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 46 毫秒
1.
The early receptor potential (ERP) was recorded intracellularly from Limulus ventral photoreceptors. The ERP in cells dissected under red light was altered by exhaustive illumination. No recovery to the original wafeform was observed, even after 1 h in the dark. The ERP waveform could be further altered by chromatic adaptation or by changes in pH. The results indicate that at pH 7.8 there are two interconvertible pigment states with only slightly different lambdamax, whereas at pH 9.6 there are two interconvertible states with very different lambdamax. Under all conditions studied the ERPs were almost identical with those previously obtained in squid retinas. This strongly suggests that light converts Limulus rhodopsin to a stable photoequilibrium mixture of rhodopsin to a stable photoequilibrium mixture of rhodopsin and metarhodopsin and that, as in squid, the lambdamax of metarhodopsin depends on pH. This conversion at pH 7.8 is associated with a small (0.7 log unit) decrease in the maximum sensitivity of the late receptor potential. Thus the component of adaptation linked to changes in rhodopsin concentration is unimportant in comparison to the "neural" component.  相似文献   

2.
Dark regeneration of rhodopsin in crayfish photoreceptors   总被引:1,自引:1,他引:0       下载免费PDF全文
The eyes of crayfish were exposed to lights of known spectral composition, and the course of regeneration was followed in the dark by measuring the content of rhodopsin and metarhodopsin in single rhabdoms isolated at various times after the adaptation, using an assay that is based on the fluorescence of metarhodopsin. Complete recovery requires several days in the dark after intense adaptation to orange light, but requires less than 2 d after blue light exposure. Following an orange light exposure with blue produces recovery kinetics characteristic of the blue light exposure alone. This quickening of recovery occurs whether the receptors are exposed to blue light either immediately or many hours after the original exposure to orange. Conversely, following blue light adaptation with orange leads to slow recovery, which is characteristic of orange alone. Recovery from long-wavelength adaptation is slower principally because many rhabdoms seem to delay the onset of regeneration. We suggest that the regeneration system is itself photosensitive, and after orange light adaptation the supply of active chromophore (presumably 11-cis retinal) limits the rate of recovery. Once started, recovery proceeds slowly and continuously, and the total pigment concentration (rhodopsin plus metarhodopsin) in the rhabdomeric membrane remains approximately constant. Within hours after intense adapting exposures, the rhabdoms become altered in appearance, the surfaces become coated with accessory pigment, and the bands of microvilli are less distinct. These changes persist until recovery of rhodopsin proceeds, which suggests that visual pigment regeneration results from addition of newly synthesized rhodopsin associated with membrane turn-over.  相似文献   

3.
1. The primary aim of the experiments described in this article was to localize the origin of the complex fluorescence in the compound eye of flies. The eye tissue was dissected and the fluorescence from cells and cell organelles was recorded by microspectrofluorometry. Using this technique, fluorophore systems were detected in the rhabdomeres, Semper cells and corneal lenses. The fluorophore systems are photoreconvertible by UV and blue light. 2. The fluorophore systems in the rhabdomeres and Semper cells are similar. The intensity of the fluorescence from the microvilli is enhanced up to 29 X by adaptation to UV light. The enhancement is inversely related to the rhodopsin content in the microvilli, indicating that the chromophoric group of the fluorophore is not a vitamin A derivative. 3. The enhancement of the fluorescence by UV light strongly depends on pH, suggesting that the photoreconvertible fluorophore systems in the microvilli and Semper cells are photosensitive redox pigments. These redox systems are probably located in the membranes of the microvilli in the photoreceptors, and in the endoplasmic reticulum of the Semper cells, or they are coupled to filaments in the cytoskeleton of both cell types. 4. Preliminary reaction schemes for the photoreactions based on the recorded excitation and emission spectra and photokinetics were developed. A primary pigment in the microvillous structure, AR, or in organelles in the Semper cells, AS, is converted by UV light into an excited state AR* or AS*, which either relaxes to the primary pigment by photon emission, or converts into an intermediate X, which by proton uptake changes into stable products, BR or BS. Blue illumination converts BR and BS into the excited states BR* and BS*, which either relax by photon emission to BR or BS, or convert into an intermediate Y, which after deprotonation reconverts into the primary pigment AR or AS. 5. Estimation of the molecular density showed that the concentration of the fluorophore in the microvilli presumably is almost equal to maximal rhodopsin concentration. The high density suggests that the fluorophores have a specific function in transduction or adaptation of the visual process.  相似文献   

4.
The early receptor current (ERC) represents molecular charge movement during rhodopsin conformational dynamics. To determine whether this time-resolved assay can probe various aspects of structure-function relationships in rhodopsin, we first measured properties of expressed normal human rhodopsin with ERC recordings. These studies were conducted in single fused giant cells containing on the order of a picogram of regenerated pigment. The action spectrum of the ERC of normal human opsin regenerated with 11-cis-retinal was fit by the human rhodopsin absorbance spectrum. Successive flashes extinguished ERC signals consistent with bleaching of a rhodopsin photopigment with a normal range of photosensitivity. ERC signals followed the univariance principle since millisecond-order relaxation kinetics were independent of the wavelength of the flash stimulus. After signal extinction, dark adaptation without added 11-cis-retinal resulted in spontaneous pigment regeneration from an intracellular store of chromophore remaining from earlier loading. After the ERC was extinguished, 350-nm flashes overlapping metarhodopsin-II absorption promoted immediate recovery of ERC charge motions identified by subsequent 500-nm flashes. Small inverted R(2) signals were seen in response to some 350-nm flashes. These results indicate that the ERC can be photoregenerated from the metarhodopsin-II state. Regeneration with 9-cis-retinal permits recording of ERC signals consistent with flash activation of isorhodopsin. We initiated structure-function studies by measuring ERC signals in cells expressing the D83N and E134Q mutant human rhodopsin pigments. D83N ERCs were simplified in comparison with normal rhodopsin, while E134Q ERCs had only the early phase of charge motion. This study demonstrates that properties of normal rhodopsin can be accurately measured with the ERC assay and that a structure-function investigation of rapid activation processes in analogue and mutant visual pigments is feasible in a live unicellular environment.  相似文献   

5.
Dark adaptation requires timely deactivation of phototransduction and efficient regeneration of visual pigment. No previous study has directly compared the kinetics of dark adaptation with rates of the various chemical reactions that influence it. To accomplish this, we developed a novel rapid-quench/mass spectrometry-based method to establish the initial kinetics and site specificity of light-stimulated rhodopsin phosphorylation in mouse retinas. We also measured phosphorylation and dephosphorylation, regeneration of rhodopsin, and reduction of all-trans retinal all under identical in vivo conditions. Dark adaptation was monitored by electroretinography. We found that rhodopsin is multiply phosphorylated and then dephosphorylated in an ordered fashion following exposure to light. Initially during dark adaptation, transduction activity wanes as multiple phosphates accumulate. Thereafter, full recovery of photosensitivity coincides with regeneration and dephosphorylation of rhodopsin.  相似文献   

6.
The rhabdoms of Euphausia superba contain one digitonin-extractable rhodopsin, lambda max 485 nm. The rhodopsin undergoes unusual pH- dependent spectral changes: above neutrality, the absorbance decreases progressively at 485 nm and rises near 370 nm. This change is reversible and appears to reflect an equilibrium between a protonated and an unprotonated form of the rhodopsin Schiff-base linkage. Near neutral pH and at 10 degrees C, the rhodopsin is partiaLly converted by 420-nm light to a stable 493-nm metarhodopsin. The metarhodopsin is partially photoconverted to rhodopsin by long-wavelength light in the absence of NH2OH; in the presence of NH2OH, it is slowly converted to retinal oxime and opsin. The rhodopsin of Meganyctiphanes norvegica measured in fresh rhabdoms by microspectrophotometry has properties very similar to those of the extracted rhodopsin of E. superba. Its lambda max is 488 nm and it is partially photoconverted by short wavelength irradiation to a stable photoconvertible metarhodopsin similar to that of E. superba. In the presence of light and NH2OH, the M. norvegica metarhodopsin is converted to retinal oxime and opsin. Our results indicate that previous determinations of euphausiid rhodopsin absorbance spectra were incorrect because of accessory pigment contamination.  相似文献   

7.
Summary Dark adaptation of living lobsters was measured by recording the ERG at several temperatures in the range 5–20 °C following adapting flashes that convert about 70% of the rhodopsin to metarhodopsin. Recovery of log threshold is rapid, and at 10–20° is nearly complete in 10 min. Only at 5 °C is dark adaptation significantly slowed. Comparison of dark adaptation with data on regeneration of pigment (Bruno et al., 1977) is consistent with the hypothesis that as rhodopsin concentration rises and falls, its only effect on sensitivity is to alter the probability of quantum catch. This interpretation is further bolstered by observations on winter lobsters that have a 70% deficiency of rhodopsin without the concomitant increase in metarhodopsin that accompanies light adaptation. No effect of metarhodopsin on sensitivity was detected. These experiments support the growing body of evidence indicating that the relationship between rhodopsin concentration and log threshold is fundamentally different in the rhabdomeric photoreceptors of invertebrates and the rods and cones of vertebrates.This work was supported by USPHS research grant EY 00222 to Yale University. S.N.B. was aided by NIH Postdoctoral Fellowship EY 52378, by funds made available through the Unidel Foundation, and by a grant from the University of Delaware Research Foundation.  相似文献   

8.
Purified preparations of cattle rhodopsin have been titrated to various pH, irradiated, and the pH changes followed thereafter until completed. In this way we have obtained the titration curves of rhodopsin, of the immediate product of irradiation, measured within 30 seconds; and of the final product of irradiation (opsin). The rhodopsin preparations display about 54 titratable groups per mole of pigment: about 34 base-binding and 20 acid-binding groups. In default of an absolute purification, one cannot be sure that all of these go with rhodopsin itself. Exposure to light induces an immediate rise of pH between pH 2 and 8, maximal at about pH 5. This—followed by its slow partial or complete reversal—is the only change of pH in the physiological range (6–7). It involves the exposure of 1 new acid-binding group per mole of rhodopsin with pK about 6.6, close therefore to that of the imidazole group of histidine. At acid and alkaline pH this immediate change is followed by slower changes, occupying up to 40 minutes at 20°C. These changes are always in the direction of neutrality. They involve increases of 5 to 6 moles acid bound at acid pH, and 7 moles base bound at alkaline pH. They are associated with the irreversible denaturation of opsin in acid and alkaline solution, as evidenced by loss of its capacity to regenerate rhodopsin. Such frank denaturation procedures as the exposure of rhodopsin to alkali or heat in the dark result in comparable acid-base changes.  相似文献   

9.
Exposure of an intact vertebrate eye to light bleaches the rhodopsin in the photoreceptor outer segments in spatially nonuniform patterns. Some axial bleaching patterns produced in toad rods were determined using microspectrophotometric techniques. More rhodopsin was bleached at the base of the outer segment than at the distal tip. The shape of the bleaching gradient varied with the extent of bleach and with the spectral content of the illuminant. Monochromatic light at the lambda max of the rhodopsin gave rise to the steepest bleaching gradients and induced the greatest changes in the form of the gradient with increasing extent of bleach. These results were consistent with a mathematical model for pigment bleaching in an unstirred sample. The model did not fit bleaching patterns resulting from special lighting conditions that promoted the photoregeneration of rhodopsin from the intermediates of bleaching. Prolonged light adaptation of toads could also produce axial rhodopsin gradients that were not fit by the bleaching model. Under certain conditions the axial gradient of rhodopsin in a rod outer segment reversed with time in the light: the rhodopsin content became highest at the base. This result could be explained by an interaction between the pattern of bleaching and the intracellular topography of regeneration.  相似文献   

10.
Cyclic GMP has been implicated in controlling the light-regulated conductance of rod photoreceptors of the vertebrate retina. However, there is little direct evidence correlating changes in cGMP concentration with the light-regulated permeability mechanism in living cells. A preparation of intact frog rod outer segments suspended in a Ringer's medium containing low Ca2+ has been used to demonstrate that initial changes in total cellular cGMP concentration parallel changes in the light-regulated membrane current over a wide range of light intensities. At light intensities bleaching from 160 to 5.6 X 10(6) rhodopsin molecules/rod/s, decreases in the response latency for the cGMP kinetics parallel decreases in the latent period of the electrical response. Further, changes in the rate of the cGMP decrease parallel the rate of membrane current suppression as the light intensity is varied. Up to 10(5) cGMP molecules are hydrolyzed per photolyzed rhodopsin, consistent with in vitro studies showing that each bleached rhodopsin can activate over 100 phosphodiesterase molecules. Addition of the Ca2+ ionophore, A23187, does not affect the initial kinetics of the cGMP decrease or of the electrical response, excluding a direct role for Ca2+ in the initial events of phototransduction. These results are consistent with cGMP being the intracellular messenger that links rhodopsin isomerization with changes in membrane permeability upon illumination. It is unlikely, however, that light-induced changes in total cGMP concentration are the sole regulators of membrane current. This is suggested by several observations: at bright light intensities, the subsecond light-induced cGMP decrease is essentially complete prior to complete suppression of membrane current; maximal light-induced decreases in cGMP concentration occur at all light intensities tested, whereas the extent of membrane current suppression varies over the same range of light intensities; changing the external Ca2+ concentration from 1 mM to 10 nM in the dark causes an increase in membrane current that is significantly more rapid than corresponding changes in cGMP concentration. Thus, light-induced changes in total cellular cGMP concentration correlate with some, but not all, aspects of the visual excitation process in vertebrate photoreceptors.  相似文献   

11.
R N Frank  S M Buzney 《Biochemistry》1975,14(23):5110-5117
Partial separation of protein kinase activity from rhodopsin in isolated bovine retinal photoreceptor outer segments was accomplished by mild ultrasonic treatment followed by ultracentrifugation. Residual kinase activity in the rhodopsin-rich sediment was destroyed by chemical denaturation which did not affect the spectral properties of the rhodopsin. The retinal outer segment kinase was found to be specific for rhodopsin, since in these preparations it alone of several bovine protein kinases was capable of phosphorylating rhodopsin in the light. The phosphorylation reaction apparently requires a specific conformation of the rhodopsin molecule since it is abolished by heat denaturation of rhodopsin, and it is greatly reduced or abolished by treatment of the visual pigment protein with potassium alum after the rhodopsin has been "bleached" by light. When kinase and rhodopsin or opsin fractions were prepared from dark-adapted and bleached outer segments and the resultant fractions were mixed in various combinations of bleached and unbleached preparations, the observed pattern of light-activated phosphorylation was consistent only with the interpretation that a conformational change in the rhodopsin molecule in the light exposes a site on the visual pigment protein to the kinase and ATP. These results rule out the possibility of a direct or indirect (rhodopsin-mediated) light activation of the kinase. Finally, phosphorylation of retinal outer segment protein in monochromatic lights of various wavelengths followed by sodium dodecyl sulfate-polyacrylamide gel electrophoresis indicates that both rhodopsin and the higher molecular weight visual pigment protein reported by several laboratories have the same action spectrum for phosphorylation. This result is consistent with the suggestion that the higher molecular weight species is a rhodopsin dimer.  相似文献   

12.
An apparatus was constructed in order to record continuously and simultaneously changes in extinction and electrical conductance of rhodopsin solutions. With this apparatus, changes in electrical conductance on exposing rhodopsin to light were investigated. On illumination solutions of rhodopsin revealed a conductance change so long as they preserved their photosensitivity. The conductance change begins almost immediately upon illumination and is almost proportional to the amount of rhodopsin decomposed, continuing until rhodopsin is converted to indicator yellow. Near pH 7 the conductance is apt to increase slightly, while it decreases considerably outside the range of pH 6–9, being accompanied by a pH change towards neutrality. The conductance change is regarded as an essential property of rhodopsin, because it occurs in aqueous suspension as well as in digitonin solution; it may be caused by hydrogen or hydroxyl ions and some other conductive substances. It is also noteworthy that the petroleum ether-soluble component of the rod outer segments—presumably the lipide—tends to increase the conductance change. In suspensions of rod outer segments and retinal homogenates, the conductance increases on illumination irrespective of pH: this may be due to secondary reactions following the photic reaction of rhodopsin. We shall discuss the significance of the conductance change in relation to the initiation of visual excitation.  相似文献   

13.
Tsutsui K  Imai H  Shichida Y 《Biochemistry》2007,46(21):6437-6445
A visual pigment consists of an opsin protein and a chromophore, 11-cis-retinal, which binds to a specific lysine residue of opsin via a Schiff base linkage. The Schiff base chromophore is protonated in pigments that absorb visible light, whereas it is unprotonated in ultraviolet-absorbing visual pigments (UV pigments). To investigate whether an unprotonated Schiff base can undergo photoisomerization as efficiently as a protonated Schiff base in the opsin environment, we measured the quantum yields of the bovine rhodopsin E113Q mutant, in which the Schiff base is unprotonated at alkaline pH, and the mouse UV pigment (mouse UV). Photosensitivities of UV pigments were measured by irradiation of the pigments followed by chromophore extraction and HPLC analysis. Extinction coefficients were estimated by comparing the maximum absorbances of the original pigments and their acid-denatured states. The quantum yield of the bovine rhodopsin E113Q mutant at pH 8.2, where the Schiff base is unprotonated, was significantly lower than that of wild-type rhodopsin, whereas the mutant gave a quantum yield almost identical to that of the wild type at pH 5.5, where the Schiff base is protonated. These results suggest that Schiff base protonation plays a role in increasing quantum yield. The quantum yield of mouse UV, which has an unprotonated Schiff base chromophore, was significantly higher than that of the unprotonated form of the rhodopsin E113Q mutant, although it was still lower than the visible-absorbing pigments. These results suggest that the mouse UV pigment has a specific mechanism for the efficient photoisomerization of its unprotonated Schiff base chromophore.  相似文献   

14.
Anabaena sensory rhodopsin is a seven transmembrane protein that uses all-trans/13-cis retinal as a chromophore. About 22 residues in the retinal-binding pocket of microbial rhodopsins are conserved and important to control the quality of absorbing light and the function of ion transport or sensory transduction. The absorption maximum is 550 nm in the presence of all-trans retinal at dark. Here, we mutated Pro206 to Glu or Asp, of which the residue is conserved as Asp among all other microbial rhodopsins, and the absorption maximum and pKa of the proton acceptor group were measured by absorption spectroscopy at various pHs. Anabaena rhodopsin was expressed best in Escherichia coli in the absence of extra leader sequence when exogenous all-trans retinal was added. The wild-type Anabaena rhodopsin showed small absorption maximum changes between pH 4 and 11. In addition, Pro206Asp showed 46 nm blue-shift at pH 7.0. Pro206Glu or Asp may change the contribution to the electron distribution of the retinal that is involved in the major role of color tuning for this pigment. The critical residue Ser86 (Asp 96 position in bacteriorhodopsin: proton donor) for the pumping activity was replaced with Asp, but it did not change the proton pumping activity of Anabaena rhodopsin.  相似文献   

15.
Nontransducing rhodopsin   总被引:1,自引:0,他引:1       下载免费PDF全文
Rhodopsin is converted by light to an active photoproduct that triggers the transduction cascade. The active photoproduct must then be inactivated by some kind of chemical modification. The question addressed here is whether photoconversion of the inactive photoproduct to rhodopsin creates a modified form of rhodopsin that is unable to support transduction. This question was investigated in ultraviolet receptors of Limulus median eye by measuring the relative quantum efficiency of excitation after photoregeneration of rhodopsin from the inactive photoproduct. The results show that when this newly created rhodopsin absorbs a photon, no receptor potential is generated; i.e., the pigment is nontransducing. A dark process requiring 30-60 min returns rhodopsin to its transducing form.  相似文献   

16.
Resonance Raman excitation profiles have been measured for the bovine visual pigment rhodopsin using excitation wavelengths ranging from 457.9 to 647.1 nm. A complete Franck-Condon analysis of the absorption spectrum and resonance Raman excitation profiles has been performed using an excited-state, time-dependent wavepacket propagation technique. This has enabled us to determine the change in geometry upon electronic excitation of rhodopsin's 11-cis-retinal protonated Schiff base chromophore along 25 normal coordinates. Intense low-frequency Raman lines are observed at 98, 135, 249, 336, and 461 cm-1 whose intensities provide quantitative, mode-specific information about the excited-state torsional deformations that lead to isomerization. The dominant contribution to the width of the absorption band in rhodopsin results from Franck-Condon progressions in the 1,549 cm-1 ethylenic normal mode. The lack of vibronic structure in the absorption spectrum is shown to be caused by extensive progressions in low-frequency torsional modes and a large homogeneous linewidth (170 cm-1 half-width) together with thermal population of low-frequency modes and inhomogeneous site distribution effects. The resonance Raman cross-sections of rhodopsin are unusually weak because the excited-state wavepacket moves rapidly (approximately 35 fs) and permanently away from the Franck-Condon geometry along skeletal stretching and torsional coordinates.  相似文献   

17.
Iodopsin   总被引:5,自引:0,他引:5       下载免费PDF全文
The iodopsin system found in the cones of the chicken retina is identical with the rhodopsin system in its carotenoids. It differs only in the protein-the opsin -with which carotenoid combines. The cone protein may be called photopsin to distinguish it from the scotopsins of the rods. Iodopsin bleaches in the light to a mixture of photopsin and all-trans retinene. The latter is reduced by alcohol dehydrogenase and cozymase to all-trans vitamin A(1). Iodopsin is resynthesized from photopsin and a cis isomer of vitamin A, neovitamin Ab or the corresponding neoretinene b, the same isomer that forms rhodopsin. The synthesis of iodopsin from photopsin and neoretinene b is a spontaneous reaction. A second cis retinene, isoretinene a, forms iso-iodopsin (lambda(max) 510 mmicro). The bleaching of iodopsin in moderate light is a first-order reaction (Bliss). The synthesis of iodopsin from neoretinene b and opsin is second-order, like that of rhodopsin, but is very much more rapid. At 10 degrees C. the velocity constant for iodopsin synthesis is 527 times that for rhodopsin synthesis. Whereas rhodopsin is reasonably stable in solution from pH 4-9, iodopsin is stable only at pH 5-7, and decays rapidly at more acid or alkaline reactions. The sulfhydryl poison, p-chloromercuribenzoate, blocks the synthesis of iodopsin, as of rhodopsin. It also bleaches iodopsin in concentrations which do not attack rhodopsin. Hydroxylamine also bleaches iodopsin, yet does not poison its synthesis. Hydroxylamine acts by competing with the opsins for retinene. It competes successfully with chicken, cattle, or frog scotopsin, and hence blocks rhodopsin synthesis; but it is less efficient than photopsin in trapping retinene, and hence does not block iodopsin synthesis. Though iodopsin has not yet been prepared in pure form, its absorption spectrum has been computed by two independent procedures. This exhibits an alpha-band with lambda(max) 562 mmicro, a minimum at about 435 mmicro, and a small beta-band in the near ultraviolet at about 370 mmicro. The low concentration of iodopsin in the cones explains to a first approximation their high threshold, and hence their status as organs of daylight vision. The relatively rapid synthesis of iodopsin compared with rhodopsin parallels the relatively rapid dark adaptation of cones compared with rods. A theoretical relation is derived which links the logarithm of the visual sensitivity with the concentration of visual pigment in the rods and cones. Plotted in these terms, the course of rod and cone dark adaptation resembles closely the synthesis of rhodopsin and iodopsin in solution. The spectral sensitivities of rod and cone vision, and hence the Purkinje phenomenon, have their source in the absorption spectra of rhodopsin and iodopsin. In the chicken, for which only rough spectral sensitivity measurements are available, this relation can be demonstrated only approximately. In the pigeon the scotopic sensitivity matches the spectrum of rhodopsin; but the photopic sensitivity is displaced toward the red, largely or wholly through the filtering action of the colored oil globules in the pigeon cones. In cats, guinea pigs, snakes, and frogs, in which no such colored ocular structures intervene, the scotopic and photopic sensitivities match quantitatively the absorption spectra of rhodopsin and iodopsin. In man the scotopic sensitivity matches the absorption spectrum of rhodopsin; but the photopic sensitivity, when not distorted by the yellow pigmentations of the lens and macula lutea, lies at shorter wave lengths than iodopsin. This discrepancy is expected, for the human photopic sensitivity represents a composite of at least three classes of cone concerned with color vision.  相似文献   

18.
Photoreceptor potentials were recorded extracellularly from the aspartate-treated, isolated retina of the skate (Raja oscellata and R. erinacea), and the effects of externally applied retinal were studied both electrophysiologically and spectrophotometrically. In the absence of applied retinal, strong light adaptation leads to an irreversible depletion of rhodopsin and a sustained elevation of receptor threshold. For example, after the bleaching of 60% of the rhodopsin initially present in dark-adapted receptors, the threshold of the receptor response stabilizes at a level about 3 log units above the dark-adapted value. The application of 11-cis retinal to strongly light-adapted photoreceptors induces both a rapid, substantial lowering of receptor threshold and a shift of the entire intensity-response curve toward greater sensitivity. Exogenous 11-cis retinal also promotes the formation of rhodopsin in bleached photoreceptors with a time-course similar to that of the sensitization measured electrophysiologically. All-trans and 13-cis retinal, when applied to strongly light-adapted receptors, fail to promote either an increase in receptor sensitivity or the formation of significant amounts of light-sensitive pigment within the receptors. However, 9-cis retinal isin. These findings provide strong evidence that the regeneration of visual pigment in the photoreceptors directly regulates the process of photochemical dark adaptation.  相似文献   

19.
The signaling state metarhodopsin II of the visual pigment rhodopsin decays to the apoprotein opsin and all-trans retinal, which are then regenerated to rhodopsin by the visual cycle. Opsin is known to have at neutral pH only a small residual constitutive activity toward its G protein transducin, which is thought to play a considerable role in light adaptation (bleaching desensitization). In this study we show with Fourier-transform infrared spectroscopy that after metarhodopsin II decay, opsin exists in two conformational states that are in a pH-dependent equilibrium at 30 degrees C with a pK of 4.1 in the presence of hydroxylamine scavenging the endogenous all-trans retinal. Despite the lack of the native agonist in its binding pocket, the low pH opsin conformation is very similar to that of metarhodopsin II and is likewise stabilized by peptides derived from rhodopsin's cognate G protein, transducin. The high pH form, on the other hand, has some conformational similarity to the inactive metarhodopsin I state. We therefore conclude that the opsin apoprotein displays intrinsic conformational states that are merely modulated by bound all-trans retinal.  相似文献   

20.
C Longstaff  R R Rando 《Biochemistry》1987,26(19):6107-6113
Bacteriorhodopsin (bR) in purple membranes was permethylated with formaldehyde and pyridine-borane with the incorporation of approximately 12 methyl groups. This new pigment, PMbR, absorbed light in the dark-adapted state with a lambda max at 558 nm, virtually the same as that of bR. Light adaptation of PMbR produced a lambda max of 564 nm with a slightly elevated epsilon. Similar changes occurred with bR. When incorporated into asolectin vesicles, PMbR was able to pump protons in the light with an efficiency similar to that of bR itself. Bleaching of PMbR exposed its active site lysine residue, which was monomethylated to form active site methylated bR (AMbR) after regeneration with all-trans-retinal. This blue pigment, which is a cyanopsin rather than a rhodopsin, showed an extraordinary red shift, absorbing light with a lambda max of 620 nm in the dark-adapted state. Light adaptation of AMbR resulted in a spectral shift to 616 nm with a decrease in epsilon. This change was completely reversible in the dark. This shift was interpreted to mean that an L-like intermediate was accumulating, as would be expected if deprotonation of the protonated Schiff base could not occur to produce the M intermediate. Furthermore, when incorporated into asolectin vesicles, AMbR proved incapable of pumping protons in the light. It was concluded from these experiments that deprotonation of the Schiff base of bR is obligate for light-induced proton pumping.  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