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 9 毫秒
1.
Cellular retinal-binding protein from bovine retina purifies with bound 11-cis-retinal and 11-cis-retinol as endogenous ligands. Inasmuch as these retinoids are interconvertible by a dehydrogenase reaction the accessibility of the aldehyde function of bound 11-cis-retinal to chemical and enzymatic reducing agents was determined. An 11-cis-retinol dehydrogenase from retinal pigment epithelial microsomes, first described by Lion, F., Rotmans, J.P., Daemen, F.J.M. and Bonting, S.L. (Biochim. Biophys. Acta 384, 283–292, 1975) was found to reduce complexed 11-cis-retinal at pH 5.5 and 37°C rapidaly and nearly quantitatively. The product of the reduction, 11-cis-retinol, remained complexed with the binding protein following the reaction. Reduction proceeded 3-times more rapidly with NADH than with NADPH. No change in geometrical isomeric configuration occurred during the reaction. The dehydrogenase from retinal pigment epithelium oxidized 11-cis-retinol complexed with cellular retinal-binding protein at pH 8.5 in the presence of NAD. In spite of the ready enzymatic reduction of 11-cis-retinal complexed with cellular retinal-binding protein, the aldehyde function was inaccessible to several chemical reducing agents. Incubation of the complex with NaBH4 at pH 7.5 and NaCNBH3 or borane dimethylamine at pH 5.5 did not result in reduction of 11-cis-retinal unless the complex had been exposed to white light, a treatment known to produce all-rans-retinal which has little affinity for the binding protein. Liver alcohol dehydrogenase produced only 10% reduction of 11-cis-retinal complexed with cellular retinal-binding protein in 15 min at 37°C when added in amounts which produced about 60% reduction of the uncomplexed retinoid. The results suggest that the interaction between the 11-cis-retinol dehydrogenase and the 11-cis-retinal complexed to cellular retinal-binding protein is a specific one of that the binding protein may function as a substrate carrier for a dehydrogenase.  相似文献   

2.
11-cis-Retinol has previously been shown in physiological experiments to promote dark adaptation and recovery of photoresponsiveness of bleached salamander red cones but not of bleached salamander red rods. The purpose of this study was to evaluate the direct interaction of 11-cis-retinol with expressed human and salamander cone opsins, and to determine by microspectrophotometry pigment formation in isolated salamander photoreceptors. We show here in a cell-free system using incorporation of radioactive guanosine 5′-3-O-(thio)triphosphate into transducin as an index of activity, that 11-cis-retinol inactivates expressed salamander cone opsins, acting an inverse agonist. Similar results were obtained with expressed human red and green opsins. 11-cis-Retinol had no significant effect on the activity of human blue cone opsin. In contrast, 11-cis-retinol activates the expressed salamander and human red rod opsins, acting as an agonist. Using microspectrophotometry of salamander cone photoreceptors before and after bleaching and following subsequent treatment with 11-cis-retinol, we show that 11-cis-retinol promotes pigment formation. Pigment was not formed in salamander red rods or green rods (containing the same opsin as blue cones) treated under the same conditions. These results demonstrate that 11-cis-retinol is not a useful substrate for rod photoreceptors although it is for cone photoreceptors. These data support the premise that rods and cones have mechanisms for handling retinoids and regenerating visual pigment that are specific to photoreceptor type. These mechanisms are critical to providing regenerated pigments in a time scale required for the function of these two types of photoreceptors.11-cis-Retinol is the precursor to 11-cis-retinal, the 11-cis-aldehyde form of vitamin A and the chromophore that combines covalently with rod and cone opsin proteins to form visual pigments. 11-cis-Retinal is consumed during visual signaling, and its continual synthesis is required. Photon absorption by the visual pigments causes the isomerization of its chromophore to the all-trans configuration. This initiates two processes critical for vision: activation of the photoreceptor cell and the eventual recovery of the original photosensitivity of the cells, requiring regeneration of the visual pigments. As cones are used for bright light vision, these two processes must work more rapidly in cones than in rods and thus cones have a higher requirement of 11-cis-retinoids as suggested by Rushton (1, 2).Photoreceptor activation begins with photoisomerization of the chromophore within the visual pigment. This results in a subsequent conformational change of the protein part of the visual pigment that is able to activate its G protein transducin, which in turn activates a PDE that lowers the concentration of cGMP and closes cGMP-gated ion channels. These steps comprise the visual signal transduction cascade (see Ref. 3 for review).The visual cycle involves regeneration of the visual pigment, which ultimately deactivates the protein and accomplishes the recovery of the photosensitivity of the photoreceptor cell. Classically, this process involves both the photoreceptor cell and the retinal pigment epithelium (RPE).4 After photoisomerization of the chromophore and formation of the active visual pigment, all-trans-retinal is released from the opsin and reduced to all-trans-retinol, which is then transported to the RPE where it is isomerized to 11-cis-retinol through a number of steps. In the RPE, 11-cis-retinol is oxidized to the aldehyde form, which is transported back to the photoreceptor cell and can be directly used by all of the opsins to regenerate an inactive pigment ready for photoactivation. The details of this model have been extensively reviewed (4, 5). Alternatively, recent work suggests that cones have an additional source of 11-cis-retinoids from Müller cells (68). Like the RPE cells, Müller cells have been shown to be able to convert all-trans-retinol to 11-cis-retinol (6). Unlike in the RPE cells, 11-cis-retinol is not oxidized to 11-cis-retinal in Müller cells.Jones et al. (9) demonstrated that administration of 11-cis-retinol to bleached salamander red cones could restore photosensitivity. A logical conclusion was that red cones were able to oxidize 11-cis-retinol to the aldehyde and regenerate visual pigments although noncovalent binding of 11-cis-retinol to red cone opsins generating a light-sensitive complex could not be excluded. On the other hand, 11-cis-retinol does not restore photosensitivity to bleached salamander rod cells but appears to directly activate the cells (9, 10). The data suggested that the rods were not able to oxidize 11-cis-retinol, but that the retinol itself could activate the signal transduction cascade, and indeed we recently demonstrated that 11-cis-retinol acts as an agonist to expressed bovine rod opsin (11). Our aim here was to study the action of 11-cis-retinol on cone opsins and cone photoreceptor cells to determine the efficacy of an alternate visual cycle for cones.The photoreceptor cells used in this study are from tiger salamander, and the expressed opsins used for biochemical experiments are those from salamander and human. Photoreceptor cells are generally identified by cell morphology and the type of opsin it contains that can be further complicated by the findings that some cone cells have multiple opsins (12, 13). Recently genetic analysis has determined that opsins fall into five classes (reviewed in Refs. 14 and 15). We have studied opsins falling into four of these classes and use common color-derived names for the opsins and photoreceptor cells. The classic rod cells used for scotopic vision contain rhodopsin, the visual pigment for the rod opsin (RH1 opsin) and appeared red and thus have been designated as red rods. Some species such as salamanders have an additional rod cell whose photosensitivity is blue-shifted from that of the red rod and thus designated as green rods. In the tiger salamander, the green rods contain the identical opsin (SWS2 opsin) found in blue cones (16). The human blue cones contain an opsin from a different class (SWS1 opsin), which is homologous to the salamander UV cone opsin. The human red and green and salamander red cone opsins all belong to the same class of opsins (M/LWS opsins). Absorption properties of visual pigments are further modulated in some animals including the tiger salamander by use of 11-cis-retinal with an additional double bond (3,4-dehydro or A2 11-cis-retinal) resulting in red-shifted absorbance from pigments containing 11-cis-retinal (A1 11-cis-retinal).We show here that 11-cis-retinol is not an agonist to cone opsins and does not itself generate a light-sensitive opsin. We further show using microspectrophotometry that both red and blue salamander cone cells regenerate visual pigments from 11-cis-retinol, whereas pigments could not be regenerated with 11-cis-retinol in bleached salamander red and green rods even though the latter contains the same opsin as the salamander blue cone. Thus, rods and cones have mechanisms for handling retinoids and regenerating visual pigment that are specific to photoreceptor type, and these mechanisms are critical to providing regenerated pigments in a time scale required for the function of these two types of photoreceptors.  相似文献   

3.

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   

4.
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   

5.
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   

6.
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°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 Km 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   

7.
Cone photoreceptors require effective pigment regeneration mechanisms to maintain their sensitivity in the light. Our previous studies in carp cones suggested the presence of an unconventional and very effective mechanism to produce 11-cis retinal, the necessary component in pigment regeneration. In this reaction (aldehyde-alcohol redox coupling reaction, AL-OL coupling reaction), formation of 11-cis retinal, i.e. oxidation of 11-cis retinol is coupled to reduction of an aldehyde at a 1:1 molar ratio without exogenous NADP(H) which is usually required in this kind of reaction. Here, we identified carp retinol dehydrogenase 13-like (RDH13L) as an enzyme catalyzing the AL-OL coupling reaction. RDH13L was partially purified from purified carp cones, identified as a candidate protein, and its AL-OL coupling activity was confirmed using recombinant RDH13L. We further examined the substrate specificity, subcellular localization, and expression level of RDH13L. Based on these results, we concluded that RDH13L contributes to a significant part, but not all, of the AL-OL coupling activity in carp cones. RDH13L contained tightly bound NADP+ which presumably functions as a cofactor in the reaction. Mouse RDH14, a mouse homolog of carp RDH13L, also showed the AL-OL coupling activity. Interestingly, although carp cone membranes, carp RDH13L and mouse RDH14 all showed the coupling activity at 15–37 °C, they also showed a conventional NADP+-dependent 11-cis retinol oxidation activity above 25 °C without addition of aldehydes. This dual mechanism of 11-cis retinal synthesis attained by carp RDH13L and mouse RDH14 probably contribute to effective pigment regeneration in cones that function in the light.  相似文献   

8.
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   

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.
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   

11.
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   

12.
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   

13.
The mechanism of retinol isomerization in the vertebrate retina visual cycle remains controversial. Does the isomerase enzyme RPE65 operate via nucleophilic addition at C11 of the all-trans substrate, or via a carbocation mechanism? To determine this, we modeled the RPE65 substrate cleft to identify residues interacting with substrate and/or intermediate. We find that wild-type RPE65 in vitro produces 13-cis and 11-cis isomers equally robustly. All Tyr-239 mutations abolish activity. Trp-331 mutations reduce activity (W331Y to ∼75% of wild type, W331F to ∼50%, and W331L and W331Q to 0%) establishing a requirement for aromaticity, consistent with cation-π carbocation stabilization. Two cleft residues modulate isomerization specificity: Thr-147 is important, because replacement by Ser increases 11-cis relative to 13-cis by 40% compared with wild type. Phe-103 mutations are opposite in action: F103L and F103I dramatically reduce 11-cis synthesis relative to 13-cis synthesis compared with wild type. Thr-147 and Phe-103 thus may be pivotal in controlling RPE65 specificity. Also, mutations affecting RPE65 activity coordinately depress 11-cis and 13-cis isomer production but diverge as 11-cis decreases to zero, whereas 13-cis reaches a plateau consistent with thermal isomerization. Lastly, experiments using labeled retinol showed exchange at 13-cis-retinol C15 oxygen, thus confirming enzymatic isomerization for both isomers. Thus, RPE65 is not inherently 11-cis-specific and can produce both 11- and 13-cis isomers, supporting a carbocation (or radical cation) mechanism for isomerization. Specific visual cycle selectivity for 11-cis isomers instead resides downstream, attributable to mass action by CRALBP, retinol dehydrogenase 5, and high affinity of opsin apoproteins for 11-cis-retinal.  相似文献   

14.
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   

15.
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   

16.
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   

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.
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   

19.
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   

20.
In the vertebrate retina, phototransduction, the conversion of light to an electrical signal, is carried out by the rod and cone photoreceptor cells1-4. Rod photoreceptors are responsible for vision in dim light, cones in bright light. Phototransduction takes place in the outer segment of the photoreceptor cell, a specialized compartment that contains a high concentration of visual pigment, the primary light detector. The visual pigment is composed of a chromophore, 11-cis retinal, attached to a protein, opsin. A photon absorbed by the visual pigment isomerizes the chromophore from 11-cis to all-trans. This photoisomerization brings about a conformational change in the visual pigment that initiates a cascade of reactions culminating in a change in membrane potential, and bringing about the transduction of the light stimulus to an electrical signal. The recovery of the cell from light stimulation involves the deactivation of the intermediates activated by light, and the reestablishment of the membrane potential. Ca2+ modulates the activity of several of the enzymes involved in phototransduction, and its concentration is reduced upon light stimulation. In this way, Ca2+ plays an important role in the recovery of the cell from light stimulation and its adaptation to background light.Another essential part of the recovery process is the regeneration of the visual pigment that has been destroyed during light-detection by the photoisomerization of its 11-cis chromophore to all-trans5-7. This regeneration begins with the release of all-trans retinal by the photoactivated pigment, leaving behind the apo-protein opsin. The released all-trans retinal is rapidly reduced in a reaction utilizing NADPH to all- trans retinol, and opsin combines with fresh 11-cis retinal brought into the outer segment to reform the visual pigment. All-trans retinol is then transferred out of the outer segment and into neighboring cells by the specialized carrier Interphotoreceptor Retinoid Binding Protein (IRBP).Fluorescence imaging of single photoreceptor cells can be used to study their physiology and cell biology. Ca2+-sensitive fluorescent dyes can be used to examine in detail the interplay between outer segment Ca2+ changes and response to light8-12 as well as the role of inner segment Ca2+ stores in Ca2+ homeostasis13,14. Fluorescent dyes can also be used for measuring Mg2+ concentration15, pH, and as tracers of aqueous and membrane compartments16. Finally, the intrinsic fluorescence of all-trans retinol (vitamin A) can be used to monitor the kinetics of its formation and removal in single photoreceptor cells17-19.Download video file.(70M, mov)  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