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.
Photochemical studies were conducted on human rhodopsin at 20 degrees C to characterize the intermediates which precede the formation of metarhodopsin II, the trigger for the enzyme cascade mechanism of visual transduction. Human rhodopsin was prepared from eyes which had previously been used for corneal donations. Time resolved absorption spectra collected from 10(-8) to 10(-6) s after photolysis of human rhodopsin in detergent suspensions displayed biexponential decay kinetics. The apparent lifetimes obtained from the data are 65 +/- 20 and 292 +/- 25 ns, almost a factor of 2 slower than the corresponding rates in bovine rhodopsin. The spectra can be fit well using a model in which human bathorhodopsin decays toward equilibrium with a blue-shifted intermediate (BSI) which then decays to lumirhodopsin. Spectra and kinetic rate constants were determined for all these intermediates using a global analysis which showed that the spectra of the human intermediates are remarkably similar to bovine intermediates. Microscopic rate constants derived from this model are 7.4 x 10(6) s-1 for bathorhodopsin decay and 7.5 x 10(6) s-1 and 4.6 x 10(6) s-1 for the forward and reverse reactions of BSI, respectively. Decay of lumirhodopsin to later intermediates was studied from 10(-6) to 10(-1) s after photolysis of rhodopsin in human disk membrane suspensions. The human metarhodopsin I in equilibrium metarhodopsin II equilibrium appears to be more forward shifted than in comparable bovine studies.  相似文献   

2.
Nonbleachable rhodopsins containing retinal moieties with fixed 11-ene structures have been prepared. When the nonbleachable rhodopsin analogue corresponding to the natural pigment was flash-photolysed at 20.8 degrees C, no absorption changes occurred at the monitoring wavelengths of 380, 480, and 580 nm for the time range of 2 microseconds--10 s. This observation is in contrast to that of natural rhodopsin which showed the formation of metarhodopsin I and its decay to meta II. Irradiation of the artificial rhodopsin, 77 K, with light of 460 and 540 nm, also gave no spectral changes; in the case of natural rhodopsin, however, the irradiation leads to formation of the red-shifted intermediate bathorhodopsin. The absence of photochemistry in the artificial pigment shows that an 11-cis to trans photoisomerization of the retinal moiety is a crucial step in inducing the chain of events in te photolysis of rhodopsin.  相似文献   

3.
Glutamic acid at position 113 in bovine rhodopsin ionizes to form the counterion to the protonated Schiff base (PSB), which links the 11-cis-retinylidene chromophore to opsin. Photoactivation of rhodopsin requires both Schiff base deprotonation and neutralization of Glu-113. To better understand the role of electrostatic interactions in receptor photoactivation, absorbance difference spectra were collected at time delays from 30 ns to 690 ms after photolysis of rhodopsin mutant E113Q solubilized in dodecyl maltoside at different pH values at 20 degrees C. The PSB form (pH 5. 5, lambda(max) = 496 nm) and the unprotonated Schiff base form (pH 8. 2, lambda(max) = 384 nm) of E113Q rhodopsin were excited using 477 nm or 355 nm light, respectively. Early photointermediates of both forms of E113Q were qualitatively similar to those of wild-type rhodopsin. In particular, early photoproducts with spectral shifts to longer wavelengths analogous to wild-type bathorhodopsin were seen. In the case of the basic form of E113Q, the absorption maximum of this intermediate was at 408 nm. These results suggest that steric interaction between the retinylidene chromophore and opsin, rather than charge separation, plays the dominant role in energy storage in bathorhodopsin. After lumirhodopsin, instead of deprotonating to form metarhodopsin I(380) on the submillisecond time scale as is the case for wild type, the acidic form of E113Q produced metarhodopsin I(480), which decayed very slowly (exponential lifetime = 12 ms). These results show that Glu-113 must be present for efficient deprotonation of the Schiff base and rapid visual transduction in vertebrate visual pigments.  相似文献   

4.
The photochemical intermediate metarhodopsin II (meta II; lambda max = 380 nm) is generally identified with rho*, the conformation of photolyzed rhodopsin which binds and activates the visual G-protein, Gt [Emeis, D., & Hoffman, K.P. (1981) FEBS Lett. 136, 201-207]. Purified bovine rhodopsin was incorporated into vesicles consisting of dimyristoylphosphatidylcholine (DMPC), and the rapid formation of a photochemical intermediate absorbing maximally at 380 nm was quantified via both flash photolysis and equilibrium spectral measurements. Kinetic and equilibrium spectral measurements performed above the Tm of DMPC showed that Gt, in the absence of GTP, enhances the production of the 380-nm-absorbing species while reducing the concentration of the 478-nm-absorbing species, metarhodopsin I (meta I), in a manner similar to that observed in the native rod outer segment disk membrane. This Gt-induced shift in the equilibrium concentration of photointermediates indicated that the species with an absorbance maximum at 380 nm was meta II. The presence of rho* in the DMPC bilayer was established via measurements of photolysis-induced exchange of tritiated GMPPNP, a nonhydrolyzable analogue of GTP, on Gt. Above Tm, the metarhodopsin equilibrium is strongly shifted toward meta I relative to the native rod outer segment disk membrane; however, at 37 degrees C, 40% of the photointermediates are in the form of meta II. The formation of meta II above Tm is slowed by a factor of ca. 2 relative to the disk membrane. Below Tm, the equilibrium is shifted still further toward meta I, and meta II forms ca. 7 times slower than in the disk membrane.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

5.
The ionization changes during the photolysis of the visual pigment, cattle rhodopsin, have been measured by simultaneous recording of spectral and pH changes. The thermal intermediates of rhodopsin and pH changes were recorded over a pH range of 4.6–8.9.In the normal sequence of intermediate changes at pH values of 5.4–7.7, the proton uptake of rhodopsin during the metarhodopsin I478 to II380 reaction is followed by a proton release in the thermal decay of metarhodopsin II380 to III465. Below pH 5.4, no proton release is observed during the thermal decay of metarhodopsin II380, and the metarhodopsin II380 appears to thermally decay directly to N-retinylidene-opsin440. Above pH 7.7, the major process appears to be a proton release and the final product is N-retinylidene-opsin365.The ionization state of certain groups in rhodopsin appears to control the metarhodopsin I478 to II380 reaction and control the products in the thermal decay of metarhodopsin II380. The pK changes of certain groups in rhodopsin may be the major factor in determining sequence of thermal intermediates and the values of the kinetic activation parameters. The reversing ionization changes may be important to the transduction process.  相似文献   

6.
7.
The torsion model with which we proposed to interpret the specific properties of the photoisomerization reaction of rhodopsin has been developed to apply to isorhodopsin I, isorhodopsin II and some intermediates. Based on this model, optical absorption wavelengths and oscillator strengths, as well as rotational strengths of visual pigments, analogues and intermediates at low temperatures are analyzed by varying twisted conformations of the chromophores. As a result, it was found that most of the optical data could be very well accounted for quantitatively by the torsion model. The twisting characters in the chromophore of rhodopsin are very similar to those of isorhodopsin. The obtained conformations of the chromophores are very similar in rhodopsin and its analogues, and in isorhodopsin and its analogues. Those of the chromophores of bathorhodopsin, lumirhodopsin and metarhodopsin I are similar to one another except that the conjugated chain of metarhodopsin I bends considerably when compared with the other intermediates.  相似文献   

8.
G Renk  R K Crouch 《Biochemistry》1989,28(2):907-912
Several analogue pigments have been prepared containing retinals altered at the cyclohexyl ring or proximal to the aldehyde group in order to examine the role of the chromophore in the formation of the metarhodopsin I and II states of visual pigments. Deletion of the 13-methyl group on the isoprenoid chain did not affect metarhodopsin formation. However, analogue pigments containing chromophores with modified rings did not show the typical absorption changes associated with the metarhodopsin transitions of native or regenerated rhodopsins. In particular, 4-hydroxyretinal pigments did not show clear transitions between the metarhodopsin I and metarhodopsin II states. Pigment formed with an acyclic retinal showed no evidence by absorption spectroscopy of metarhodopsin formation. A retinal altered by substitution of a five-membered ring containing a nitroxide required a more acidic pH than the native pigment for formation of the metarhodopsin II state. ESR data suggest that the ring remains buried within the protein through the metarhodopsin II state. However, the Schiff base linkage is susceptible to hydrolysis of hydroxylamine in the metarhodopsin II state. These data indicate that (1), in the transition from rhodopsin to metarhodopsin II, major protein conformational changes are occurring near the lysine-retinal linkage whereas the ring portion of the chromophore remains deeply buried within the protein and (2) pigment absorptions characteristic of the metarhodopsin I and II states may be due to specific protein-chromophore interactions near the region of the chromophore ring.  相似文献   

9.
Lewis JW  Szundi I  Kazmi MA  Sakmar TP  Kliger DS 《Biochemistry》2006,45(17):5430-5439
The role of ionizable amino acid side chains in the bovine rhodopsin activation mechanism was studied in mutants E134Q, E134R/R135E, H211F, and E122Q. All mutants exhibited bathorhodopsin stability on the 30 ns to 1 micros time scale similar to that of the wild type. Lumirhodopsin decay was also similar to that of the wild type except for the H211F mutant where early decay (20 micros) to a second form of lumirhodopsin was seen, followed by formation of an extremely long-lived Meta I(480) product (34 ms), an intermediate which forms to a much reduced extent, if at all, in dodecyl maltoside suspensions of wild-type rhodopsin. A smaller amount of a similar long-lived Meta I(480) product was seen after photolysis of E122Q, but E134Q and E134R/R135Q displayed kinetics much more similar to those of the wild type under these conditions (i.e., no Meta I(480) product). These results support the idea that specific interaction of His211 and Glu122 plays a significant role in deprotonation of the retinylidene Schiff base and receptor activation. Proton uptake measurements using bromcresol purple showed that E122Q was qualitatively similar to wild-type rhodopsin, with at least one proton being released during lumirhodopsin decay per Meta I(380) intermediate formed, followed by uptake of at least two protons per rhodopsin bleached on a time scale of tens of milliseconds. Different results were obtained for H211F, E134Q, and E134R/R135E, which all released approximately two protons per rhodopsin bleached. These results show that several ionizable groups besides the Schiff base imine are affected by the structural changes involved in rhodopsin activation. At least two proton uptake groups and probably at least one proton release group in addition to the Schiff base are present in rhodopsin.  相似文献   

10.
Enthalpy changes (Delta H) of the photointermediates that appear in the photolysis of octopus rhodopsin were measured at physiological temperatures by the laser-induced transient grating method. The enthalpy from the initial state, rhodopsin, to bathorhodopsin, lumirhodopsin, mesorhodopsin, transient acid metarhodopsin, and acid metarhodopsin were 146 +/- 15 kJ/mol, 122 +/- 17 kJ/mol, 38 +/- 8 kJ/mol, 12 +/- 5 kJ/mol, and 12 +/- 5 kJ/mol, respectively. These values, except for lumirhodopsin, are similar to those obtained for the cryogenically trapped intermediate species by direct calorimetric measurements. However, the Delta H of lumirhodopsin at physiological temperatures is quite different from that at low temperature. The reaction volume changes of these processes were determined by the pulsed laser-induced photoacoustic method along with the above Delta H values. Initially, in the transformation between rhodopsin and bathorhodopsin, a large volume expansion of +32 +/- 3 ml/mol was obtained. The volume changes of the subsequent reaction steps were rather small. These results are compared with the structural changes of the chromophore, peptide backbone, and water molecules within the membrane helixes reported previously.  相似文献   

11.
P A Baldwin  W L Hubbell 《Biochemistry》1985,24(11):2633-2639
When rhodopsin is incorporated into the saturated short-chain phospholipid dimyristoylphosphatidylcholine, photolysis of the protein results in an abnormal sequence of spectral transitions, and the dominant product of metarhodopsin I decay is free retinal plus opsin [Baldwin, P. A., & Hubbell, W. L. (1985) Biochemistry (preceding paper in this issue)]. By incorporation of rhodopsin into a series of phosphatidylcholines of defined composition, we have determined the properties of the lipid environment that are responsible for the altered spectral behavior. Metarhodopsin II is not found in appreciable amounts in bilayers containing acyl chains that are too short (14 or fewer carbon atoms in length), in the presence of only n-alkyl chains, or below the characteristic phase-transition temperature of recombinant membranes. Double bonds are not required for the formation of the metarhodopsin II intermediate, as it is observed in diphytanoylphosphatidylcholine recombinants.  相似文献   

12.
We have previously described [H, Kühn et al. (1981) Proc. Natl Acad. Sci. USA, 78, 6873-6877] a light-induced scattering change ('binding signal') associated with a stoichiometric binding between photoexcited rhodopsin and a peripheral membrane protein, the GTP-binding protein, in bovine rod outer segment suspensions. We have attempted here to identify the rhodopsin intermediate R* which is responsible for this interaction, by studying its dependence on pH, temperature and ionic strength. The results strongly suggest that the active state is metarhodopsin II (M II). 1. The initial phase of the binding signal is slightly slower than the formation of metarhodopsin II (2-37 degrees C, pH 5.5-9). 2. The kinetics of the decay of the active rhodopsin state are similar to those of the metarhodopsin II leads to metarhodopsin III transition (37 degrees C, pH 7.3). 3. All conditions which lead to light-induced binding of the GTP-binding protein to R* also lead to the formation of M II. At 2 degrees C, pH 8.3, in particular where no M II is formed in the absence of GTP-binding protein, binding signals and light-induced attachment of the GTP-binding protein to the membrane are still observed. Consistently, addition of GTP-binding protein to a suspension of extracted membranes bleached at 2 degrees C (pH 8.3) shifts the metarhodopsin I in equilibrium metarhodopsin II equilibrium towards metarhodopsin II. The shift is reversed by GTP, which dissociates the rhodopsin--GTP-binding protein complex. 4. At low ionic strength, where the GTP-binding protein is soluble in the dark (instead of being associated to the membrane as in the above experiments) M II still induces the binding whereas M I does not, indicating a much lower affinity of the GTP-binding protein for MI.  相似文献   

13.
Resonance Raman spectroscopy of octopus rhodopsin and its photoproducts   总被引:2,自引:0,他引:2  
C Pande  A Pande  K T Yue  R Callender  T G Ebrey  M Tsuda 《Biochemistry》1987,26(16):4941-4947
We report here the resonance Raman spectra of octopus rhodopsin and its photoproducts, bathorhodopsin and acid metarhodopsin. These studies were undertaken in order to make comparisons with the well-studied bovine pigments, so as to understand the similarities and the differences in pigment structure and photochemical processes between vertebrates and invertebrates. The flow method was used to obtain the Raman spectrum of rhodopsin at 13 degrees C. The bathorhodopsin spectrum was obtained by computer subtraction of the spectra containing different photostationary mixtures of rhodopsin, isorhodopsin, hypsorhodopsin, and bathorhodopsin, obtained at 12 K using the pump-probe technique and from measurements at 80 K. Like their bovine counterparts, the Schiff base vibrational mode appears at approximately 1660 cm-1 in octopus rhodopsin and the photoproducts, bathorhodopsin and acid metarhodopsin, suggesting a protonated Schiff base linkage between the chromophore and the protein. Differences between the Raman spectra of octopus rhodopsin and bathorhodopsin indicate that the formation of bathorhodopsin is associated with chromophore isomerization. This inference is substantiated by the chromophore chemical extraction data which show that, like the bovine system, octopus rhodopsin is an 11-cis pigment, while the photoproducts contain an all-trans pigment, in agreement with previous work. The octopus rhodopsin and bathorhodopsin spectra show marked differences from their bovine counterparts in other respects, however. The differences are most dramatic in the structure-sensitive fingerprint and the HOOP regions. Thus, it appears that although the two species differ in the specific nature of the chromophore-protein interactions, the general process of visual transduction is the same.  相似文献   

14.
The torsion model with which we proposed to interpret the specific properties of the photoisomerization reaction of rhodopsin has been developed to apply to isorhodopsin I, isorhodopsin II and some intermediates. Based on this model, optical absorption wavelengths and oscillator strengths, as well as rotational strengths of visual pigments, analogues and intermediates at low temperatures are analyzed by varying twisted conformations of the chromophores. As a result, it was found that most of the optical data could be very well accounted for quantitatively by the torsion model. The twisting characters in the chromophore of rhodopsin are very similar to those of isorhodopsin. The obtained conformations of the chromophores are very similar in rhodopsin and its analogues, and in isorhodopsin and its analogues. Those of the chromophores of bathorhodopsin, lumirhodopsin and metarhodopsin I are similar to one another except that the conjugated chain of metarhodopsin I bends considerably when compared with the other intermediates.A part of this work was performed while one of the authors (T.K.) was a Visiting Investigator of Japan Society for the Promotion of Science at Kyoto University from April, 1977 to March, 1978  相似文献   

15.
Flash photolysis studies have shown that the membrane lipid environment strongly influences the ability of rhodopsin to form the key metarhodopsin II intermediate. Here we have used plasmon-waveguide resonance (PWR) spectroscopy, an optical method sensitive to both mass and conformation, to probe the effects of lipid composition on conformational changes of rhodopsin induced by light and due to binding and activation of transducin (G(t)). Octylglucoside-solubilized rhodopsin was incorporated by detergent dilution into solid-supported bilayers composed either of egg phosphatidylcholine or various mixtures of a nonlamellar-forming lipid (dioleoylphosphatidylethanolamine; DOPE) together with a lamellar-forming lipid (dioleoylphosphatidylcholine; DOPC). Light-induced proteolipid conformational changes as a function of pH correlated well with previous flash photolysis studies, indicating that the PWR spectral shifts monitored metarhodopsin II formation. The magnitude of these effects, and hence the extent of the conformational transition, was found to be proportional to the DOPE content. Our data are consistent with previous suggestions that lipids having a negative spontaneous curvature favor elongation of rhodopsin during the activation process. In addition, measurements of the G(t)/rhodopsin interaction in a DOPC/DOPE (25:75) bilayer at pH 5 demonstrated that light activation increased the affinity for G(t) from 64 nM to 0.7 nM, whereas G(t) affinity for dark-adapted rhodopsin was unchanged. By contrast, in DOPC bilayers the affinity of G(t) for light-activated rhodopsin was only 18 nM at pH 5. Moreover exchange of GDP for GTP gamma S was also monitored by PWR spectroscopy. Only the light-activated receptor was able to induce this exchange which was unaffected by DOPE incorporation. These findings demonstrate that nonbilayer-forming lipids can alter functionally linked conformational changes of G-protein-coupled receptors in membranes, as well as their interactions with downstream effector proteins.  相似文献   

16.
Early photolysis intermediates of native bovine rhodopsin (RHO) are investigated by nanosecond laser photolysis near physiological temperature. Absorption difference spectra are collected after excitation with 477-, 532-, and 560-nm laser pulses of various energies and with 477-nm laser excitation at 5, 12, 17, 21, and 32 degrees C. The data are analyzed by using singular-value decomposition (SVD) and a global exponential fitting routine. Two rate constants associated with distinct spectral changes are observed during the time normally associated with the decay of bathorhodopsin to lumirhodopsin. Various models consistent with this observation are considered. A sequential model in which there is a reversible step between a bathorhodopsin intermediate and a new intermediate (BSI), which is blue-shifted relative to lumirhodopsin, is shown to best fit the data. The temperature dependence of the observed and calculated rate constants leads to linear Arrhenius plots. Extrapolation of the temperature dependence suggests that BSI should not be observable after rhodopsin photolysis at temperatures below -100 degrees C. The results are discussed with regard to the artificial visual pigments cis-5,6-dihydroisorhodopsin and 13-demethylrhodopsin. It is proposed that the rate of the BATHO to BSI transition is limited by the relaxation of the strained all-trans-retinal chromophore within a tight protein environment. The transition to LUMI involves chromophore relaxation concurrent with protein relaxation. While the first process is strongly affected by changes in the chromophore, the second transition seems to be determined more by protein relaxation.  相似文献   

17.
Activation of guanosine 3',5'-cyclic monophosphate (cGMP) phosphodiesterase (EC 3.1.4.35.) in frog rod outer segment membrane by rhodopsin and its analogues was investigated. The Schiff-base linkage between opsin and retinal in rhodopsin was not always necessary for the phosphodiesterase activation. The binding of beta-ionone ring of retinal to a hydrophobic region of opsin was not enough to induce the enzyme activation. A striking photo-activation of the enzyme was induced by photo-isomerization of rhodopsin analogues from cis to trans form. It seems probable that an "expanded" conformation of opsin around the retinylidene chromophore induced by the cis to trans isomerization may be the trigger for the activation of phosphodiesterase. On the other hand, the phosphodiesterase in frog rod outer segment was activated by warming of bathorhodopsin to -12 degrees C and then incubating it at the same temperature. Thus, metarhodopsin II or an earlier intermediate than metarhodopsin II should be a direct intermediate for the enzyme activation.  相似文献   

18.
Lewis JW  Szundi I  Kazmi MA  Sakmar TP  Kliger DS 《Biochemistry》2004,43(39):12614-12621
The role of glutamic acid 181 in the bovine rhodopsin retinylidene chromophore pocket was studied by expressing E181 mutants in COS cells and measuring, as a function of time, the absorbance changes produced after excitation of lauryl maltoside pigment suspensions with 7 ns laser pulses. All mutants studied except E181D showed accelerated decay of bathorhodopsin compared to wild type. Even for E181D, an anomalously large blue shift was observed in the absorption spectrum of the bathorhodopsin decay product, BSI. These observations support the idea that E181 plays a significant role in the earliest stages of receptor activation. E181 mutations have a pronounced effect on the decay of the lumirhodopsin photointermediate, primarily affecting the size of the red shift that occurs in the lumirhodopsin I to lumirhodopsin II transition that takes place on the 10 micros time scale after wild-type photoexcitation. While the spectral change that occurs in the lumirhodopsin I to lumirhodopsin II transition in wild-type rhodopsin is very small ( approximately 2 nm), making it difficult to detect, it is larger in E181D ( approximately 6 nm), making it evident even in the lower signal-to-noise ratio measurements possible with rhodopsin mutants. The change seen is even larger for the E181F mutant where significant amounts of a deprotonated Schiff base intermediate are produced with the 10 micros time constant of lumirhodopsin II formation. The E181Q mutant shows lumirhodopsin decay more similar to wild-type behavior, and no lumirhodopsin I to lumirhodopsin II transition can be resolved. The addition of chloride ion to E181Q increases the lumirhodopsin I-lumirhodopsin II spectral shift and slows the deprotonation of the Schiff base. The latter result is consistent with the idea that a negative charge at position 181 contributes to protonated Schiff base stability in the later intermediates.  相似文献   

19.
P A Baldwin  W L Hubbell 《Biochemistry》1985,24(11):2624-2632
Photolysis of bovine rhodopsin in dimyristoylphosphatidylcholine recombinant membranes results in the production of a relatively stable metarhodopsin I like photointermediate that decays slowly to a species with a broad absorbance maximum centered at about 380 nm [O'Brien, D. F., Costa, L. F., & Ott, R. A. (1977) Biochemistry 16, 1295-1303]. On the basis of the results of a variety of chemical and spectroscopic tests, we show that this process corresponds to the production of free retinal plus opsin and not to the slow production of metarhodopsin II. Electron spin resonance studies using a novel disulfide spin-label that is covalently linked to rhodopsin indicate that the apparent arrest of the protein at the metarhodopsin I stage is not due to simple aggregation of the protein in this short-chain, saturated lipid bilayer but must be understood in terms of the effect of the lipid host on the conformational energies of individual protein molecules. Limited production of metarhodopsin II is observed under acidic conditions. Thus, the rhodopsin-dimyristoylphosphatidylcholine recombinants offer a unique system for the study of the effect of the phospholipid bilayer environment on the conformation of an intrinsic membrane protein.  相似文献   

20.
The decay reactions of metarhodopsin II and the dissociation of the complex between rhodopsin (in the metarhodopsin II state) and the GTP-binding protein (G-protein) (in its inactive, GDP-binding form) have been compared at various concentrations of hydroxylamine. The reactions of the chromophore were measured by absorption changes in the visible range, the complex dissociation by changes in the near-in-frared scattering. An additional monitor of the complex was given by the G-protein-dependent equilibrium between metarhodopsin I and metarhodopsin II. For all measurements, fragments of isolated bovine rod outer segments in suspension were used. In the absence of hydroxylamine, the rhodopsin-G-protein complex dissociated within 20–30 min at room temperature. The presence of hydroxylamine greatly accelerated (e.g., 5-fold at 1 mM NH2OH) the dissociation. Under all conditions, the free, dissociated G-protein can reassociate to metarhodopsin II produced by subsequent bleaching. Dissociation of the metarhodopsin II-G-protein complex required the decay of photoproducts with a maximal absorbance of 380 nm, but was not affected by the simultaneous presence of metarhodopsin III or metarhodopsin III — like photoproducts with a maximal absorbance between 450 and 470 nm. Despite the acceleration of metarhodopsin II-G-protein dissociation by NH2OH, metarhodopsin II-G-protein was relatively stabilized as compared to free metarhodopsin II. The ratio of the decay rates of free metarhodopsin II and metarhodopsin III-G-protein was increased as much as 10-fold in the presence of 25 mM NH2OH. The results indicate a mutual interdependence of retinal, opsin and 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