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 187 毫秒
1.
The chromophore in halorhodopsin (HR) which acts as a light-driven chloride pump in halobacteria shares many properties with its counterpart in bacteriorhodopsin (BR): (i) a similar retinal protein interaction, (ii) trans to cis isomerization and (iii) similar intermediates of its photocycle. One major difference between the two chromoproteins is that the HR chromophore does not become deprotonated during its photocycle. A mechanism for the photocycle of HR is presented, which, in close analogy to an earlier proposed mechanism for BR, involves the sequence of all-trans 13-cis, 14s-cis 13-cis all-trans isomerizations of the chromophore, a Schiff base of retinal. In contrast to the situation in BR the 13-cis, 14s-cis13-cis isomerization is induced not by deprotonation of the retinal Schiff base chromophore but rather by the movement of an anion (Cl-) towards the protonated nitrogen of the Schiff's base. The suggested mechanism involves the Schiff base directly in the chloride translocation in halorhodopsin.  相似文献   

2.
THE initial reaction following absorption of light in the retina is the isomerization of the 11-cis retinal chromophore of the visual pigment1. Isolated 11cis retinal will undergo the same isomerization to the all-trans form when excited by light of wavelength shorter than about 450 nm and this reaction can be sensitized to light of longer wavelengths by the addition of trace amounts of iodine to the solution2.  相似文献   

3.
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   

4.
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   

5.
We discuss to what extent the vibrational spectra of bacteriorhodopsin that have been observed and assigned by Smith et al. (1, 2) by means of resonance Raman and by Gerwert and Siebert (EMBO (Eur. Mol. Biol. Organ.) J. In press) by means of infrared absorption experiments are in agreement with a photo-cycle of bacteriorhodopsin that involves the sequence BR, IO(all-trans) → K(13,14-cis) → L(13,14-cis) → M(13-cis) → N(13-cis) → O(all-trans). Our discussion is based on a quantumchemical modified neglect of diatomic overlap [MNDO] calculation of the vibrational spectra of the relevant isomers of the protonated retinal Schiff base. In particular, we investigated in these calculations the effects of different charge environments on the frequencies of the relevant C-C single bond stretching vibrations of these isomers.  相似文献   

6.
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   

7.
Halorhodopsin from Natronomonas pharaonis (NpHR) is a member of the retinal protein group and serves as a light-driven chloride pump in which chloride ions are transported through the membrane following light absorption by the retinal chromophore. In this study, we examined two main issues: (1) factors controlling the binding of the retinal chromophore to the NpHR opsin and (2) the ability of the NpHR opsin to catalyze the thermal isomerization of retinal isomers. We have revealed that the reconstitution process of pharaonis HR (NpHR) pigment from its apoprotein and all-trans retinal depends on the pH, and the process has a pKa of 5.8 ± 0.1. It was proposed that this pKa is associated with the pKa of the lysine residue that binds the retinal chromophore (Lys256). The pigment formation is regulated by the concentration of sodium chloride, and the maximum yield was observed at 3.7 M NaCl. The low yield of pigment in a lower concentration of NaCl (< 3 M) may be due to an altered conformation adopted by the apomembrane, which is not capable of forming the pigment. Unexpectedly and unlike the apomembrane of bacteriorhodopsin, NpHR opsin produces pigments with 11-cis retinal and 9-cis retinal owing to the thermal isomerization of these retinal isomers to all-trans retinal. The isomerization rate depends on the pH, and it is faster at a higher pH. The pKa value of the isomerization process is similar to the pKa of the binding process of these retinals, which suggests that Lys256 is also involved in the isomerization process. The isomerization is independent of the sodium chloride concentration. However, in the absence of sodium chloride, the apoprotein adopts such a conformation, which does not prevent the isomerization of retinal, but it prevents a covalent bond formation with the lysine residue. The rate and the thermodynamic parameter analysis of the retinal isomerization by NpHR apoprotein led to the conclusion that the apomembrane catalyzes the isomerization via a triplet mechanism.  相似文献   

8.
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   

9.
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   

10.
11.
Photoisomerization of the chromophore of squid rhodopsin is dependent upon the irradiation temperature. Above 0°C, only 11-cis ? all-trans reaction proceeds and the all-trans → 9-cis reaction is limited to extremely low frequency. At liquid nitrogen temperature, 11-cis ? all-trans ? 9-cis reaction takes place. At intermediary low temperatures (?80°C to ?15°C) another isomer of retinal may be produced by the irradiation, which forms a pigment having an absorbance maximum at 465 nm (P-465). The formation of P-465 decreases remarkably in the narrow temperature range from ?30°C to 0°C where mesorhodopsin converts to metarhodopsin. Mesorhodopsin is quite different from metharhodopsin in the photoisomerization of the chromophore because P-465 is produced from the former but not from the latter. No P-465 is produced both at liquid nitrogen temperature and above 0°C. P-465 is more labile than any of the other photoproducts so far known, that is isorhodopsin, alkaline and acid metarhodopsins. P-465 is converted to metarhodopsin by irradiation.  相似文献   

12.
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   

13.
Opsins are light-sensitive pigments in the vertebrate retina, comprising a G protein-coupled receptor and an 11-cis-retinaldehyde chromophore. Absorption of a photon by an opsin pigment induces isomerization of its chromophore to all-trans-retinaldehyde. After a brief period of activation, opsin releases all-trans-retinaldehyde and becomes insensitive to light. Restoration of light sensitivity to the apo-opsin involves the conversion of all-trans-retinaldehyde back to 11-cis-retinaldehyde via an enzyme pathway called the visual cycle. The critical isomerization step in this pathway is catalyzed by Rpe65. Rpe65 is strongly associated with membranes but contains no membrane-spanning segments. It was previously suggested that the affinity of Rpe65 for membranes is due to palmitoylation of one or more Cys residues. In this study, we re-examined this hypothesis. By two independent strategies involving mass spectrometry, we show that Rpe65 is not palmitoylated nor does it appear to undergo other post-translational modifications at significant stoichiometry. Instead, we show that Rpe65 binds the acidic phospholipids, phosphatidylserine, phosphatidylglycerol, and cardiolipin, but not phosphatidic acid. No binding of Rpe65 to basic phospholipids or neutral lipids was observed. The affinity of Rpe65 to acidic phospholipids was strongly pH-dependent, suggesting an electrostatic interaction of basic residues in Rpe65 with negatively charged phospholipid headgroups. Binding of Rpe65 to liposomes containing phosphatidylserine or phosphatidylglycerol, but not the basic or neutral phospholipids, allowed the enzyme to extract its insoluble substrate, all-trans-retinyl palmitate, from the lipid bilayer for synthesis of 11-cis-retinol. The interaction of Rpe65 with acidic phospholipids is therefore biologically relevant.  相似文献   

14.
细菌视紫红质的质子传输机理   总被引:2,自引:0,他引:2  
细菌视紫红质(bR)是嗜盐菌紫膜中的唯一蛋白质成分, 具有质子泵、电荷分离和光致变色功能. bR分子中的发色团视黄醛通过质子化席夫碱以共价键与Lys216相连. bR分子受可见光照射后, 视黄醛发生从全-反到13-顺式构型的异构化, 导致席夫碱的去质子化,继之以可极化基团位置的改变. 力场的变化引起包括蛋白质三级结构在内的诸多变化, 这些变化促进并保证了质子从细胞质侧向细胞外侧的定向传输.  相似文献   

15.
P. Hegemann  W. Grtner    R. Uhl 《Biophysical journal》1991,60(6):1477-1489
Orientation of the green alga Chlamydomonas in light (phototaxis and stop responses) is controlled by a visual system with a rhodopsin as the functional photoreceptor. Here, we present evidence that in Chlamydomonas wild-type cells all-trans retinal is the predominant isomer and that it is present in amounts similar to that of the rhodopsin itself.

The ability of different retinal isomers and analog compounds to restore photosensitivity in blind Chlamydomonas cells (strain CC2359) was tested by means of flash-induced light scattering transients or by measuring phototaxis in a taxigraph. All-trans retinal reconstitutes behavioral light responses within one minute, whereas cis-isomers require at least 50 × longer incubation times, suggesting that the retinal binding site is specific for all-trans retinal. Experiments with 13-demethyl(dm)-retinal and short-chained analogs reveal that only chromophores with a β-methyl group and at least three double bonds in conjugation with the aldehyde mediate function. Because neither 13-dm-retinal, nor 9,12-phenylretinal restores a functional rhodopsin, a trans/13-cis isomerisation seems to take place in the course of the activation mechanism. We conclude that with respect to its chromophore, Chlamydomonas rhodopsin bears a closer resemblence to bacterial rhodopsins than to visual rhodopsins of higher animals.

  相似文献   

16.
Studies have shown that trans-cis isomerization of retinal is the primary photoreaction in the photocycle of the light-driven proton pump bacteriorhodopsin (BR) from Halobacterium salinarum, as well as in the photocycle of the chloride pump halorhodopsin (HR). The transmembrane proteins HR and BR show extensive structural similarities, but differ in the electrostatic surroundings of the retinal chromophore near the protonated Schiff base. Point mutation of BR of the negatively charged aspartate D85 to a threonine T (D85T) in combination with variation of the pH value and anion concentration is used to study the ultrafast photoisomerization of BR and HR for well-defined electrostatic surroundings of the retinal chromophore. Variations of the pH value and salt concentration allow a switch in the isomerization dynamics of the BR mutant D85T between BR-like and HR-like behaviors. At low salt concentrations or a high pH value (pH 8), the mutant D85T shows a biexponential initial reaction similar to that of HR. The combination of high salt concentration and a low pH value (pH 6) leads to a subpopulation of 25% of the mutant D85T whose stationary and dynamic absorption properties are similar to those of native BR. In this sample, the combination of low pH and high salt concentration reestablishes the electrostatic surroundings originally present in native BR, but only a minor fraction of the D85T molecules have the charge located exactly at the position required for the BR-like fast isomerization reaction. The results suggest that the electrostatics in the native BR protein is optimized by evolution. The accurate location of the fixed charge at the aspartate D85 near the Schiff base in BR is essential for the high efficiency of the primary reaction.  相似文献   

17.
G-protein-coupled receptors transmit stimuli (light, taste, hormone, neurotransmitter, etc.) to the intracellular signaling systems, and rhodopsin (Rh) is the most-studied G-protein-coupled receptor. Rh possesses an 11-cis retinal as the chromophore, and 11-cis to all-trans photoisomerization leads to the protein structural changes in the cytoplasmic loops to activate G-protein. Microbial rhodopsins are similar heptahelical membrane proteins that function as bacterial sensors, light-driven ion-pumps, or light-gated channels. Microbial rhodopsins possess an all-trans retinal, and all-trans to 13-cis photoisomerization triggers protein structural changes for each function. Despite these similarities, there is no sequence homology between visual and microbial rhodopsins, and microbial rhodopsins do not activate G-proteins. However, it was reported that bacteriorhodopsin (BR) chimeras containing the third cytoplasmic loop of bovine Rh are able to activate G-protein, suggesting a common mechanism of protein structural changes. Here we design chimeric proteins for Natronomonas pharaonis sensory rhodopsin II (SRII, also called pharaonis phoborhodopsin), which has a two-orders-of-magnitude slower photocycle than BR. Light-dependent transducin activation was observed for most of the nine SRII chimeras containing the third cytoplasmic loop of bovine Rh (from Y223, G224, Q225 to T251, R252, and M253), but the activation level was 30,000–140,000 times lower than that of bovine Rh. The BR chimera, BR/Rh223-253, activates a G-protein transducin, whereas the activation level was 37,000 times lower than that of bovine Rh. We interpret the low activation by the chimeric proteins as reasonable, because bovine Rh must have been optimized for activating a G-protein transducin during its evolution. On the other hand, similar activation level of the SRII and BR chimeras suggests that the lifetime of the M intermediates is not the simple determinant of activation, because SRII chimeras have two-orders-of-magnitude's slower photocycle than the BR chimera. Activation mechanism of visual and microbial rhodopsins is discussed on the basis of these results.  相似文献   

18.
The redox potentials of the oriented films of the wild-type, the E194Q-, E204Q- and D96N-mutated bacteriorhodopsins (bR), prepared by adsorbing purple membrane (PM) sheets or its mutant on a Pt electrode, have been examined. The redox potentials (V) of the wild-type bR were −470 mV for the 13-cis configuration of the retinal Shiff base in bR and −757 mV for the all-trans configuration in H2O, and −433 mV for the 13-cis configuration and −742 mV for the all-trans configuration in D2O. The solvent isotope effect (ΔV=V(D2O)−V(H2O)), which shifts the redox potential to a higher value, originates from the cooperative rearrangements of the extensively hydrogen-bonded water molecules around the protonated CN part in the retinal Schiff base. The redox potential of bR was much higher for the 13-cis configuration than that for the all-trans configuration. The redox potentials for the E194Q mutant in the extracellular region were −507 mV for the 13-cis configuration and −788 mV for the all-trans configuration; and for the E204Q mutant they were −491 mV for the 13-cis configuration and −769 mV for the all-trans configuration. Replacement of the Glu194 or Glu204 residues by Gln weakened the electron withdrawing interaction to the protonated CN bond in the retinal Schiff base. The E204 residue is less linked with the hydrogen-bonded network of the proton release pathway compared with E194. The redox potentials of the D96N mutant in the cytoplasmic region were −471 mV for the 13-cis configuration and −760 mV for the all-trans configuration which were virtually the same as those of the wild-type bR, indicating that the D to N point mutation of the 96 residue had no influence on the interaction between the D96 residue and the CN part in the Schiff base under the light-adapted condition. The results suggest that the redox potential of bR is closely correlated to the hydrogen-bonded network spanning from the retinal Schiff base to the extracellular surface of bR in the proton transfer pathway.  相似文献   

19.
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   

20.
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   

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