首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
The mammalian visual dim‐light photoreceptor rhodopsin is considered a prototype G protein‐coupled receptor. Here, we characterize the kinetics of its light‐activation process. Milligram quantities of α,ε‐15N‐labeled tryptophan rhodopsin were produced in stably transfected HEK293 cells. Assignment of the chemical shifts of the indole signals was achieved by generating the single‐point‐tryptophan to phenylalanine mutants, and the kinetics of each of the five tryptophan residues were recorded. We find kinetic partitioning in rhodopsin decay, including three half‐lives, that reveal two parallel processes subsequent to rhodopsin activation that are related to the photocycle. The meta II and meta III states emerge in parallel with a relative ratio of about 3:1. Transient formation of the meta III state was confirmed by flash photolysis experiments. From analysis of the site‐resolved kinetic data we propose the involvement of the E2‐loop in the formation of the meta III state.  相似文献   

2.
The RHO gene encodes the G‐protein‐coupled receptor (GPCR) rhodopsin. Numerous mutations associated with impaired visual cycle have been reported; the G90D mutation leads to a constitutively active mutant form of rhodopsin that causes CSNB disease. We report on the structural investigation of the retinal configuration and conformation in the binding pocket in the dark and light‐activated state by solution and MAS‐NMR spectroscopy. We found two long‐lived dark states for the G90D mutant with the 11‐cis retinal bound as Schiff base in both populations. The second minor population in the dark state is attributed to a slight shift in conformation of the covalently bound 11‐cis retinal caused by the mutation‐induced distortion on the salt bridge formation in the binding pocket. Time‐resolved UV/Vis spectroscopy was used to monitor the functional dynamics of the G90D mutant rhodopsin for all relevant time scales of the photocycle. The G90D mutant retains its conformational heterogeneity during the photocycle.  相似文献   

3.
Solid‐state NMR spectroscopy gives a powerful avenue for investigating G protein‐coupled receptors and other integral membrane proteins in a native‐like environment. This article reviews the use of solid‐state 2H NMR to study the retinal cofactor of rhodopsin in the dark state as well as the meta I and meta II photointermediates. Site‐specific 2H NMR labels have been introduced into three regions (methyl groups) of retinal that are crucially important for the photochemical function of rhodopsin. Despite its phenomenal stability 2H NMR spectroscopy indicates retinal undergoes rapid fluctuations within the protein binding cavity. The spectral lineshapes reveal the methyl groups spin rapidly about their three‐fold (C3) axes with an order parameter for the off‐axial motion of For the dark state, the 2H NMR structure of 11‐cis‐retinal manifests torsional twisting of both the polyene chain and the β‐ionone ring due to steric interactions of the ligand and the protein. Retinal is accommodated within the rhodopsin binding pocket with a negative pretwist about the C11=C12 double bond. Conformational distortion explains its rapid photochemistry and reveals the trajectory of the 11‐cis to trans isomerization. In addition, 2H NMR has been applied to study the retinylidene dynamics in the dark and light‐activated states. Upon isomerization there are drastic changes in the mobility of all three methyl groups. The relaxation data support an activation mechanism whereby the β‐ionone ring of retinal stays in nearly the same environment, without a large displacement of the ligand. Interactions of the β‐ionone ring and the retinylidene Schiff base with the protein transmit the force of the retinal isomerization. Solid‐state 2H NMR thus provides information about the flow of energy that triggers changes in hydrogen‐bonding networks and helix movements in the activation mechanism of the photoreceptor.  相似文献   

4.
Accumulation of all‐trans‐retinal (all‐trans‐RAL), reactive vitamin A aldehyde, is one of the key factors in initiating retinal photodamage. This photodamage is characterized by progressive retinal cell death evoked by light exposure in both an acute and chronic fashion. Photoactivated rhodopsin releases all‐trans‐RAL, which is subsequently transported by ATP‐binding cassette transporter 4 and reduced to all‐trans‐retinol by all‐trans‐retinol dehydrogenases located in photoreceptor cells. Any interruptions in the clearing of all‐trans‐RAL in the photoreceptors can cause an accumulation of this reactive aldehyde and its toxic condensation products. This accumulation may result in the manifestation of retinal dystrophy including human retinal degenerative diseases such as Stargardt’s disease and age‐related macular degeneration. Herein, we discuss the mechanisms of all‐trans‐RAL clearance in photoreceptor cells by sequential enzymatic reactions, the visual (retinoid) cycle, and potential molecular pathways of retinal photodamage. We also review recent imaging technologies to monitor retinal health status as well as novel therapeutic strategies preventing all‐trans‐RAL‐associated retinal photodamage.  相似文献   

5.
All‐trans retinal is a potent photosensitizer that is released in photoreceptor outer segments by the photoactivated visual pigment following the detection of light. Photoreceptor outer segments also contain high concentrations of polyunsaturated fatty acids, and are thus particularly susceptible to oxidative damage such as that initiated by light via a photosensitizer. Upon its release, all‐trans retinal is reduced within the outer segment to all‐trans retinol, through a reaction requiring metabolic input in the form of NADPH. The phototoxic potential of physiologically generated all‐trans retinal was examined in single living rod photoreceptors obtained from frog (Rana pipiens) retinas. Light‐induced oxidation was measured with fluorescence imaging using an oxidation‐sensitive indicator dye from the shift in fluorescence between the intact and oxidized forms. Light‐induced oxidation was highest in metabolically compromised rod outer segments following photoactivation of the visual pigment rhodopsin, and after a time interval, sufficiently long to ensure the release of all‐trans retinal. Furthermore, light‐induced oxidation increased with the concentration of exogenously added all‐trans retinal. The results show that the all‐trans retinal generated during the detection of light can mediate light‐induced oxidation. Its removal through reduction to all‐trans retinol protects photoreceptor outer segments against light‐induced oxidative damage.  相似文献   

6.
The RHO gene encodes the G-protein-coupled receptor (GPCR) rhodopsin. Numerous mutations associated with impaired visual cycle have been reported; the G90D mutation leads to a constitutively active mutant form of rhodopsin that causes CSNB disease. We report on the structural investigation of the retinal configuration and conformation in the binding pocket in the dark and light-activated state by solution and MAS-NMR spectroscopy. We found two long-lived dark states for the G90D mutant with the 11-cis retinal bound as Schiff base in both populations. The second minor population in the dark state is attributed to a slight shift in conformation of the covalently bound 11-cis retinal caused by the mutation-induced distortion on the salt bridge formation in the binding pocket. Time-resolved UV/Vis spectroscopy was used to monitor the functional dynamics of the G90D mutant rhodopsin for all relevant time scales of the photocycle. The G90D mutant retains its conformational heterogeneity during the photocycle.  相似文献   

7.
The photocycle of channelrhodopsin‐2 is investigated in a comprehensive study by ultrafast absorption and fluorescence spectroscopy as well as flash photolysis in the visible spectral range. The ultrafast techniques reveal an excited‐state decay mechanism analogous to that of the archaeal bacteriorhodopsin and sensory rhodopsin II from Natronomonas pharaonis. After a fast vibrational relaxation of the excited‐state population with 150 fs its decay with mainly 400 fs is observed. Hereby, both the initial all‐trans retinal ground state and the 13‐cis‐retinal K photoproduct are populated. The reaction proceeds with a 2.7 ps component assigned to cooling processes. Small spectral shifts are observed on a 200 ps timescale. They are attributed to conformational rearrangements in the retinal binding pocket. The subsequent dynamics progresses with the formation of an M‐like intermediate (7 and 120 μs), which decays into red‐shifted states within 3 ms. Ground‐state recovery including channel closing and reisomerization of the retinal chromophore occurs in a triexponential manner (6 ms, 33 ms, 3.4 s). To learn more about the energy barriers between the different photocycle intermediates, temperature‐dependent flash photolysis measurements are performed between 10 and 30 °C. The first five time constants decrease with increasing temperature. The calculated thermodynamic parameters indicate that the closing mechanism is controlled by large negative entropy changes. The last time constant is temperature independent, which demonstrates that the photocycle is most likely completed by a series of individual steps recovering the initial structure.  相似文献   

8.
The phototransduction cascade is perhaps the best understood model system for G protein‐coupled receptor (GPCR) signaling. Phototransduction links the absorption of a single photon of light to a decrease in cytosolic cGMP. Depletion of the cGMP pool induces closure of cGMP‐gated cation channels resulting in the hyperpolarization of photoreceptor cells and consequently a neuronal response. Many biochemical and both low‐ and high‐resolution structural approaches have been utilized to increase our understanding of rhodopsin, the key molecule of this signaling cascade. Rhodopsin, a member of the GPCR or seven‐transmembrane spanning receptor superfamily, is composed of a chromophore, 11‐cis‐retinal that is covalently bound by a protonated Schiff base linkage to the apo‐protein opsin at Lys296 (in bovine opsin). Upon absorption of a photon, isomerization of the chromophore to an all‐trans‐retinylidene conformation induces changes in the rhodopsin structure, ultimately converting it from an inactive to an activated state. This state allows it to activate the heterotrimeric G protein, transducin, by triggering nucleotide exchange. To fully understand the structural and functional aspects of rhodopsin it is necessary to critically examine crystal structures of its different photointermediates. In this review we summarize recent progress on the structure and activation of rhodopsin in the context of other GPCR structures.  相似文献   

9.
The reaction of the di‐lithiated oxamide of 1 with two equivalents of SnCl2 provided the tin trans‐oxamide 3 . In solution, spectroscopic analysis suggests exclusively the formation of a trans‐oxamide (trans‐ 3 ). However, the solid state shows an atypical cis‐oxamide (cis‐ 3 ), where the oxamide fragment acts as an anti‐Janus head ligand. An 119Sn‐NMR variable temperature experiment ([D8]THF) of the trans‐oxamide (trans‐ 3 ) was performed however, at lower temperature no additional signal was observed, which confirmed the absence of a dynamic equilibrium. Dispersion‐corrected density functional calculations revealed that the cis conformation of this tin(II) oxamide complex is more stable than the trans isomer by 1.4 kcal · mol–1.  相似文献   

10.
The cytoplasmic surface of G protein‐coupled receptors plays a central role for activation and deactivation of the receptor. To understand the molecular mechanisms which underlie these processes, we determined the surface charge density and its changes upon activation directly at the cytoplasmic surface of bovine rhodopsin and correlated these changes with key events in receptor activation. The surface charge density was calculated from the ionic strength dependence of the apparent pKa of the surface‐bound pH‐indicator dye fluorescein according to the Gouy‐Chapman theory. The surface charge density at pH 6.5 changes by 0.8 ± 0.2 elementary charge/1000 Å2 in rod outer segment disk membranes and by 0.4 ± 0.2 elementary charge/1000 Å2 in rhodopsin/dodecylmaltoside micelles upon formation of the active metarhodopsin‐II state. By comparison of these surface charge density values determined with and without the native lipid environment, we calculated the charge change to about 1 elementary charge/cytoplasmic rhodopsin surface. The more positive surface charge density in metarhodopsin‐II decreases back to the dark state level of σ = ?2.0 ± 0.2 elementary charges/1000 Å2 in the opsin state, providing further evidence that the cytoplasmic surface properties after metarhodopsin‐II decay resemble almost those of the dark state.  相似文献   

11.
The physico‐chemical properties as well as the conformation of the cytoplasmic surface of the 7‐helix retinal proteins bacteriorhodopsin (bR) and visual rhodopsin change upon light activation. A recent study found evidence for a transient softening of bR in its key intermediate M [Pieper et al. (2008) Phys. Rev. Lett. 100 , 228103] as a direct proof for the functional significance of protein flexibility. In this report we compare environmental and flexibility changes at the cytoplasmic surface of light‐activated bR and rhodopsin detected by time‐resolved fluorescence spectroscopy. The changes in fluorescence of covalently bound fluorescent probes and protein real‐time dynamics were investigated. We found that in fluorescently labeled bR and rhodopsin the intensity of fluorescein and Atto647 increased upon formation of the key intermediates M and metarhodopsin‐II, respectively, suggesting different surface properties compared to the dark state. Furthermore, time‐resolved fluorescence anisotropy experiments reveal an increase in steric restriction of loop flexibility because of changes in the surrounding protein environment in both the M‐intermediate as well as the active metarhodopsin‐II state. The kinetics of the fluorescence changes at the rhodopsin surface uncover multiple transitions, suggesting metarhodopsin‐II substates with different surface properties. Proton uptake from the aqueous bulk phase correlates with the first transition, while late proton release seems to parallel the second transition. The last transition between states of different surface properties correlates with metarhodopsin‐II decay.  相似文献   

12.
The trans‐bis(trimethylsilyl)chalcogenolate palladium complexes, trans‐[Pd(ESiMe3)2(PnBu3)2] [E = S ( 1 ) and Se ( 2 )] were synthesized in good yields and high purity by reacting trans‐[PdCl2(PBu3)2] with LiESiMe3 (E = S, Se), respectively. These complexes were characterized by 1H, 13C{1H}, 31P{1H} (and 77Se{1H}) NMR spectroscopy and single‐crystal X‐ray analysis. The reaction of 2 with propionyl chloride led to the formation of trans‐[Pd(SeC(O)CH2CH3)2(PnBu3)2] ( 3 ), a trans‐bis(selenocarboxylato) palladium complex and thus established a new method for the formation of this type of complex. Complex 3 was characterized by 1H, 13C{1H}, 31P{1H} and 77Se{1H} NMR spectroscopy and a single‐crystal X‐ray structure analysis.  相似文献   

13.
Photoinduced electron‐transfer processes within a precatalyst for intramolecular hydrogen evolution [(tbbpy)2Ru(tpphz)PdCl2]2+ ( RuPd ; tbbpy=4,4′‐di‐tert‐butyl‐2,2′‐bipyridine, tpphz=tetrapyrido[3,2‐a:2′,3′c:3′′,2′′,‐h:2′′′,3′′′‐j]phenazine) have been studied by resonance Raman and ultrafast time‐resolved absorption spectroscopy. By comparing the photophysics of the [(tbbpy)2Ru(tpphz)]2+ subunit Ru with that of the supramolecular catalyst RuPd , the individual electron‐transfer steps are assigned to kinetic components, and their dependence on solvent is discussed. The resonance Raman data reveal that the initial excitation of the molecular ensemble is spread over the terminal tbbpy and the tpphz ligands. The subsequent excited‐state relaxation of both Ru and RuPd on the picosecond timescale involves formation of the phenazine‐centered intraligand charge‐transfer state, which in RuPd precedes formation of the Pd‐reduced state. The photoreaction in the heterodinuclear supramolecular complex is completed on a subnanosecond timescale. Taken together, the data indicate that mechanistic investigations must focus on potential rate‐determining steps other than electron transfer between the photoactive center and the Pd unit. Furthermore, structural variations should be directed towards increasing the directionality of electron transfer and the stability of the charge‐separated states.  相似文献   

14.
The photodynamics of the recombinant rhodopsin fragment of the histidine kinase rhodopsin HKR1 from Chlamydomonas reinhardtii was studied by absorption and fluorescence spectroscopy. The retinal cofactor of HKR1 exists in two Schiff base forms RetA and RetB. RetA is the deprotonated 13‐cis‐retinal Schiff base (RSB) absorbing in the UVA spectral region. RetB is the protonated all‐trans RSB absorbing in the blue spectral region. Blue light exposure converts RetB fully to RetA. UVA light exposure converts RetA to RetB and RetB to RetA giving a mixture determined by their absorption cross sections and their conversion efficiencies. The quantum efficiencies of conversion of RetA to RetB and RetB to RetA were determined to be 0.096 ± 0.005 and 0.405 ± 0.01 respectively. In the dark thermal equilibration between RetA and RetB with dominant RetA content occurred with a time constant of about 3 days at room temperature. The fluorescence emission behavior of RetA and RetB was studied, and fluorescence quantum yields of ?F(RetA) = 0.00117 and ?F(RetB) = 9.4 × 10?5 were determined. Reaction coordinate schemes of the photodynamics are developed.  相似文献   

15.
The structure of trans‐[Cr(Me2tn)2Cl2]2ZnCl4 (Me2tn = 2,2‐dimethylpropane‐1,3‐diamine) was determined by a single‐crystal X‐ray diffraction study at 173 K. The analysis reveals that there are three crystallographically independent chromium(III) complex cations in the title compound. The chromium(III) atoms are coordinated by four nitrogen atoms of Me2tn and two chlorine atoms in a trans arrangement, displaying a distorted octahedral geometry. The two six‐membered chelate rings in three complex cations are oriented in an anti chair–chair conformation with respect to each other. The Cr–N and Cr–Cl bond lengths average 2.0862(2) and 2.3112(6) Å, respectively. The ZnCl42– have slightly distorted tetrahedral arrangement with Zn–Cl lengths and the Cl–Zn–Cl angles are influenced by hydrogen bonding. The resolved absorption maxima in the electronic d–d spectrum were fitted with a secular determinant for a quartet energy state of the d3 configuration in a tetragonal field. It is confirmed that the nitrogen atoms of the Me2tn ligand are strong σ donors, but the chloro ligands have weak σ‐ and π donor properties toward the chromium(III) ion.  相似文献   

16.
Complete and unambiguous 1H and 13C NMR chemical shift assignments for all‐trans‐retinal, 13‐cis‐retinal, 11‐cis‐retinal and 9‐cis‐retinal (1–4) have been established by means of two‐dimensional COSY, HSQC, HMBC and NOESY spectroscopic experiments. Copyright © 2013 John Wiley & Sons, Ltd.  相似文献   

17.
The RPE65 protein of the retinal pigment epithelium (RPE) enables the conversion of retinyl esters to the visual pigment chromophore 11‐cis retinal. Fresh 11‐cis retinal is generated from retinyl esters following photoisomerization of the visual pigment chromophore to all‐trans during light detection. Large amounts of esters accumulate in Rpe65?/? mice, indicating their continuous formation when 11‐cis retinal generation is blocked. We hypothesized that absence of light, by limiting the conversion of esters to 11‐cis retinal, would also result in the build‐up of retinyl esters in the RPE of wild‐type mice. We used HPLC to quantify ester levels in organic extracts of the RPE from wild‐type and Rpe65?/? mice. Retinyl ester levels in Sv/129 wild‐type mice that were dark adapted for various intervals over a 4‐week period were similar to those in mice raised in cyclic light. In C57BL/6 mice however, which contain less Rpe65 protein, dark adaptation was accompanied by an increase in ester levels compared to cyclic light controls. Retinyl ester levels were much higher in Rpe65?/? mice compared to wild type and kept increasing with age. The results suggest that the RPE65 role in retinyl ester homeostasis extends beyond enabling the formation of 11‐cis retinal.  相似文献   

18.
Anthocyanins are a class of phytochemicals that confer color to flowers, fruits, vegetables and leaves. They are part of our regular diet and serve as dietary supplements because of numerous health benefits, including improved vision. Recent studies have shown that the anthocyanin cyanidin‐3‐O‐glucoside (C3G) increased regeneration of the dim‐light photoreceptor rhodopsin (Matsumoto et al. [2003] J. Agric. Food Chem., 51 , 3560–3563). In an accompanying study (Yanamala et al. [2009] Photochem. Photobiol.), we show that C3G directly binds to rhodopsin in a pH‐dependent manner. In this study, we investigated the functional consequences of C3G binding to rhodopsin. As observed previously in rod outer segments, regeneration of purified rhodopsin in detergent micelles is also accelerated in the presence of C3G. Thermal denaturation and stability studies using circular dichroism, fluorescence and UV/visible absorbance spectroscopy show that C3G exerts a destabilizing effect on rhodopsin structure while it only modestly alters G‐protein activation and the rates at which the light‐activated Metarhodopsin II state decays to opsin and free retinal. These results indicate that the mechanism of C3G‐enhanced regeneration may be based on changes in opsin structure promoting access to the retinal binding pocket.  相似文献   

19.
The photoisomerization of retinal   总被引:1,自引:0,他引:1  
Abstract— –Quantum efficiencies have been measured for the photoisomerization of four stereoisomers of retinal (all-trans, 13-cis, 11 cis, and 9-cis) in two solvents at different wavelengths of irradiation and at various temperatures. In heane at 25°C the quantum efficiencies for isomerization at 365 nm are: 9-cis to trans, 0.5; 13-cis to trans, 0.4; 11-cis to trans, 0.2; all-trans to monocis isomers, 0.2-0.06, depending upon assumptions made regarding the stereo-isomeric composition of the product. These values vary somewhat with the wavelength of the irradiating light. The quantum efficiency for the photoisomerization of all-trans retinal in hexane decreases by a factor of 30 when the temperature is lowered from 25° to – 65°C; the activation energy for this photoisomerization is about 5 kcal/mole. The quantum efficiencies for the isomerization of the monocis isomers to all-trans retinal in hexane are virtually independent of temperature. In ethanol the rates of photoisomerization from trans to cis or cis to trans depend only slightly on the temperature between 25° and – 65°C. The photosensitivities of the stereoisomers of retinal are of the same order of magnitude as those of the retinylidene chromophores of rhodopsin (11 -cis), metarhodopsin I (all-trans), and isorhodopsin (9-cis); but it is not yet possible to derive the photochemistry of rhodopsin uniquely and quantitatively from that of retinal.  相似文献   

20.
The retinal chromophore (11Z)‐3‐diazo‐4‐oxoretinal ( 1 ) with two photo‐labile moieties has been synthesized by semi‐hydrogenation of an 11‐yne precursor with activated Zn in aqueous media. Incorporation of 1 into opsin yielded diazoketo rhodopsin (DK‐Rh), which, upon bleaching, gave rise to intermediates batho‐Rh, lumi‐Rh, meta‐Rh, and meta‐II‐Rh corresponding to those of native Rh but at lower temperatures. Photoaffinity labeling of DK‐Rh and these bleaching intermediates showed that the ionone ring cross‐linked to Trp265 of helix F in DK‐Rh and batho intermediate, and to Ala169 of helix D in lumi, meta‐I, and meta‐II intermediates. These results demonstrate the occurrence of large conformational changes along the visual transduction path, which, in turn, is responsible for activation of the G‐protein.  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