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 15 毫秒
1.
The deactivation of the bovine G-protein-coupled receptor, rhodopsin, is a two-step process consisting of the phosphorylation of specific serine and threonine residues in the cytoplasmic tail of rhodopsin by rhodopsin kinase. Subsequent binding of the regulatory protein arrestin follows this phosphorylation. Previous results find that at least three phosphorylatable sites on the rhodopsin tail (T340) and at least two of the S338, S334, or S343 sites are needed for complete arrestin-mediated deactivation. Thus, to elucidate the details of the interaction between rhodopsin with arrestin, we have employed both a computational and an in vitro experimental approach. In this work, we first simulated the interaction of the carboxy tail of rhodopsin with arrestin using a Monte Carlo simulated annealing method. Since at this time phosphorylation of specific serines and threonines is not possible in our simulations, we substitute either aspartic or glutamic acid residues for the negatively charged phosphorylated residues required for binding. A total of 17 simulations were performed and analysis of this shows specific charge-charge interactions of the carboxy tail of rhodopsin with arrestin. We then confirmed these computational results with assays of comparable constructed rhodopsin mutations using our in vitro assay. This dual computational/experimental approach indicates that sites S334, S338, and T340 in rhodopsin and K14 and K15 on arrestin are indeed important in the interaction of rhodopsin with arrestin, with a possible weaker S343 (rhodopsin)/K15 (arrestin) interaction.  相似文献   

2.
Brannock MT  Weng K  Robinson PR 《Biochemistry》1999,38(12):3770-3777
Many recent reports have demonstrated that rhodopsin's carboxyl-terminal serine residues are the main targets for phosphorylation by rhodopsin kinase. Phosphorylation at the serines would therefore be expected to promote high-affinity arrestin binding. We have examined the roles of the carboxyl serine and threonine residues during arrestin-mediated deactivation of rhodopsin using an in vitro transducin activation assay. Mutations were introduced into a synthetic bovine rhodopsin gene and expressed in COS-7 cells. Individual serine and threonine residues were substituted with neutral amino acids. The ability of the mutants to act as substrates for rhodopsin kinase was analyzed. The effect of arrestin on the activities of the phosphorylated mutant rhodopsins was measured in a GTPgammaS binding assay involving purified bovine arrestin, rhodopsin kinase, and transducin. A rhodopsin mutant lacking the carboxyl serine and threonine residues was not phosphorylated by rhodopsin kinase, demonstrating that phosphorylation is restricted to the seven putative phosphorylation sites. A rhodopsin mutant possessing a single phosphorylatable serine at 338 demonstrated no phosphorylation-dependent quench by arrestin. These results suggest that singly phosphorylated rhodopsin is deactivated through a mechanism that does not involve arrestin. Analysis of additional mutants revealed that the presence of threonine in the carboxyl tail of rhodopsin provides for greater arrestin-mediated quench than does serine. These results suggest that phosphorylation site selection could serve as a mechanism to modulate the ability of arrestin to quench rhodopsin.  相似文献   

3.
On stimulation, rhodopsin, the light-sensing protein in the rod cells of the retina, is phosphorylated at several sites on its C terminus as the first step in deactivation. We have developed a mass spectrometry-based method to quantify the kinetics of phosphorylation at each site in vivo. After exposing either a freshly dissected mouse retina or the eye of an anesthetized mouse to a flash of light, phosphorylation and dephosphorylation reactions are terminated by rapidly homogenizing the retina or enucleated eye in 8 M urea. The C-terminal peptide containing all known phosphorylation sites is cleaved from rhodopsin, partially purified by ultracentrifugation, and analyzed by liquid chromatography coupled with mass spectrometry (LCMS). The mass spectrometer responds linearly to the peptide from 10 fmole to 100 pmole. The relative sensitivity to peptides with zero to five phosphates was determined using purified phosphopeptide standards. High pressure liquid chromatography (HPLC) coupled with tandem mass spectrometry (LCMS/MS) was used to distinguish the three primary sites of phosphorylation, Ser 334, Ser 338, and Ser 343. Peptides monophosphorylated on Ser 334 were separable from those monophosphorylated on Ser 338 and Ser 343 by reversed-phase HPLC. Although peptides monophosphorylated at Ser 338 and Ser 343 normally coelute, the relative amounts of each species in the single peak could be determined by monitoring the ratio of specific daughter ions characteristic of each peptide. Doubly phosphorylated rhodopsin peptides with different sites of phosphorylation also were distinguished by LCMS/MS. The sensitivity of these methods was evaluated by using them to measure rhodopsin phosphorylation stimulated either by light flashes or by continuous illumination over a range of intensities.  相似文献   

4.
Invertebrate rhodopsins activate a G-protein signalling pathway in microvillar photoreceptors. In contrast to the transducin-cyclic GMP phosphodiesterase pathway found in vertebrate rods and cones, visual transduction in cephalopod (squid, octopus, cuttlefish) invertebrates is signalled via Gq and phospholipase C. Squid rhodopsin contains the conserved residues of the G-protein coupled receptor (GPCR) family, but has only 35% identity with mammalian rhodopsins. Unlike vertebrate rhodopsins, cephalopod rhodopsin is arranged in an ordered lattice in the photoreceptor membranes. This organization confers sensitivity to the plane of polarized light and also provides the optimal orientation of the linear retinal chromophores in the cylindrical microvillar membranes for light capture. Two-dimensional crystals of squid rhodopsin show a rectilinear arrangement that is likely to be related to the alignment of rhodopsins in vivo.Here, we present a three-dimensional structure of squid rhodopsin determined by cryo-electron microscopy of two-dimensional crystals. Docking the atomic structure of bovine rhodopsin into the squid density map shows that the helix packing and extracellular plug structure are conserved. In addition, there are two novel structural features revealed by our map. The linear lattice contact appears to be made by the transverse C-terminal helix lying on the cytoplasmic surface of the membrane. Also at the cytoplasmic surface, additional density may correspond to a helix 5-6 loop insertion found in most GPCRs relative to vertebrate rhodopsins. The similarity supports the conservation in structure of rhodopsins (and other G-protein-coupled receptors) from phylogenetically distant organisms. The map provides the first indication of the structural basis for rhodopsin alignment in the microvillar membrane.  相似文献   

5.
Light-induced rhodopsin signaling is turned off with sub-second kinetics by rhodopsin phosphorylation followed by arrestin-1 binding. To test the availability of the arrestin-1 pool in dark-adapted outer segment (OS) for rhodopsin shutoff, we measured photoresponse recovery rates of mice with arrestin-1 content in the OS of 2.5%, 5%, 60%, and 100% of wild type (WT) level by two-flash ERG with the first (desensitizing) flash at 160, 400, 1000, and 2500 photons/rod. The time of half recovery (t(half)) in WT retinas increases with the intensity of the initial flash, becoming ~2.5-fold longer upon activation of 2500 than after 160 rhodopsins/rod. Mice with 60% and even 5% of WT arrestin-1 level recovered at WT rates. In contrast, the mice with 2.5% of WT arrestin-1 had a dramatically slower recovery than the other three lines, with the t(half) increasing ~28 fold between 160 and 2500 rhodopsins/rod. Even after the dimmest flash, the rate of recovery of rods with 2.5% of normal arrestin-1 was two times slower than in other lines, indicating that arrestin-1 level in the OS between 100% and 5% of WT is sufficient for rapid recovery, whereas with lower arrestin-1 the rate of recovery dramatically decreases with increased light intensity. Thus, the OS has two distinct pools of arrestin-1: cytoplasmic and a separate pool comprising ~2.5% that is not immediately available for rhodopsin quenching. The observed delay suggests that this pool is localized at the periphery, so that its diffusion across the OS rate-limits the recovery. The line with very low arrestin-1 expression is the first where rhodopsin inactivation was made rate-limiting by arrestin manipulation.  相似文献   

6.
The deactivation of visual pigments involved in phototransduction is critical for recovering sensitivity after exposure to light in rods and cones of the vertebrate retina. In rods, phosphorylation of rhodopsin by rhodopsin kinase (GRK1) and the subsequent binding of visual arrestin completely terminates phototransduction. Although signal termination in cones is predicted to occur via a similar mechanism as in rods, there may be differences due to the expression of related but distinct gene products. While rods only express GRK1, cones in some species express only GRK1 or GRK7 and others express both GRKs. In the mouse, cone opsin is phosphorylated by GRK1, but this has not been demonstrated in mammals that express GRK7 in cones. We compared cone opsin phosphorylation in intact retinas from the 13-lined ground squirrel (GS) and pig, cone- and rod-dominant mammals, respectively, which both express GRK7. M opsin phosphorylation increased during continuous exposure to light, then declined between 3 and 6 min. In contrast, rhodopsin phosphorylation continued to increase during this time period. In GS retina homogenates, anti-GS GRK7 antibody blocked M opsin phosphorylation by 73%. In pig retina homogenates, only 20% inhibition was observed, possibly due to phosphorylation by GRK1 released from rods during homogenization. Our results suggest that GRK7 phosphorylates M opsin in both of these mammals. Using an in vitro GTPgammaS binding assay, we also found that the ability of recombinant M opsin to activate G(t) was greatly reduced by phosphorylation. Therefore, phosphorylation may participate directly in the termination of phototransduction in cones by decreasing the activity of M opsin.  相似文献   

7.
Rhodopsins are densely packed in rod outer-segment membranes to maximize photon absorption, but this arrangement interferes with transducin activation by restricting the mobility of both proteins. We attempted to explore this phenomenon in transgenic mice that overexpressed rhodopsin in their rods. Photon capture was improved, and, for a given number of photoisomerizations, bright-flash responses rose more gradually with a reduction in amplification—but not because rhodopsins were more tightly packed in the membrane. Instead, rods increased their outer-segment diameters, accommodating the extra rhodopsins without changing the rhodopsin packing density. Because the expression of other phototransduction proteins did not increase, transducin and its effector phosphodiesterase were distributed over a larger surface area. That feature, as well as an increase in cytosolic volume, was responsible for delaying the onset of the photoresponse and for attenuating its amplification.  相似文献   

8.
The eye tissues and liver of the alligator contain vitamin A1 alone. The retina contains rhodopsin, typical in absorption spectrum (λmax 500 mµ); but synthesized in solution from neo-b retinene and opsin much more rapidly than are the frog, mammalian, or chicken rhodopsins previously examined. In this regard alligator rhodopsin resembles the rhodopsins and porphyropsins of fishes, all of which so far investigated are synthesized rapidly in solution. The rates of synthesis in vitro of frog and alligator rhodopsins are matched closely by the rates of rod dark adaptation in living frogs and alligators, measured electrophysiologically at the same temperature. Alligator rods dark-adapt, and alligator rhodopsin is synthesized in solution, at rates characteristically associated with cones and cone pigments in frogs, mammals, and birds.  相似文献   

9.
Eukaryotic microbial rhodopsins are widespread bacteriorhodopsin-like proteins found in many lower eukaryotic groups including fungi. Many fungi contain multiple rhodopsins, some significantly diverged from the original bacteriorhodopsin template. Although few fungal rhodopsins have been studied biophysically, both fast-cycling light-driven proton pumps and slow-cycling photosensors have been found. The purpose of this study was to characterize photochemically a new subgroup of fungal rhodopsins, the so-called auxiliary group. The study used the two known rhodopsin genes from the fungal wheat pathogen, Phaeosphaeria nodorum. One of the genes is a member of the auxiliary group while the other is highly similar to previously characterized proton-pumping Leptosphaeria rhodopsin. Auxiliary rhodopsin genes from a range of species form a distinct group with a unique primary structure and are located in carotenoid biosynthesis gene cluster. Amino acid conservation pattern suggests that auxiliary rhodopsins retain the transmembrane core of bacteriorhodopsins, including all residues important for proton transport, but have unique polar intramembrane residues. Spectroscopic characterization of the two yeast-expressed Phaeosphaeria rhodopsins showed many similarities: absorption spectra, conformation of the retinal chromophore, fast photocycling, and carboxylic acid protonation changes. It is likely that both Phaeosphaeria rhodopsins are proton-pumping, at least in vitro. We suggest that auxiliary rhodopsins have separated from their ancestors fairly recently and have acquired the ability to interact with as yet unidentified transducers, performing a photosensory function without changing their spectral properties and basic photochemistry.  相似文献   

10.
Vertebrate retinas have two types of photoreceptor cells, rods and cones, which contain visual pigments with different molecular properties. These pigments diverged from a common ancestor, and their difference in molecular properties originates from the difference in their amino acid residues. We previously reported that the difference in decay times of G protein-activating meta-II intermediates between the chicken rhodopsin and green-sensitive cone (chicken green) pigments is about 50 times. This difference only originates from the differences of two residues at positions 122 and 189 (Kuwayama, S., Imai, H., Hirano, T., Terakita, A., and Shichida, Y. (2002) Biochemistry 41, 15245-15252). Here we show that the meta-III intermediates exhibit about 700 times difference in decay times between the two pigments, and the faster decay in chicken green can be converted to the slower decay in rhodopsin by replacing the residues in chicken green with the corresponding rhodopsin residues. However, the inverse directional conversion did not occur when the two residues in rhodopsin were replaced by those of chicken green. Analysis using chimerical mutants derived from these pigments has demonstrated that amino acid residues responsible for the slow rhodopsin meta-III decay are situated at several positions throughout the C-terminal half of rhodopsin. Considering that rhodopsins evolved from cone pigments, it has been suggested that the molecular properties of rhodopsin have been optimized by mutations at several positions, and the chicken green mutants at two positions could be rhodopsin-like pigments transiently produced in the course of molecular evolution.  相似文献   

11.
The phosphorylation of a synthetic peptide, corresponding to the C-terminal 11 amino acids of bovine rhodopsin (VII, residues 338-348), was studied under different conditions. The peptide was only phosphorylated in the presence of photoactivated rhodopsin. Using the same protocol, 12 other peptides, mapping in the rhodopsin C-terminal, were screened for their effectiveness as substrates for rhodopsin kinase. It was found that the peptides became poorer substrates with increasing length, and the best substrates comprised the most C-terminal 9-12 amino acids as opposed to other parts of the C-terminus. It was noted that the absence of the two-terminal residues Pro347 and Ala348 impaired peptide phosphorylation. The effect of the decay of metarhodopsin II on the phosphorylation of rhodopsin and the peptides was determined, and it was found that the rhodopsin and peptide phosphorylations decayed with half times of approximately 33 min and 28 min, respectively. The sites of phosphorylation on the peptides were determined and in all cases the phosphorylation was found to be predominantly on serine residues. Only the 11-residue peptide (VII, residues 338-348) contained significant threonine phosphorylation, which was about 25% that on serine residues. Cumulatively, the results suggest that Ser343 is the preferred site of phosphorylation in vitro. The reason for the poor substrate effectiveness of the larger peptides was examined by competitive experiments in which it was shown that a poorly phosphorylated larger peptide successfully inhibited the phosphorylation of a 'good' peptide substrate. The studies above support a mechanism for rhodopsin kinase that we have termed the 'kinase-activation hypothesis'. This requires that the kinase exists in an inactive form and is activated only after binding to photoactivated rhodopsin.  相似文献   

12.
Following the discovery of the bacteriorhodopsin proton pump in Halobacterium halobium (salinarum), not only the halorhodopsin halide pump and two photosensor rhodopsins (sensory rhodopsin and phoborhodopsin) in the same species, but also homologs of these four rhodopsins in strains of other genera of Halobacteriaceae have been reported. Twenty-eight full (and partial) sequences of the genomic DNA of these rhodopsins have been analyzed. The deduced amino acid sequences have led to new strategies and tactics for understanding bacterial rhodopsins on a comparative basis, as summarized briefly in this article. The data discussed include (i) alignment of the sequences to qualify/characterize the conserved residues; (ii) assignment of residues that cause differences in function(s)/properties; and (iii) phylogeny of the halobacterial rhodopsins to suggest their evolutionary paths. The four kinds of rhodopsin in each strain are assumed, on the basis of their genera-specific distributions, to have arisen by at least two gene-duplication processes during evolution prior to generic speciation. The first duplication of the rhodopsin ancestor gene yielded two genes, each of which was duplicated again to give four genes in the ancestor halobacterium. The bacterium carrying four rhodopsin genes, after accumulating mutations, became ready for generic speciation and the delivery of four rhodopsins to each species. The original rhodopsin ancestor is speculated to be closest to the proton pump (bacteriorhodopsin).  相似文献   

13.
The single photon responses of retinal rod cells are remarkably reproducible, allowing the number and timing of photon absorptions to be encoded accurately. This reproducibility is surprising because the elementary response arises from a single rhodopsin molecule, and typically signals from single molecules display large intertrial variations. We have investigated the mechanisms that make the rod's elementary response reproducible. Our experiments indicate that reproducibility cannot be explained by saturation within the transduction cascade, by Ca2+ feedback, or by feedback control of rhodopsin shutoff by any known element of the cascade. We suggest instead that deactivation through a series of previously unidentified transitions allows the catalytic activity of a single rhodopsin molecule to decay with low variability. Two observations are consistent with this view. First, the time course of rhodopsin's catalytic activity could not be accounted for by the time required for the known steps in rhodopsin deactivation-phosphorylation and arrestin binding. Second, the variability of the elementary response increased when phosphorylation was made rate-limiting for rhodopsin shutoff.  相似文献   

14.
Signal transduction in rod cells begins with photon absorption by rhodopsin and leads to the generation of an electrical response. The response profile is determined by the molecular properties of the phototransduction components. To examine how the molecular properties of rhodopsin correlate with the rod-response profile, we have generated a knock-in mouse with rhodopsin replaced by its E122Q mutant, which exhibits properties different from those of wild-type (WT) rhodopsin. Knock-in mouse rods with E122Q rhodopsin exhibited a photosensitivity about 70% of WT. Correspondingly, their single-photon response had an amplitude about 80% of WT, and a rate of decline from peak about 1.3 times of WT. The overall 30% lower photosensitivity of mutant rods can be explained by a lower pigment photosensitivity (0.9) and the smaller single-photon response (0.8). The slower decline of the response, however, did not correlate with the 10-fold shorter lifetime of the meta-II state of E122Q rhodopsin. This shorter lifetime became evident in the recovery phase of rod cells only when arrestin was absent. Simulation analysis of the photoresponse profile indicated that the slower decline and the smaller amplitude of the single-photon response can both be explained by the shift in the meta-I/meta-II equilibrium of E122Q rhodopsin toward meta-I. The difference in meta-III lifetime between WT and E122Q mutant became obvious in the recovery phase of the dark current after moderate photobleaching of rod cells. Thus, the present study clearly reveals how the molecular properties of rhodopsin affect the amplitude, shape, and kinetics of the rod response.  相似文献   

15.
H Chong  J Lee  K L Guan 《The EMBO journal》2001,20(14):3716-3727
Activating and inhibitory phosphorylation mechanisms play an essential role in regulating Raf kinase activity. Here we demonstrate that phosphorylation of C-Raf in the kinase activation loop (residues T491 and S494) is necessary, but not sufficient, for activation. C-Raf has additional activating phosphorylation sites at S338 and Y341. Mutating all four of these residues to acidic residues, S338D/Y341D/T491E/S494D (DDED), in C-Raf results in constitutive activity. However, acidic residue substitutions at the corresponding activation loop sites in B-Raf are sufficient to confer constitutive activity. B-Raf and C-Raf also utilize similar inhibitory phosphorylation mechanisms to regulate kinase activity. B-Raf has multiple inhibitory phosphorylation sites necessary for full kinase inhibition where C-Raf requires only one. We examined the functional significance of these inhibitory and activating phosphorylations in Caenorhabditis elegans lin-45 Raf. Eliminating the inhibitory phosphorylation or mimicking activating phosphorylation sites is sufficient to confer constitutive activity upon lin-45 Raf and induce multi-vulva phenotypes in C.elegans. Our results demonstrate that different members of the Raf family kinases have both common and distinct phosphorylation mechanisms to regulate kinase activity and biological function.  相似文献   

16.
Visual opsins bind 11-cis retinal at an orthosteric site to form rhodopsins but increasing evidence suggests that at least some are capable of binding an additional retinoid(s) at a separate, allosteric site(s). Microspectrophotometric measurements on isolated, dark-adapted, salamander photoreceptors indicated that the truncated retinal analog, β-ionone, partitioned into the membranes of green-sensitive rods; however, in blue-sensitive rod outer segments, there was an enhanced uptake of four or more β-ionones per rhodopsin. X-ray crystallography revealed binding of one β-ionone to bovine green-sensitive rod rhodopsin. Cocrystallization only succeeded with extremely high concentrations of β-ionone and binding did not alter the structure of rhodopsin from the inactive state. Salamander green-sensitive rod rhodopsin is also expected to bind β-ionone at sufficiently high concentrations because the binding site is present on its surface. Therefore, both blue- and green-sensitive rod rhodopsins have at least one allosteric binding site for retinoid, but β-ionone binds to the latter type of rhodopsin with low affinity and low efficacy.  相似文献   

17.
Characterization of Drosophila melanogaster rhodopsin   总被引:1,自引:0,他引:1  
A polypeptide present in Drosophila eye homogenates was identified as opsin. This polypeptide pI 7.8, with Mr 39,000 is a retina-specific protein. It has the spectral characteristics of rhodopsin contained in the R1-6 photoreceptors and decreases in amount with vitamin A deprivation. It contains a chromophore derived from vitamin A and linked to the protein moiety by a Schiff base. Moreover, the polypeptide identified corresponds to a retina-specific polypeptide that was shown previously to undergo light-dependent phosphorylation in living flies. These results indicate that many properties of Drosophila rhodopsin do not differ significantly from those reported for rhodopsins of other organisms. However, the isoelectric point of Drosophila opsin is considerably more basic than those reported for vertebrate rhodopsins.  相似文献   

18.
A multi-technique approach to identification and mapping of phosphorylation on protein kinase A (PKA) is described. X-ray crystallography revealed phosphorylation at T197 and S338 while mass spectrometry (MS) on the intact protein suggested phosphorylation at three sites. Tryptic digestion, followed by MS, confirmed the presence of three phosphates. However, metal affinity treatment of the digest prior to MS revealed the presence of a fourth phosphopeptide. Subsequent analysis of the digests using liquid chromatography (LC) coupled with quadrupole ion trap (QIT) MS confirmed phosphorylation at S10 and S338 and suggested phosphorylation at S139 and T195/197. Unfortunately, identification of pS139 was inconclusive due to low signal intensity and early elution in reversed-phase LC while poor MS/MS data prevented localization of the phosphate to T195 or T197. Phosphopeptide modification with ethanethiol, followed by LC QIT-MS/MS, identified four phosphopeptides in a single experiment. In addition, the fragmentation data provided significantly more sequence information than data obtained from unmodified peptides. Data from this study suggested that PKA was completely phosphorylated at S10, T197, and S338 and partially phosphorylated at S139. These results illustrate that critical information can be lost unless multiple MS techniques are used for identification and validation of phosphorylation.  相似文献   

19.
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   

20.
A rhodopsin phosphorylation reaction that occurs with high-gain is observed if measurements are made in electropermeabilized frog rod outer segments (ROS) stimulated by a dim flash of light in the operating range of the photoreceptor. Flashes of light exciting 1000 or fewer of the 3 x 10(9) rhodopsins present/ROS results in the incorporation of 1400 phosphates from ATP into the rhodopsin pool for each excited rhodopsin (Rho*). This amplification decreases with increasing light intensity, falling most sharply after each disk has absorbed one photon. The high-gain reaction is lost if the ROS are broken into vesicles by shearing, leaving a low-gain rhodopsin phosphorylation characterized in previous studies using brighter illumination. The high-gain but not the low-gain phosphorylation appears to be regulated by G-protein and by calcium levels in the range over which intracellular calcium changes when rod photoreceptors are illuminated. Kinetic measurements made on the phosphorylation observed at higher light intensities shows that it initially occurs rapidly enough for a role in terminating the photoresponse. The high-gain phosphorylation observed at lower light intensities may play a global role in regulating light-adaptation of the rod photoreceptor, and its existence suggests that a search for a similar high-gain modification in systems using the homologous beta-adrenergic or muscarinic acetylcholine receptors might be rewarding.  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