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 20 毫秒
1.
A new method of high-performance liquid chromatography (HPLC) analysis to quantify isomers of retinol, retinal and retinoic acid simultaneously was established. The HPLC system consisted of a silica gel absorption column and a linear gradient with two kinds of solvents containing n-Hexane, 2-propanol, and glacial acetic acid in different ratios. It separated six retinoic acid isomers (13-cis, 9-cis, all-trans, all-trans-4-oxo, 9-cis-4-oxo, 13-cis-4-oxo), three retinal isomers (13-cis-, 9-cis-, and all-trans) and two retinol isomers (13-cis- and all-trans). Human serum samples were subjected to this HPLC analysis and at least, all-trans retinol, 13-cis retinol, and all-trans retinoic acid were detectable. This HPLC system is useful for evaluating retinoic acid formation from retinol via a two-step oxidation pathway. Moreover, it could be applied to monitoring the concentrations of various retinoids, including all-trans retinoic acid in human sera.  相似文献   

2.
Liver alcohol dehydrogenase (E.C.1.1.1.1) is an NAD+/NADH dependent enzyme with a broad substrate specificity being active on an assortment of primary and secondary alcohols. It catalyzes the reversible oxidation of a wide variety of alcohols to the corresponding aldehydes and ketones as well as the oxidation of certain aldehydes to their related carboxylic acids. Although the bioinorganic and bioorganic aspects of the enzymatic mechanism, as well as the structures of various ternary complexes, have been extensively studied, the kinetic significance of certain intermediates has not been fully evaluated. Nevertheless, the availability of computer-assisted programs for kinetic simulation and molecular modeling make it possible to describe the biochemical mechanism more completely. Although the true physiological substrates of this zinc metalloenzyme are unknown, alcohol dehydrogenase effectively catalyzes not only the interconversion of all-trans-retinol and all-trans-retinal but also the oxidation of all-trans-retinal to the corresponding retinoic acid. Retinal and related vitamin A derivatives play fundamental roles in many physiological processes, most notably the vision process. Furthermore, retinoic acid is used in dermatology as well as in the prevention and treatment of different types of cancer. The enzyme-NAD+-retinol complex has an apparent pKa value of 7.2 and loses a proton rapidly. Proton inventory modeling suggests that the transition state for the hydride transfer step has a partial negative charge on the oxygen of retinoxide. Spectral evidence for an intermediate such as E-NAD+-retinoxide was obtained with enzyme that has cobalt(II) substituted for the active site zinc(II). Biophysical considerations of water in these biological processes coupled with the inverse solvent isotope effect lead to the conclusion that the zinc-bound alkoxide makes a strong hydrogen bond with the hydroxyl group of Ser48 and is thus activated for hydride transfer. Moderate pressure accelerates enzyme action indicative of a negative volume of activation. The data with retinol is discussed in terms of enzyme stability, mechanism, adaptation to extreme conditions, as well as water affinities of substrates and inhibitors. Our data concern all-trans, 9-cis, 11-cis, and 13-cis retinols as well as the corresponding retinals. In all cases the enzyme utilizes an approximately ordered mechanism for retinol–retinal interconversion and for retinal–retinoic acid transformation.  相似文献   

3.
Mouse ADH4 (purified, recombinant) has a low catalytic efficiency for ethanol and acetaldehyde, but very high activity with longer chain alcohols and aldehydes, at pH 7.3 and temperature 37 degrees C. The observed turnover numbers and catalytic efficiencies for the oxidation of all-trans-retinol and the reduction of all-trans-retinal and 9-cis-retinal are low relative to other substrates; 9-cis-retinal is more reactive than all-trans-retinal. The reduction of all-trans- or 9-cis-retinals coupled to the oxidation of ethanol by NAD(+) is as efficient as the reduction with NADH. However, the Michaelis constant for ethanol is about 100 mM, which indicates that the activity would be lower at physiologically relevant concentrations of ethanol. Simulations of the oxidation of retinol to retinoic acid with mouse ADH4 and human aldehyde dehydrogenase (ALDH1), using rate constants estimated for all steps in the mechanism, suggest that ethanol (50 mM) would modestly decrease production of retinoic acid. However, if the K(m) for ethanol were smaller, as for human ADH4, the rate of retinol oxidation and formation of retinoic acid would be significantly decreased during metabolism of 50 mM ethanol. These studies begin to describe quantitatively the roles of enzymes involved in the metabolism of alcohols and carbonyl compounds.  相似文献   

4.

Background

Cellular retinol binding-protein I (CRBPI) and cellular retinol binding-protein II (CRBPII) serve as intracellular retinoid chaperones that bind retinol and retinal with high affinity and facilitate substrate delivery to select enzymes that catalyze retinoic acid (RA) and retinyl ester biosynthesis. Recently, 9-cis-RA has been identified in vivo in the pancreas, where it contributes to regulating glucose-stimulated insulin secretion. In vitro, 9-cis-RA activates RXR (retinoid × receptors), which serve as therapeutic targets for treating cancer and metabolic diseases. Binding affinities and structure–function relationships have been well characterized for CRBPI and CRBPII with all-trans-retinoids, but not for 9-cis-retinoids. This study extended current knowledge by establishing binding affinities for CRBPI and CRBPII with 9-cis-retinoids.

Methods

We have determined apparent dissociation constants, K′d, through monitoring binding of 9-cis-retinol, 9-cis-retinal, and 9-cis-RA with CRBPI and CRBPII by fluorescence spectroscopy, and analyzing the data with non-linear regression. We compared these data to the data we obtained for all-trans- and 13-cis-retinoids under identical conditions.

Results

CRBPI and CRBPII, respectively, bind 9-cis-retinol (K′d, 11 nM and 68 nM) and 9-cis-retinal (K′d, 8 nM and 5 nM) with high affinity. No significant 9-cis-RA binding was observed with CRBPI or CRBPII.

Conclusions

CRBPI and CRBPII bind 9-cis-retinol and 9-cis-retinal with high affinities, albeit with affinities somewhat lower than for all-trans-retinol and all-trans-retinal.

General significance

These data provide further insight into structure–binding relationships of cellular retinol binding-proteins and are consistent with a model of 9-cis-RA biosynthesis that involves chaperoned delivery of 9-cis-retinoids to enzymes that recognize retinoid binding-proteins.  相似文献   

5.
The role of 9-cis-β-carotene (9-cis-β-C) as a potential precursor of 9-cis-retinoic acid (9-cis-RA) has been examined in human intestinal microcosa in vitro. By using HPLC, uv spectra, and chemical derivatization analysis, both 9-cis-RA and all-trans-retinoic acid (all-trans-RA) have been identified in the postnuclear fraction of human intestinal microcosa after incubation with 9-cis-β-C at 37°C. The biosynthesis of both 9-cis-RA and all-trans-RA from 9-cis-β-C was linear with increasing concentrations of 9-cis-β-C (2-30 μM) and was linear with respect to tissue protein concentration up to 0.75 mg/ml. Retinoic acid was not detected when a boiled incubation mixture was incubated in the presence of 9-cis-β-C. The rate of synthesis of 9-cis- and all-trans-RA from 4 μM 9-cis-β-C were 16 ± 1 and 18 ± 2 pmol/hr/mg of protein, respectively. However, when 2 μM all-trans-β-C was added to the 4 μM 9-cis-β-C, the rate of all-trans-RA synthesis was increased to 38 ± 6 pmol/hr/mg of protein, whereas the rate of 9-cis-RA synthesis remained the same. These results suggest that 9-cis-RA is produced directly from 9-cis-β-C. Furthermore, incubations of either 0.1 μM 9-cis- or all-trans-retinal under the same incubation conditions showed that 9-cis-RA could also arise through oxidative conversion of 9-cis-retinal. Although only 9-cis-RA was detected when 9-cis-RA was used as the substrate, the isomerization of the all-trans-RA to 9-cis-RA cannot be ruled out, since both all-trans-RA and trace amounts of 9-cis-RA were detected when all-trans-retinal was incubated as the substrate. These data indicate that 9-cis-β-C can be a source of 9-cis-RA in the human. This conversion may have a significance in the anticarcinogenic action of β-C.  相似文献   

6.
Our previous study suggested the presence of a novel cone-specific redox reaction that generates 11-cis-retinal from 11-cis-retinol in the carp retina. This reaction is unique in that 1) both 11-cis-retinol and all-trans-retinal were required to produce 11-cis-retinal; 2) together with 11-cis-retinal, all-trans-retinol was produced at a 1:1 ratio; and 3) the addition of enzyme cofactors such as NADP(H) was not necessary. This reaction is probably part of the reactions in a cone-specific retinoid cycle required for cone visual pigment regeneration with the use of 11-cis-retinol supplied from Müller cells. In this study, using purified carp cone membrane preparations, we first confirmed that the reaction is a redox-coupling reaction between retinals and retinols. We further examined the substrate specificity, reaction mechanism, and subcellular localization of this reaction. Oxidation was specific for 11-cis-retinol and 9-cis-retinol. In contrast, reduction showed low specificity: many aldehydes, including all-trans-, 9-cis-, 11-cis-, and 13-cis-retinals and even benzaldehyde, supported the reaction. On the basis of kinetic studies of this reaction (aldehyde-alcohol redox-coupling reaction), we found that formation of a ternary complex of a retinol, an aldehyde, and a postulated enzyme seemed to be necessary, which suggested the presence of both the retinol- and aldehyde-binding sites in this enzyme. A subcellular fractionation study showed that the activity is present almost exclusively in the cone inner segment. These results suggest the presence of an effective production mechanism of 11-cis-retinal in the cone inner segment to regenerate visual pigment.  相似文献   

7.
Bacteriorhodopsin monomer dispersed in a solution of the detergent L-1690 could maintain the specific interaction between retinal and protein in the pH range 9.0-0.0 at 25°C. λmax of the absorbance spectrum was 550 nm at pH 9.0, 556 nm at pH 5.5, 609 nm at pH 2.1 and 570 nm at pH 0.0. Increasing the NaCl concentration in the solution promoted formation of the 609 nm product at pH 5.0-3.0 and also its transition to the 570 nm product at pH 2.5-1.0. Retinal isomer analysis gave a ratio of 13-cis- to all-trans-retinal of 53 : 47 at pH 5.5. When the pH of the solution was reduced, the relative content of all-trans-retinal increased and the ratio of 13-cis- to all-trans-retinal was 14 : 86 at pH 0.0. Illumination of the solution at pH 7.2 yielded a product containing 9-cis-retinal or 9-cis, 13-cis-retinal, which may be due to a reaction other than the photoreaction cycle.  相似文献   

8.
All-trans-retinoic acid (all-trans-RA) and 13-cis-retinoic acid (13-cis-RA), due to their effects on cell differentiation, proliferation and angiogenesis, improved treatment results in some malignancies. Pharmacokinetic studies of all-trans-RA and 13-cis-RA along with monitoring of retinoic acid metabolites may help to optimize retinoic acid therapy and to develop new effective strategies for the use of retinoic acids in cancer treatment. Therefore, we developed a HPLC method for the simultaneous determination in human plasma of the physiologically important retinoic acid isomers, all-trans-, 13-cis- and 9-cis-retinoic acid, their 4-oxo metabolites, 13-cis-4-oxoretinoic acid (13-cis-4-oxo-RA) and all-trans-4-oxoretinoic acid (all-trans-4-oxo-RA), and vitamin A (all-trans-retinol). Analysis performed on a silica gel column with UV detection at 350 nm using a binary multistep gradient composed on n-hexane, 2-propanolol and glacial acetic acid. For liquid-liquid extraction a mixture of n-hexane, dichloromethane and 2-propanolol was used. The limits of detection were 0.5 ng/ml for retinoic acids and 10 ng/ml for all-trans-retinol. The method showed good reproducibility for all components (within-day C.V.: 3.02–11.70%; day-to-day C.V.: 0.01–11.34%. Furthermore, 9-cis-4-oxoretinoic acid (9-cis-4-oxo-RA) is separated from all-trans-4-oxo-RA and 13-cis-4-oxo-RA. In case of clinical use of 9-cis-retinoic acid (9-cis-RA) the pharmacokinetics and metabolism of this retinoic acid isomer can also be examined.  相似文献   

9.
The G protein coupled receptor rhodopsin contains a pocket within its seven-transmembrane helix (TM) structure, which bears the inactivating 11-cis-retinal bound by a protonated Schiff-base to Lys296 in TM7. Light-induced 11-cis-/all-trans-isomerization leads to the Schiff-base deprotonated active Meta II intermediate. With Meta II decay, the Schiff-base bond is hydrolyzed, all-trans-retinal is released from the pocket, and the apoprotein opsin reloaded with new 11-cis-retinal. The crystal structure of opsin in its active Ops* conformation provides the basis for computational modeling of retinal release and uptake. The ligand-free 7TM bundle of opsin opens into the hydrophobic membrane layer through openings A (between TM1 and 7), and B (between TM5 and 6), respectively. Using skeleton search and molecular docking, we find a continuous channel through the protein that connects these two openings and comprises in its central part the retinal binding pocket. The channel traverses the receptor over a distance of ca. 70 Å and is between 11.6 and 3.2 Å wide. Both openings are lined with aromatic residues, while the central part is highly polar. Four constrictions within the channel are so narrow that they must stretch to allow passage of the retinal β-ionone-ring. Constrictions are at openings A and B, respectively, and at Trp265 and Lys296 within the retinal pocket. The lysine enforces a 90° elbow-like kink in the channel which limits retinal passage. With a favorable Lys side chain conformation, 11-cis-retinal can take the turn, whereas passage of the all-trans isomer would require more global conformational changes. We discuss possible scenarios for the uptake of 11-cis- and release of all-trans-retinal. If the uptake gate of 11-cis-retinal is assigned to opening B, all-trans is likely to leave through the same gate. The unidirectional passage proposed previously requires uptake of 11-cis-retinal through A and release of photolyzed all-trans-retinal through B.  相似文献   

10.
The retina-specific ATP binding cassette transporter, ABCA4 protein, is associated with a broad range of inherited macular degenerations, including Stargardt disease, autosomal recessive cone rod dystrophy, and fundus flavimaculatus. In order to understand its role in retinal transport in rod out segment discs, we have investigated the interactions of the soluble domains of ABCA4 with both 11-cis- and all-trans-retinal. Using fluorescence anisotropy-based binding analysis and recombinant polypeptides derived from the amino acid sequences of the four soluble domains of ABCA4, we demonstrated that the nucleotide binding domain 1 (NBD1) specifically bound 11-cis-retinal. Its affinity for all-trans-retinal was markedly reduced. Stargardt disease-associated mutations in this domain resulted in attenuation of 11-cis-retinal binding. Significant differences in 11-cis-retinal binding affinities were observed between NBD1 and other cytoplasmic and lumenal domains of ABCA4. The results suggest a possible role of ABCA4 and, in particular, the NBD1 domain in 11-cis-retinal binding. These results also correlate well with a recent report on the in vivo role of ABCA4 in 11-cis-retinal transport.  相似文献   

11.
Opn3/TMT opsins belong to one of the opsin groups with vertebrate visual and non-visual opsins, and are widely distributed in eyes, brains and other internal organs in various vertebrates and invertebrates. Vertebrate Opn3/TMT opsins are further classified into four groups on the basis of their amino acid identities. However, there is limited information about molecular properties of these groups, due to the difficulty in preparing the recombinant proteins. Here, we successfully expressed recombinant proteins of TMT1 and TMT2 opsins of medaka fish (Oryzias latipes) in cultured cells and characterized their molecular properties. Spectroscopic and biochemical studies demonstrated that TMT1 and TMT2 opsins functioned as blue light-sensitive Gi/Go-coupled receptors, but exhibited spectral properties and photo-convertibility of the active state different from each other. TMT1 opsin forms a visible light-absorbing active state containing all-trans-retinal, which can be photo-converted to 7-cis- and 9-cis-retinal states in addition to the original 11-cis-retinal state. In contrast, the active state of TMT2 opsin is a UV light-absorbing state having all-trans-retinal and does not photo-convert to any other state, including the original 11-cis-retinal state. Thus, TMT opsins are diversified so as to form a different type of active state, which may be responsible for their different functions.  相似文献   

12.
A major goal in vision research over the past few decades has been to understand the molecular details of retinoid processing within the retinoid (visual) cycle. This includes the consequences of side reactions that result from delayed all-trans-retinal clearance and condensation with phospholipids that characterize a variety of serious retinal diseases. Knowledge of the basic retinoid biochemistry involved in these diseases is essential for development of effective therapeutics. Photoisomerization of the 11-cis-retinal chromophore of rhodopsin triggers a complex set of metabolic transformations collectively termed phototransduction that ultimately lead to light perception. Continuity of vision depends on continuous conversion of all-trans-retinal back to the 11-cis-retinal isomer. This process takes place in a series of reactions known as the retinoid cycle, which occur in photoreceptor and RPE cells. All-trans-retinal, the initial substrate of this cycle, is a chemically reactive aldehyde that can form toxic conjugates with proteins and lipids. Therefore, much experimental effort has been devoted to elucidate molecular mechanisms of the retinoid cycle and all-trans-retinal-mediated retinal degeneration, resulting in delineation of many key steps involved in regenerating 11-cis-retinal. Three particularly important reactions are catalyzed by enzymes broadly classified as acyltransferases, short-chain dehydrogenases/reductases and carotenoid/retinoid isomerases/oxygenases. This article is part of a Special Issue entitled: Retinoid and Lipid Metabolism.  相似文献   

13.
The vertebrate visual photoreceptor rhodopsin (Rho) is a unique G protein-coupled receptor as it utilizes a covalently tethered inverse agonist (11-cis-retinal) as the native ligand. Previously, electrophysiological studies showed that ligand binding of 11-cis-retinal in dark-adapted Rho was essentially irreversible with a half-life estimated to be 420 years, until after thermal isomerization to all-trans-retinal, which then slowly dissociates. This long lifetime of 11-cis-retinal binding was considered to be physiologically important for minimizing background signal (dark noise) of the visual system. However, in vitro biochemical studies on the thermal stability of Rho showed that Rho decays with a half-life on the order of days. In this study, we resolve the discrepancy by measuring the chromophore exchange rate of the bound 11-cis-retinal chromophore with free 9-cis-retinal from Rho in an in vitro phospholipid/detergent bicelle system. We conclude that the thermal decay of Rho primarily proceeds through spontaneous breaking of the covalent linkage between opsin and 11-cis-retinal, which was overlooked in the electrophysiological recording. We estimate that this slow spontaneous release of 11-cis-retinal from Rho should result in 104 to 105 free opsin molecules in a dark-adapted rod cell—a number that is three orders of magnitude higher than previously expected. We also discuss the physiological implications of these findings on the basal activity of opsins and the associated dark noise in the visual system.  相似文献   

14.
Exposure to bright light can cause visual dysfunction and retinal photoreceptor damage in humans and experimental animals, but the mechanism(s) remain unclear. We investigated whether the retinoid cycle (i.e. the series of biochemical reactions required for vision through continuous generation of 11-cis-retinal and clearance of all-trans-retinal, respectively) might be involved. Previously, we reported that mice lacking two enzymes responsible for clearing all-trans-retinal, namely photoreceptor-specific ABCA4 (ATP-binding cassette transporter 4) and RDH8 (retinol dehydrogenase 8), manifested retinal abnormalities exacerbated by light and associated with accumulation of diretinoid-pyridinium-ethanolamine (A2E), a condensation product of all-trans-retinal and a surrogate marker for toxic retinoids. Now we show that these mice develop an acute, light-induced retinopathy. However, cross-breeding these animals with lecithin:retinol acyltransferase knock-out mice lacking retinoids within the eye produced progeny that did not exhibit such light-induced retinopathy until gavaged with the artificial chromophore, 9-cis-retinal. No significant ocular accumulation of A2E occurred under these conditions. These results indicate that this acute light-induced retinopathy requires the presence of free all-trans-retinal and not, as generally believed, A2E or other retinoid condensation products. Evidence is presented that the mechanism of toxicity may include plasma membrane permeability and mitochondrial poisoning that lead to caspase activation and mitochondria-associated cell death. These findings further understanding of the mechanisms involved in light-induced retinal degeneration.The retinoid cycle is a fundamental metabolic process in the vertebrate retina responsible for continuous generation of 11-cis-retinal from its all-trans-isomer (1-3). Because 11-cis-retinal is the chromophore of rhodopsin and cone visual pigments (4), disabling mutations in genes encoding proteins of the retinoid cycle can cause a spectrum of retinal diseases affecting sight (3). Moreover, the efficiency of the mammalian visual system and health of photoreceptors and retinal pigment epithelium (RPE)2 decrease significantly with age. Even in the presence of a functional retinoid cycle, A2E, retinal dimer (RALdi), and other toxic all-trans-retinal condensation products (5-7) can accumulate as a consequence of aging (8). Under experimental conditions, these compounds can produce toxic effects on RPE cells (9-11). Patients affected by age-related macular degeneration, Stargardt disease, or other retinal diseases associated with accumulation of surrogate markers, such as A2E, all develop retinal degeneration (12). Thus, elucidating the fundamental causes of these age-dependent changes is of increasing importance. Encouragingly, our understanding of both retinoid metabolism outside the eye and production of 11-cis-retinal unique to the eye has accelerated recently (Scheme 1) (1-3), and genetic mouse models are readily available to study these processes and their potential aberrations in vivo (13). Thus, a central question can be addressed, namely what initiates the death of photoreceptor cells and the underlining RPE?Open in a separate windowSCHEME 1.Retinoid flow and all-trans-retinal clearance in the visual cycle. After diffusion from the RPE, the visual chromophore, 11-cis-retinal, combines with rhodopsin and then is photoisomerized to all-trans-retinal. Most of the all-trans-retinal dissociates from opsin into the cytoplasm, where it is reduced to all-trans-retinol by RDHs, including RDH8. The fraction of all-trans-retinal that dissociates into the disc lumen is transported by ABCA4 into the cytoplasm (23) before it is reduced. All-trans-retinol then is translocated to the RPE, esterified by LRAT, and recycled back to 11-cis-retinal. Mutations of ABCA4 are associated with human macular degeneration, Stargardt disease, and age-related macular degeneration (55, 56).Several mechanisms associated with retinoid metabolism may contribute to different retinopathies (1). For example, lack of retinoids in LRAT (lecithin:retinol acyltransferase) or chromophore in retinoid isomerase knock-out (Rpe65-/-) mice leads to rapid degeneration of cone photoreceptors and slowly progressive death of rods (14). Such mice do not produce toxic condensation products from all-trans-retinal. Instead, their retinopathies have been attributed to continuous activation of visual phototransduction (15) due to either the basal activity of opsin (16-18) or disordered vectorial transport of cone visual pigments without bound chromophore (19). Paradoxically, an abnormally high flux of retinoids through the retinoid cycle can also lead to retinopathy in other mouse models (20, 21). Animal models featuring anomalies in the retinoid cycle illustrate the importance of chromophore regeneration and provide an approach to elucidating mechanisms involved in human retinal dysfunction and disease.Recently, we showed that mice carrying a double knock-out of Rdh8 (retinol dehydrogenase 8), one of the main enzymes that reduces all-trans-retinal in rod and cone outer segments (22), and Abca4 (ATP-binding cassette transporter 4), which transports all-trans-retinal from the inside to the outside of disc membranes (23), rapidly accumulate all-trans-retinal condensation products and exhibit accentuated RPE/photoreceptor dystrophy at an early age (24). Although these studies suggest retinoid toxicity, it is still unclear if the elevated levels of retinal and/or its condensation products, such as A2E, are the cause of this retinopathy or merely a nonspecific reflection of impaired retinoid metabolism. Here, we report that spent chromophore, all-trans-retinal, is most likely responsible for photoreceptor degeneration in Rdh8-/-Abca4-/- mice. Toxic effects of all-trans-retinal include caspase activation and mitochondria-associated cell death.  相似文献   

15.
The metabolism of 13-cis-[11-3H]retinoic acid has been examined in vitamin A-normal rats. Within 24 h after intravenous administration of the parent retinoid (15 μg/kg) to animals with biliary fistulas, 69 ± 9% of the dose was detected in the bile with 9 ± 6% being found in the urine. Analysis of the bile by reverse-phase high-pressure liquid chromatography demonstrated that the retinoic acid was being metabolized to several more polar compounds. A number of these compounds were sensitive to incubation with β-glucuronidase as evidenced by a change in their chromatographic behavior after treatment with the enzyme. Two of the metabolites have been identified as 13-cis-4-oxoretinoyl-β-glucuronide (8.1 ± 1.0% of the dose during the first 4 h after administration of the parent compound) and 13-cis-retinoyl-β-glucuronide (7.0 ± 4.4% of the dose). A comparison of the chromatographic profiles of bile from 13-cis- versus all-trans-retinoic acid-treated rats indicated a difference in their metabolism, with a greater proportion of the all-trans-retinoic acid being converted to compounds that eluted in the more polar regions of the column effluent.  相似文献   

16.
The biological activity of various natural retinoids and the time "window" when vitamin A activity is required for normal cardiovascular development were examined in vitamin A-deprived Japanese quail embryos. The administration of 1 μg of retinol at the beginning of incubation resuited in normal cardiovascular development in 97% of embryos; retinoic acid was toxic at this dose level. Treatment of embryos with 0.1 μg of all-trans-retinol or 13-cis-retinoic acid at the beginning of incubation resulted in normal cardiovascular development in 47 and 12% of embryos, respectively; administration of these retinoids at other time points attenuated the percentage of embryos with normal cardiovascular development. Didehydroretinol, 0.1 μg, and 9-cis-retinoic acid, 0.1 μg, were inactive at all time points examined; 9-cis-retinoic acid did not enhance the biological activity of all-trans-retinoic acid. All-trans-retinoic acid, 0.1 μg, administered during 22-28 hr of incubation induced normal cardiovascular development in 20-34% of embryos; biological activity was optimal when it was administered at 24 hr. All retinoids tested were inactive in establishing normal cardiovascular development when administered at 36 hr of incubation or later. The studies suggest that all-trans-retinoic acid is the biologically active form of vitamin A required for normal cardiovascular development in the avian embryo. There is a critical time point within the first 22-28 hr of quail embryogenesis when all-trans-retinoic acid initiates events that lead to normal cardiovascular development.  相似文献   

17.
A recombinant alcohol dehydrogenase (ADH) from Kangiella koreensis was purified as a 40 kDa dimer with a specific activity of 21.3 nmol min?1 mg?1, a K m of 1.8 μM, and a k cat of 1.7 min?1 for all-trans-retinal using NADH as cofactor. The enzyme showed activity for all-trans-retinol using NAD + as a cofactor. The reaction conditions for all-trans-retinol production were optimal at pH 6.5 and 60 °C, 2 g enzyme l?1, and 2,200 mg all-trans-retinal l?1 in the presence of 5 % (v/v) methanol, 1 % (w/v) hydroquinone, and 10 mM NADH. Under optimized conditions, the ADH produced 600 mg all-trans-retinol l?1 after 3 h, with a conversion yield of 27.3 % (w/w) and a productivity of 200 mg l?1 h?1. This is the first report of the characterization of a bacterial ADH for all-trans-retinal and the biotechnological production of all-trans-retinol using ADH.  相似文献   

18.
An isocratic reversed-phase high-performance liquid chromatographic procedure for the determination of all-trans-retinoic acid (all-trans-RA) and its metabolites, all-trans-4-oxo-RA, 5,6-epoxy-RA, 9-cis-RA and13-cis-RA, in mouse plasma and embryo and in new in vitro potential test systems for development toxicology has been developed. These compounds, their biological precursor retinol (vitamin A) and the internal standard were resolved on a Spherisorb ODS-2 (5 μm) column (250×4.6 mm I.D.) with acetonitrile-water-methanol-n-butyl alcohol (56:37:4:3, v/v) containing 100 mM ammonium acetate and 70 mM acetic acid as the elution system with a total run time of 23 min. The assay was linear over a wide range, with a lower limit of quantitation of 50 ng/ml or 10 ng/ml of protein for all-trans-RA, 13-cis-RA and 9-cis-RA and of 25 ng/ml or 5 ng/ml protein for the 4-oxo- and 5,6-epoxy-metabolites. At these concentrations, intra-assay coefficients of variation (C.V.) of the retinoids were 3–9%. Mean intra-assay C.V. averaged 5–7% in the tissues studied. Its use is discussed for RA measurements in some of the new test systems — Drosophila melanogaster, sea urchin embryos and cultured human keratinocytes — that have to be evaluated in toxicological testing, supplementary to standard assays in mammals.  相似文献   

19.
Studies on a Missing Reaction in the Visual Cycle   总被引:1,自引:0,他引:1  
DEVELOPMENT in the biochemistry of vision during the past twenty-five years can be summarized by equations (1) and (2) in Fig. 1, which envisage1 that 11-cis-retinal combines with the visual protein opsin in a dark reaction to form the photolabile complex rhodopsin, λmax 497 nm. When rhodopsin absorbs light it stimulates, through a process whose mechanism is not understood, the transmission of impulses, which are responsible for the visual sensation, although much is known about the biochemical changes accompanying the absorption of light by rhodopsin. These changes culminate in the formation of all-trans-retinal (λmax 385 nm) and opsin (equation (2), Fig. 1), through a number of intermediates2 and for the completion of the cycle one needs a molecular process which may regenerate 11-cis-retinal from all-trans-retinal (equation (3), Fig. 1).  相似文献   

20.
A reversed-phase high-performance liquid chromatographic method for the simultaneous analysis of retinol, all-trans-retinoic acid, 13-cis-retinoic acid and 13-cis-4-oxoretinoic acid in human plasma and cell culture medium is described. Sample preparation involves precipitation of proteins and extraction of retinoids with 60% acetonitrile. After centrifugation, the acetonitrile content of the supernatant is reduced to 45%, allowing on-column concentration of analytes. Injection volumes up to 2.0 ml (equivalent to 0.525 ml of sample) can be used without compromising chromatographic resolution of all-trans-retinoic acid and 13-cis-retinoic acid. Retinoids were stable in this extract and showed no isomerization when stored in the dark in a cooled autosampler, allowing automated analysis of large series of samples. Recoveries from spiked plasma samples were between 95 and 103%. Although no internal standard was used, the inter-assay precision for all retinoids was better than 6% and 4% at concentrations of 30 nM and 100 nM, respectively. The method is a valuable tool for the study of cellular metabolism of all-trans-retinoic acid, as polar metabolites of this compound can be detected with high sensitivity in cell culture media.  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