolysis of the urine gave a radioactive fraction with a polarity similar to that of retinoic acid. Retinoyl fluoride also reacts readily with glycine to yield N-retinoyl glycine. Thus, the biological activity of retinoyl fluoride can be attributed to the formation of retinoic acid, probably by way of N-retinoyl derivatives. A possible pathway for its metabolism is presented.  相似文献   

7.
The steady-state concentrations of retinol in rat tissues varied as a function of dietary α-tocopherol. The liver, kidney, and intestinal retinol concentrations increased in animals fed an α-tocopherol-deficient diet despite a decrease (liver) or no change (kidney and intestine) in the concentrations of total vitamin A. In contrast, in lung the concentrations of both retinol and total vitamin A decreased. α-Tocopherol inhibited retinyl palmitate hydrolase in vitro in liver, kidney, and intestine; had minimal effect on the testes hydrolase; and stimulated the lung hydrolase. Fifty percent inhibition of the liver hydrolase was provided by an α-tocopherol concentration (100 μm), close to that reported in livers of rats fed a purified diet, constituted with moderately low amounts of α-tocopheryl acetate. Phylloquinone (vitamin K1) inhibited the retinyl palmitate hydrolase in vitro in all tissues tested, and was about fivefold more potent than α-tocopherol. The effects of phylloquinone and α-tocopherol on the liver hydrolase were additive, not synergistic. The antioxidant N,N′-diphenyl-p-phenylenediamine, the most effective synthetic vitamin E substitute known, had little effect on the hydrolase. These data show that α-tocopherol effects vitamin A metabolism in several tissues, and suggest that it may be a physiological effector of tissue retinol homeostasis.  相似文献   

8.
The isomerization of all-trans retinol (vitamin A) to 11-cis retinol in the retinal pigment epithelium (RPE) is a key step in the visual process for the regeneration of the visual pigment chromophore, 11-cis retinal. LRAT and RPE65 are recognized as the minimal isomerase catalytic components. However, regulators of this rate-limiting step are not fully identified and could account for the phenotypic variability associated with inherited retinal degeneration (RD) caused by mutations in the RPE65 gene. To identify new RPE65 partners, we screened a porcine RPE mRNA library using a yeast two-hybrid assay with full-length human RPE65. One identified clone (here named FATP1c), containing the cytosolic C-terminal sequence from the fatty acid transport protein 1 (FATP1 or SLC27A1, solute carrier family 27 member 1), was demonstrated to interact dose-dependently with the native RPE65 and with LRAT. Furthermore, these interacting proteins colocalize in the RPE. Cellular reconstitution of human interacting proteins shows that FATP1 markedly inhibits 11-cis retinol production by acting on the production of all-trans retinyl esters and the isomerase activity of RPE65. The identification of this new visual cycle inhibitory component in RPE may contribute to further understanding of retinal pathogenesis.  相似文献   

9.
10.
Isotope dilution is currently the most accurate technique in humans to determine vitamin A status and bioavailability/bioconversion of provitamin A carotenoids such as β-carotene. However, limits of MS detection, coupled with extensive isolation procedures, have hindered investigations of physiologically-relevant doses of stable isotopes in large intervention trials. Here, a sensitive liquid chromatography-tandem mass spectrometry (LC/MS/MS) analytical method was developed to study the plasma response from coadministered oral doses of 2 mg [13C10]β-carotene and 1 mg [13C10]retinyl acetate in human subjects over a 2 week period. A reverse phase C18 column and binary mobile phase solvent system separated β-carotene, retinol, retinyl acetate, retinyl linoleate, retinyl palmitate/retinyl oleate, and retinyl stearate within a 7 min run time. Selected reaction monitoring of analytes was performed under atmospheric pressure chemical ionization in positive mode at m/z 537→321 and m/z 269→93 for respective [12C]β-carotene and [12C] retinoids; m/z 547→330 and m/z 274→98 for [13C10]β-carotene and [13C5] cleavage products; and m/z 279→100 for metabolites of [13C10]retinyl acetate. A single one-phase solvent extraction, with no saponification or purification steps, left retinyl esters intact for determination of intestinally-derived retinol in chylomicrons versus retinol from the liver bound to retinol binding protein. Coadministration of [13C10]retinyl acetate with [13C10]β-carotene not only acts as a reference dose for inter-individual variations in absorption and chylomicron clearance rates, but also allows for simultaneous determination of an individual''s vitamin A status.  相似文献   

11.
The four most important non-specific carboxylesterases from rat liver were assayed for their ability to hydrolyse retinyl esters. Only the esterases with pI 6.2 and 6.4 (= esterase ES-4) are able to hydrolyse retinyl palmitate. Their specific activities strongly depend on the emulsifier used (maximum rate: 440 nmol of retinol liberated/h per mg of esterase). Beside retinyl palmitate, these esterases cleave palmitoyl-CoA and monoacylglycerols with much higher rates, as well as certain drugs (e.g. aspirin and propanidid). However, no transacylation between palmitoyl-CoA and retinol occurs. Retinyl acetate also is a substrate for the above esterases and for another one with pI 5.6 (= esterase ES-3). Again the emulsifier influences the hydrolysis by these esterases (maximum rates: 475 nmol/h per mg for ES-4 and 200 nmol/h per mg for ES-3). Differential centrifugation of rat liver homogenate reveals that retinyl palmitate hydrolase activity is highly enriched in the plasma membranes, but only moderately so in the endoplasmic reticulum, where the investigated esterases are located. Since the latter activity can be largely inhibited with the selective esterase inhibitor bis-(4-nitrophenyl) phosphate, it is concluded that the esterases with pI 6.2 and 6.4 (ES-4) represent the main retinyl palmitate hydrolase of rat liver endoplasmic reticulum. In view of this cellular localization, the enzyme could possibly be involved in the mobilization of retinol from the vitamin A esters stored in the liver. However, preliminary experiments in vivo have failed to demonstrate such a biological function.  相似文献   

12.
The mechanism responsible for the metabolism of vitamin A during hypervitaminosis is largely unknown. This study investigated hepatic 13C-retinol uptake in hypervitaminotic A rhesus monkeys. We hypothesized that individual retinyl esters would be enriched in 13C after a physiologic dose of 13C2-retinyl acetate, thus suggesting de novo in vivo hepatic retinol esterification. Male rhesus macaques (n = 16; 11.8 ± 2.9 y) each received 3.5 µmol 14, 15-13C2-retinyl acetate. Blood was drawn at baseline and 5 h and 2, 4, 7, 14, 21, and 28 d after administration. Liver biopsies were collected 7 d before and 2 d after dose administration (n = 4) and at 7, 14, and 28 d after dose administration (n = 4 per time point). 13C enrichments of retinol and retinyl esters HPLC-purified from liver samples were measured by using gas chromatography–combustion–isotope ratio mass spectrometry. 13C enrichment of total vitamin A and individual retinyl esters were significantly greater 2 d after dose administration compared with baseline levels. In contrast, the concentration of isolated retinyl esters did not always increase 2 d after treatment. Given that the liver biopsy site differed between monkeys, these data suggest that the accumulation of hepatic retinyl esters is a dynamic process that is better represented by combining analytical techniques. This sensitive methodology can be used to characterize vitamin A trafficking after physiologic doses of 13C-retinol. In this nonhuman primate model of hypervitaminosis A, hepatic retinyl esters continued to accumulate with high liver stores.Abbreviations: GCCIRMS, gas chromatography–combustion–isotope ratio mass spectrometry, IRMS, isotope ratio mass spectrometry, PS/LO, ratio of retinyl palmitate plus stearate to retinyl linoleate plus oleateVitamin A is critical for vision, reproduction, and cellular differentiation.18 All tissue vitamin A originates as dietary vitamin A, which is predominantly available as preformed vitamin A (that is, retinyl esters)3 and the provitamin A carotenoids. Retinyl esters are cleaved to retinol in the intestinal lumen, and unesterified retinol is absorbed into the enterocyte.25 Retinol is esterified within the intestinal mucosa23,31 and then packaged into chylomicrons. These particles are exocytosed and transported into the general circulation,4 where they are degraded to chylomicron remnants and taken up by the liver.5,9 Once in hepatic parenchymal cells, retinyl esters are rapidly hydrolyzed to retinol and, depending on the animals’ vitamin A status, retinol is secreted back into plasma bound to retinol-binding protein or transferred to stellate cells, reesterified, and stored.6 Thus, vitamin A in the liver occurs in 2 forms: as retinol and esterified to various fatty acids. Analysis of hepatic retinyl esters within 30 min of intravenous injection of labeled chylomicron retinyl ester in vitamin-A–sufficient rats recovered 80% to 90% of the dose. During vitamin A sufficiency, a majority of labeled chylomicron retinyl esters are taken up by the liver before subsequent hydrolysis.5Little is known about the storage and metabolism of vitamin A during hypervitaminosis A,32,49 despite the wide use of retinoids pharmaceutically.48 An early study involving hypervitaminotic A rats characterized increased concentrations of retinyl esters as percentages of total vitamin A in the plasma profile during excessive consumption of vitamin A.32 This study further identified that retinyl esters were transported in the serum by means of lipoproteins, thus mediating the vitamin''s nonspecific delivery to body tissues.32 Increased plasma levels of retinyl esters as a percentage of total vitamin A also occurred in human patients10,28,49 and rhesus monkeys41 from the same colony as those used in the current study. Numerous case studies in humans document various symptoms of hypervitaminosis A, including dermatologic, hepatic, and neurologic pathologies.27,30,35,47,48 Hepatic pathologies include abnormal liver function tests consistent with accumulation of lipid-storing droplets. Similar accumulation of lipid droplets has been reported to occur in rhesus monkeys from the same colony as those in the current study.39Previously, liver vitamin A concentrations in captive rhesus monkeys were reported to range from 11.9 ± 5.4 to 18.8 ± 6.4 μmol retinol/g liver,8,33,39 which are several fold higher than the concentrations considered excessive (that is, 0.70 to 1.05 μmol/g liver) and toxic (that is, 1.05 μmol/g liver) in humans.36 Hepatic vitamin A concentrations in 2 wild-caught control rhesus monkeys in a vitamin A deficiency study were 1.07 and 1.08 μmol retinol/g liver.38 Systematic inquiry to uncover the source of the high liver vitamin A concentrations found that the dietary vitamin A intake of captive rhesus monkeys exceeds National Research Council recommendations.40 Consistent with these excessive dietary vitamin A levels, clinical markers of hypervitaminosis A were present.8 To monitor the trafficking of a vitamin A dose during chronic hypervitaminosis A, we treated rhesus monkeys from the same colony as those cited earlier with 13C2-retinyl acetate and collected liver biopsies as part of a study reported elsewhere.8 The main goal of the previous report was to validate a vitamin A assessment technique using the heavy stable isotope of carbon.In the current study, we hypothesized that isotope ratio mass spectrometry (IRMS) could be used to detect accumulation of individual hepatic retinyl esters and provide evidence of de novo in vivo hepatic retinol esterification after treatment of rhesus macaques with a physiologic dose of 13C2-retinyl acetate. The 13C abundance of HPLC-purified hepatic retinyl esters collected at baseline was compared with that of posttreatment samples. Here we show that hepatic retinyl esters continue to accumulate in a nonhuman primate model by using state-of-the-art analytical methods and 13C-labeled retinol as a tracer.  相似文献   

13.
P D Bishop  M D Griswold 《Biochemistry》1987,26(23):7511-7518
When cultured Sertoli cells derived from 20-day-old weanling rats were supplied [3H]retinol bound to serum retinol binding protein-transthyretin complex, [3H]retinol was rapidly incorporated and [3H]retinyl esters were synthesized. Within 28 h after administration, 83% of the labeled retinoids were accounted for as retinyl esters (64% as retinyl palmitate). Sertoli cells derived from vitamin A deficient rats and supplied [3H]retinol in culture under identical conditions likewise incorporated [3H]retinol and synthesized retinyl esters. In contrast to normal Sertoli cells, vitamin A deficient Sertoli cells eventually metabolized virtually all of the cellular [3H]retinol to retinyl esters. The primary metabolic fate of retinol administered to Sertoli cell cultures was the synthesis of retinyl esters under all conditions tested. However, administration of [3H]retinol bound to serum retinol binding protein gave metabolic profiles having a higher proportion of retinyl esters and lower proportions of unresolved polar material than administration of [3H]retinol alone. The kinetics of retinol uptake and intracellular retinyl ester synthesis in cultured Sertoli cells was complex. An initial, rapid phase of [3H]retinol incorporation lasting 30 min was followed by a slower rate of incorporation and a concomitant decrease in the intracellular concentration of [3H]retinol. During the time course the specific activity of [3H]retinyl palmitate eventually exceeded that of intracellular [3H]retinol. These observations suggest that two intracellular pools of retinol may exist in Sertoli cells.  相似文献   

14.
Retinoids (vitamin A) are known to be involved in many key biological functions in mammals, such as embryonic development, reproduction or vision. Besides standard vitamin A forms, freshwater fish tissues contain high levels of didehydroretinoids or vitamin A2 forms. However, the tissue distribution, metabolism and function of both standard and particularly the didehydroretinoids are still poorly known in fish. In this study, we have quantified the levels of retinoids, including retinol, retinaldehyde, retinyl palmitate and their corresponding didehydro forms, as well as the levels of the active polar retinoids all-trans-, 9-cis- and 13-cis-retinoic acid in distinct tissues of juvenile rainbow trout. Our results indicate that the liver is clearly the main retinoid storage tissue in juvenile rainbow trout. Didehydroretinoids were dominant over retinoids in all analyzed tissues with the exception of plasma. Additionally, significant differences among tissues were observed between retinoids and didehydroretinoids, such as differences in the ester profiles and the proportions between free and esterified forms, suggesting that mechanisms that favor the utilization or storage of one of the other groups of compounds might exist in fish. Our data also show the presence of polar retinoids in different tissues of fish at the fmol/g scale. Overall, this study clearly demonstrates the presence of tissue-specific patterns of accumulation of both polar and nonpolar retinoids in fish tissues. The biological relevance of these findings should be the focus of future studies.  相似文献   

15.
Previous studies have shown that rats treated with 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD) show signs of toxicity that are similar to the responses of animals to a vitamin A-deficient diet. These include hypophagia, loss of body weight, loss of hepatic vitamin A, and accumulation of renal retinoids. Male Sprague-Dawley rats treated with 10, 30, or 100 nmol/kg of TCDD accumulated renal vitamin A, with retinyl palmitate concentrations reaching 8 times those of control animals, similar to that of male rats fed a vitamin A-free diet for 26 days. Acyl CoA:retinol acyltransferase (ACARAT) activities in both TCDD-treated rats and rats fed a vitamin A-free diet for 26 days were similarly elevated, and were strongly and positively correlated with the renal retinyl palmitate concentrations. Retinol concentrations in the kidneys of rats treated with TCDD or fed a vitamin A-free diet were only slightly elevated when compared to control rats. We suggest that accumulation of retinyl esters in the kidneys of rats treated with TCDD or fed a vitamin A-free diet occurs as a result of increased rates of retinol esterification.  相似文献   

16.
Hepatic stellate cells (HSC) store vitamin A as retinyl esters and control circulating retinol levels. Upon liver injury, quiescent (q)HSC lose their vitamin A and transdifferentiate to myofibroblasts, e.g. activated (a)HSC, which promote fibrosis by producing excessive extracellular matrix. Adipose triglyceride lipase/patatin-like phospholipase domain-containing protein 2 (ATGL/PNPLA2) and adiponutrin (ADPN/PNPLA3) have so far been shown to mobilize retinol from retinyl esters in HSC. Here, we studied the putative role of hormone-sensitive lipase (HSL/LIPE) in HSC, as it is the major retinyl ester hydrolase (REH) in adipose tissue.Lipe/HSL expression was analyzed in rat liver and primary human and rat qHSC and culture-activated aHSC. Retinyl hydrolysis was analyzed after Isoproterenol-mediated phosphorylation/activation of HSL.Primary human HSC contain 2.5-fold higher LIPE mRNA levels compared to hepatocytes. Healthy rat liver contains significant mRNA and protein levels of HSL/Lipe, which predominates in qHSC and cells of the portal tree. Q-PCR comparison indicates that Lipe mRNA levels in qHSC are dominant over Pnpla2 and Pnpla3. HSL is mostly phosphorylated/activated in qHSC and partly colocalizes with vitamin A-containing lipid droplets. Lipe/HSL and Pnpla3 expression is rapidly lost during HSC culture-activation, while Pnpla2 expression is maintained. HSL super-activation by isoproterenol accelerates loss of lipid droplets and retinyl palmitate from HSC, which coincided with a small, but significant reduction in HSC proliferation and suppression of Collagen1A1 mRNA and protein levels.In conclusion, HSL participates in vitamin A metabolism in qHSC. Equivalent activities of ATGL and ADPN provide the healthy liver with multiple routes to control circulating retinol levels.  相似文献   

17.
The key step in the visual pigment regeneration process is an enzyme-catalyzedtrans tocis retinoid isomerization reaction. This reaction is of substantial general interest, because it requires the input of metabolic energy. The energy is needed because the 11-cis-retinoid reaction products are approximately 4kcal/mol higher in energy than their all-trans congeners. In the retinal pigment epithelium a novel enzymatic system has been discovered which is capable of converting all-trans-retinol into all-trans retinyl esters, by means of a lecithin retinol acyl transferase (LRAT), followed by the direct processing of the ester into 11-cis-retinol. In this process the free energy of hydrolysis of a retinyl ester, estimated to be approximately –5kcal/mol, is coupled to the endothermic (+4kcal/mol) isomerization reaction, resulting in an overall exothermic process. The overall process is analogous to ATP-dependent group transfer reactions, but here the energy is provided by the membrane phospholipids. This process illustrates a new role for membranes: they can serve as an energy source.  相似文献   

18.
1. retro-Retinyl acetate was shown to exert its biological activity by conversion into vitamin A. 2. When administered orally, retro-retinyl acetate was hydrolysed to retro-retinol in the intestine, isomerized to retinol and esterified before being transported to the liver for storage. 3. Administration of the compound at as high a dose as 4·0mg./day for 4 days led to the accumulation of both vitamin A and retro-vitamin A in the liver. The amount of retro-vitamin A in liver gradually decreased until it was almost completely converted into vitamin A in 18 days. 4. Intraperitoneal administration of the compound led to the accumulation of both vitamin A and retro-vitamin A in liver and other tissues. No vitamin A was detected in any tissue of rats receiving retro-retinyl acetate intraperitoneally after enterectomy. 5. The small intestine is the major site of conversion of retro-vitamin A into vitamin A. The conversion could also be demonstrated by everted intestinal sacs. 6. The biological potency of retro-retinyl acetate determined by the rat-growth assay was 20·5% that of all-trans-retinyl acetate, when given orally.  相似文献   

19.
Approximately 80–90% of all retinoids in the body are stored as retinyl esters (REs) in the liver. Adipose tissue also contributes significantly to RE storage. The present studies, employing genetic and nutritional interventions, explored factors that are responsible for regulating RE accumulation in the liver and adipose tissue and how these influence levels of retinoic acid (RA) and RA-responsive gene expression. Our data establish that acyl-CoA:retinol acyltransferase (ARAT) activity is not involved in RE synthesis in the liver, even when mice are nutritionally stressed by feeding a 25-fold excess retinol diet or upon ablation of cellular retinol-binding protein type I (CRBPI), which is proposed to limit retinol availability to ARATs. Unlike the liver, where lecithin:retinol acyltransferase (LRAT) is responsible for all RE synthesis, this is not true for adipose tissue where Lrat-deficient mice display significantly elevated RE concentrations. However, when CrbpI is also absent, RE levels resemble wild-type levels, suggesting a role for CrbpI in RE accumulation in adipose tissue. Although expression of several RA-responsive genes is elevated in Lrat-deficient liver, employing a sensitive liquid chromatography tandem mass spectrometry protocol and contrary to what has been assumed for many years, we did not detect elevated concentrations of all-trans-RA. The elevated RA-responsive gene expression was associated with elevated hepatic triglyceride levels and decreased expression of Pparδ and its downstream Pdk4 target, suggesting a role for RA in these processes in vivo.  相似文献   

20.
Previous studies have shown that rats treated with 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD) show signs of toxicity that are similar to the responses of animals to a vitamin A-deficient diet. These include hypophagia, loss of body weight, loss of hepatic vitamin A, and accumulation of renal retinoids. Male Sprague-Dawley rats treated with 10, 30, or 100 nmol/kg of TCDD accumulated renal vitamin A, with retinyl palmitate concentrations reaching 8 times those of control animals, similar to that of male rats fed a vitamin A-free diet for 26 days. Acyl CoA:retinol acyltransferase (ACARAT) activities in both TCDD-treated rats and rats fed a vitamin A-free diet for 26 days were similarly elevated, and were strongly and positively correlated with the renal retinyl palmitate concentrations. Retinol concentrations in the kidneys of rats treated with TCDD or fed a vitamin A-free diet were only slightly elevated when compared to control rats. We suggest that accumulation of retinyl esters in the kidneys of rats treated with TCDD or fed a vitamin A-free diet occurs as a result of increased rates of retinol esterification.  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