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.
In many crustacea, including the lobster, the bulk of the vitamin A of the whole animal is concentrated in the eyes. Recently Fisher, Kon, and Thompson found that vitamin A extracted from the eyes of euphausiid crustacea has only about one half the biological potency of the same amount of the all-trans acetate or fish liver vitamin A. In the present experiments the vitamin A of the lobster eye is found to consist almost entirely of the hindered cis isomer, neo-b, the precursor in the vertebrate retina of the visual pigments rhodopsin and iodopsin. This isomer is known to have a low biological potency in the rat, only about one quarter that of all-trans vitamin A. In the lobster eye it is virtually all extractable with petroleum ether, about 30 per cent in the form of free alcohol, about 70 per cent in the form of esters. It was identified by its absorption spectrum, as derived from measurements on crude extracts, and measured directly in purified preparations; the changes in absorption which accompany isomerization; oxidation to the corresponding retinene; and synthesis from the latter of rhodopsin. The examination of an extract of euphausiid eyes, provided by Dr. Kon, also revealed the presence of neo-b vitamin A virtually alone. This may be the characteristic condition in the eyes of Eucarid crustacea. It is peculiar in that the neo-b isomer, being a sterically hindered form, is ordinarily expected to be represented in any equilibrium mixture of geometric isomers in very small amount. Apparently certain crustacea have ways of circumventing the difficulties implicit in producing and retaining this isomer, and store it in the eye virtually alone.  相似文献   

2.
Rhodopsin, the red photosensitive pigment of rod vision, is composed of a specific cis isomer of retinene, neo-b (11-cis), joined as chromophore to a colorless protein, opsin. We have investigated the thermal denaturation of cattle rhodopsin and opsin in aqueous digitonin solution, and in isolated rod outer limbs. Both rhodopsin and opsin are more stable in rods than in solution. In solution as well as in rods, moreover, rhodopsin is considerably more stable than opsin. The chromophore therefore protects opsin against denaturation. This is true whether rhodopsin is extracted from dark-adapted retinas, or synthesized in vitro from neo-b retinene and opsin. Excess neo-b retinene does not protect rhodopsin against denaturation. The protection involves the specific relationship between the chromophore and opsin. Similar, though somewhat less, protection is afforded opsin by the stereoisomeric iso-a (9-cis) chromophore in isorhodopsin. The Arrhenius activation energies (Ea) and entropies of activation (ΔS‡) are much greater for thermal denaturation of rhodopsin and isorhodopsin than of opsin. Furthermore, these values differ considerably for rhodopsins from different species —frog, squid, cattle—presumably due to species differences in the opsins. Heat or light bleaches rhodopsin by different mechanisms, yielding different products. Light stereoisomerizes the retinene chromophore; heat denatures the opsin. Photochemical bleaching therefore yields all-trans retinene and native opsin; thermal bleaching, neo-b retinene and denatured opsin.  相似文献   

3.
Rhodopsin is formed by the condensation of opsin with a cis isomer of retinene, called neo-b. The bleaching of rhodopsin releases all-trans retinene which must be isomerized back to neo-b in order for rhodopsin to regenerate. Both retinene isomers are in equilibrium with the corresponding isomers of vitamin A, through the alcohol dehydrogenase system. An enzyme is found in cattle retinas and frog pigment layers which catalyzes the interconversion of all-trans and neo-b retinene. We call it "retinene isomerase." It is soluble in neutral phosphate buffer, and precipitates between 20 and 35 per cent saturation with ammonium sulfate. In the dark, the isomerase converts all-trans and neo-b retinene to an equilibrium mixture of 5 parts neo-b and 95 parts all-trans. With opsin present to trap neo-b, the isomerase catalyzes the synthesis of rhodopsin from all-trans retinene. This reaction, however, is too slow to account for dark adaptation. Retinene is isomerized by light, but too slowly to supply the retina with neo-b. In aqueous solution the pseudoequilibrium mixture contains about 15 per cent neo-b. When all-trans retinene is irradiated in the presence of isomerase, the rate of formation of neo-b is increased about 5 times, and the pseudoequilibrium shifted so that the mixture now contains about 32 per cent neo-b. The isomerase is specific for all-trans and neo-b retinene. It does not act on two other cis isomers of retinene, nor on all-trans or neo-b vitamin A. The role of the isomerase in vision appears to be as follows: in the light, as rhodopsin is bleached to opsin and all-trans retinene, the latter is in part converted to the neo-b isomer and stored in the pigment epithelium as neo-b vitamin A. During dark adaptation, the dominant process is the trapping by opsin of neo-b retinene supplied from stores of neo-b vitamin A, and the slow isomerase-catalyzed "dark" conversion of all-trans to neo-b retinene.  相似文献   

4.
Cis-trans isomers of vitamin A and retinene in the rhodopsin system   总被引:28,自引:14,他引:14  
Vitamin A and retinene, the carotenoid precursors of rhodopsin, occur in a variety of molecular shapes, cis-trans isomers of one another. For the synthesis of rhodopsin a specific cis isomer of vitamin A is needed. Ordinary crystalline vitamin A, as also the commercial synthetic product, both primarily all-trans, are ineffective. The main site of isomer specificity is the coupling of retinene with opsin. It is this reaction that requires a specific cis isomer of retinene. The oxidation of vitamin A to retinene by the alcohol dehydrogenase-cozymase system displays only a low degree of isomer specificity. Five isomers of retinene have been isolated in crystalline condition: all-trans; three apparently mono-cis forms, neoretinenes a and b and isoretinene a; and one apparently di-cis isomer, isoretinene b. Neoretinenes a and b were first isolated in our laboratory, and isoretinenes a and b in the Organic Research Laboratory of Distillation Products Industries. Each of these substances is converted to an equilibrium mixture of stereoisomers on simple exposure to light. For this reaction, light is required which retinene can absorb; i.e., blue, violet, or ultraviolet light. Yellow, orange, or red light has little effect. The single geometrical isomers of retinene must therefore be protected from low wave length radiation if their isomerization is to be avoided. By incubation with opsin in the dark, the capacity of each of the retinene isomers to synthesize rhodopsin was examined. All-trans retinene and neoretinene a are inactive. Neoretinene b yields rhodopsin indistinguishable from that extracted from the dark-adapted retina (λmax· 500 mµ). Isoretinene a yields a similar light-sensitive pigment, isorhodopsin, the absorption spectrum of which is displaced toward shorter wave lengths (λmax· 487 mµ). Isoretinene b appears to be inactive, but isomerizes preferentially to isoretinene a, which in the presence of opsin is removed to form isorhodopsin before the isomerization can go further. The synthesis of rhodopsin in solution follows the course of a bimolecular reaction, as though one molecule of neoretinene b combines with one of opsin. The synthesis of isorhodopsin displays similar kinetics. The bleaching of rhodopsin, whether by chemical means or by exposure to yellow or orange (i.e., non-isomerizing) light, yields primarily or exclusively all-trans retinene. The same appears to be true of isorhodopsin. The process of bleaching is therefore intrinsically irreversible. The all-trans retinene which results must be isomerized to active configurations before rhodopsin or isorhodopsin can be regenerated. A cycle of isomerization is therefore an integral part of the rhodopsin system. The all-trans retinene which emerges from the bleaching of rhodopsin must be isomerized to neoretinene b before it can go back; or if first reduced to all-trans vitamin A, this must be isomerized to neovitamin Ab before it can regenerate rhodopsin. The retina obtains new supplies of the neo-b isomer: (a) by the isomerization of all-trans retinene in the eye by blue or violet light; (b) by exchanging all-trans vitamin A for new neovitamin Ab from the blood circulation; and (c) the eye tissues may contain enzymes which catalyze the isomerization of retinene and vitamin A in situ. When the all-trans retinene which results from bleaching rhodopsin in orange or yellow light is exposed to blue or violet light, its isomerization is accompanied by a fall in extinction and a shift of absorption spectrum about 5 mµ toward shorter wave lengths. This is a second photochemical step in the bleaching of rhodopsin. It converts the inactive, all-trans isomer of retinene into a mixture of isomers, from which mixtures of rhodopsin and isorhodopsin can be regenerated. Isorhodopsin, however, is an artefact. There is no evidence that it occurs in the retina; nor has isovitamin Aa or b yet been identified in vivo. In rhodopsin and isorhodopsin, the prosthetic groups appear to retain the cis configurations characteristic of their retinene precursors. In accord with this view, the β-bands in the absorption spectra of both pigments appear to be cis peaks. The conversion to the all-trans configuration occurs during the process of bleaching. The possibility is discussed that rhodopsin may represent a halochromic complex of a retinyl ion with opsin. The increased resonance associated with the ionic state of retinene might then be responsible both for the color of rhodopsin and for the tendency of retinene to assume the all-trans configuration on its release from the complex. A distinction must be made between the immediate precursor of rhodopsin, neovitamin Ab, and the vitamin A which must be fed in order that rhodopsin be synthesized in vivo. Since vitamin A isomerizes in the body, it is probable that any geometrical isomer can fulfill all the nutritional needs for this vitamin.  相似文献   

5.
11-cis-Retinol has previously been shown in physiological experiments to promote dark adaptation and recovery of photoresponsiveness of bleached salamander red cones but not of bleached salamander red rods. The purpose of this study was to evaluate the direct interaction of 11-cis-retinol with expressed human and salamander cone opsins, and to determine by microspectrophotometry pigment formation in isolated salamander photoreceptors. We show here in a cell-free system using incorporation of radioactive guanosine 5′-3-O-(thio)triphosphate into transducin as an index of activity, that 11-cis-retinol inactivates expressed salamander cone opsins, acting an inverse agonist. Similar results were obtained with expressed human red and green opsins. 11-cis-Retinol had no significant effect on the activity of human blue cone opsin. In contrast, 11-cis-retinol activates the expressed salamander and human red rod opsins, acting as an agonist. Using microspectrophotometry of salamander cone photoreceptors before and after bleaching and following subsequent treatment with 11-cis-retinol, we show that 11-cis-retinol promotes pigment formation. Pigment was not formed in salamander red rods or green rods (containing the same opsin as blue cones) treated under the same conditions. These results demonstrate that 11-cis-retinol is not a useful substrate for rod photoreceptors although it is for cone photoreceptors. These data support the premise that rods and cones have mechanisms for handling retinoids and regenerating visual pigment that are specific to photoreceptor type. These mechanisms are critical to providing regenerated pigments in a time scale required for the function of these two types of photoreceptors.11-cis-Retinol is the precursor to 11-cis-retinal, the 11-cis-aldehyde form of vitamin A and the chromophore that combines covalently with rod and cone opsin proteins to form visual pigments. 11-cis-Retinal is consumed during visual signaling, and its continual synthesis is required. Photon absorption by the visual pigments causes the isomerization of its chromophore to the all-trans configuration. This initiates two processes critical for vision: activation of the photoreceptor cell and the eventual recovery of the original photosensitivity of the cells, requiring regeneration of the visual pigments. As cones are used for bright light vision, these two processes must work more rapidly in cones than in rods and thus cones have a higher requirement of 11-cis-retinoids as suggested by Rushton (1, 2).Photoreceptor activation begins with photoisomerization of the chromophore within the visual pigment. This results in a subsequent conformational change of the protein part of the visual pigment that is able to activate its G protein transducin, which in turn activates a PDE that lowers the concentration of cGMP and closes cGMP-gated ion channels. These steps comprise the visual signal transduction cascade (see Ref. 3 for review).The visual cycle involves regeneration of the visual pigment, which ultimately deactivates the protein and accomplishes the recovery of the photosensitivity of the photoreceptor cell. Classically, this process involves both the photoreceptor cell and the retinal pigment epithelium (RPE).4 After photoisomerization of the chromophore and formation of the active visual pigment, all-trans-retinal is released from the opsin and reduced to all-trans-retinol, which is then transported to the RPE where it is isomerized to 11-cis-retinol through a number of steps. In the RPE, 11-cis-retinol is oxidized to the aldehyde form, which is transported back to the photoreceptor cell and can be directly used by all of the opsins to regenerate an inactive pigment ready for photoactivation. The details of this model have been extensively reviewed (4, 5). Alternatively, recent work suggests that cones have an additional source of 11-cis-retinoids from Müller cells (68). Like the RPE cells, Müller cells have been shown to be able to convert all-trans-retinol to 11-cis-retinol (6). Unlike in the RPE cells, 11-cis-retinol is not oxidized to 11-cis-retinal in Müller cells.Jones et al. (9) demonstrated that administration of 11-cis-retinol to bleached salamander red cones could restore photosensitivity. A logical conclusion was that red cones were able to oxidize 11-cis-retinol to the aldehyde and regenerate visual pigments although noncovalent binding of 11-cis-retinol to red cone opsins generating a light-sensitive complex could not be excluded. On the other hand, 11-cis-retinol does not restore photosensitivity to bleached salamander rod cells but appears to directly activate the cells (9, 10). The data suggested that the rods were not able to oxidize 11-cis-retinol, but that the retinol itself could activate the signal transduction cascade, and indeed we recently demonstrated that 11-cis-retinol acts as an agonist to expressed bovine rod opsin (11). Our aim here was to study the action of 11-cis-retinol on cone opsins and cone photoreceptor cells to determine the efficacy of an alternate visual cycle for cones.The photoreceptor cells used in this study are from tiger salamander, and the expressed opsins used for biochemical experiments are those from salamander and human. Photoreceptor cells are generally identified by cell morphology and the type of opsin it contains that can be further complicated by the findings that some cone cells have multiple opsins (12, 13). Recently genetic analysis has determined that opsins fall into five classes (reviewed in Refs. 14 and 15). We have studied opsins falling into four of these classes and use common color-derived names for the opsins and photoreceptor cells. The classic rod cells used for scotopic vision contain rhodopsin, the visual pigment for the rod opsin (RH1 opsin) and appeared red and thus have been designated as red rods. Some species such as salamanders have an additional rod cell whose photosensitivity is blue-shifted from that of the red rod and thus designated as green rods. In the tiger salamander, the green rods contain the identical opsin (SWS2 opsin) found in blue cones (16). The human blue cones contain an opsin from a different class (SWS1 opsin), which is homologous to the salamander UV cone opsin. The human red and green and salamander red cone opsins all belong to the same class of opsins (M/LWS opsins). Absorption properties of visual pigments are further modulated in some animals including the tiger salamander by use of 11-cis-retinal with an additional double bond (3,4-dehydro or A2 11-cis-retinal) resulting in red-shifted absorbance from pigments containing 11-cis-retinal (A1 11-cis-retinal).We show here that 11-cis-retinol is not an agonist to cone opsins and does not itself generate a light-sensitive opsin. We further show using microspectrophotometry that both red and blue salamander cone cells regenerate visual pigments from 11-cis-retinol, whereas pigments could not be regenerated with 11-cis-retinol in bleached salamander red and green rods even though the latter contains the same opsin as the salamander blue cone. Thus, rods and cones have mechanisms for handling retinoids and regenerating visual pigment that are specific to photoreceptor type, and these mechanisms are critical to providing regenerated pigments in a time scale required for the function of these two types of photoreceptors.  相似文献   

6.
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   

7.
A number of mutant strains of the green alga, Scenedesmus obliquus, when grown in the dark, accumulated ζ-carotene as their major carotenoid together with appreciable concentrations of phytoene and phytofluene. The phytoene was almost entirely the 15-cis isomer, and phytofluene was also present mainly as the 15-cis form, whereas the ζ-carotene could be separated into three isomers, predominantly all-trans ζ-carotene accompanied by the 15-cis and an unidentified cis isomer. All the ζ-carotene isomers, when illuminated in the presence of iodine, gave the same equilibrium mixture of stereo-isomers, including a product with unusual spectroscopic and chromatographic properties, which may be a cyclic compound. The pathway of carotenoid biosynthesis in S. obliquus is discussed. On illumination, most of the ζ-carotenic strains were killed, but PGI strain survived, due to the production of cyclic carotenoids with chromophores long enough to protect chlorophyll from photosensitized oxidation.  相似文献   

8.
To assign the observed vibrationsl modes in the resonance Raman spectrum of the retinylidene chromophore of rhodopsin, we have studied chemically modified retinals. The series of analogs investigated are the n-butyl retinals substituted at C9 and C13. The results obtained for the 11-cis isomer have clearly assigned the CCH3 vibrational frequencies observed in the spectrum of the retinylidene chromophore. The data show that the C(9)CH3 stretching vibration can be assigned to the vibrational mode observed in the 1017 cm?1 region, and the vibration detected at 997 cm?1 can be assigned to the C(13CH3 vibration. The C(5)CH3 stretching mode does not contribute to the vibrations observed in this region. The splitting in the C(n)CH3 (n = 9, 13) vibration is characteristic of the 11-cis conformation. The results on the modified retinals do not support the hypothesis that the splitting arises from equilibrium mixtures of 11-cis, 12-s-cis and 11-cis, 12-s-trans in solution. Thus, this splitting cannot be used to determine whether the chromophore in rhodopsin is in a 12-s-cis or 12-s-trans conformation. However, our results demonstrate that there are other vibrational modes in the spectra which are sensitive to this conformational equilibrium and we use the presence of a strong ~ 1271 cm?1 mode in bovine and squid rhodopsin spectra as an indication that the chromophore in these pigments is 11-cis, 12-s-trans.  相似文献   

9.
The major product from the reduction of (2R,3R)-dihydroquercetin with sodium borohydride is the 2,3-trans-3,4-trans isomer of leucocyanidin [(2R,3S,4R-3,3′,4,4′,5,7-hexahydroxyflavan] whereas the enzymatic reduction product is the 2,3-trans-3,4-cis isomer [(2R,3S,4S)-3,3′,4,4′,5,7-hexahydroxyflavan]. The 3,4-trans isomer may be partly converted to the 3,4-cis isomer under mild acid conditions. The 3,4-cis isomer is more acid-labile, and more reactive both chemically with thiols and enzymatically with a diol reductase, than the 3,4-trans isomer.  相似文献   

10.
The G protein coupled receptor rhodopsin contains a pocket within its seven-transmembrane helix (TM) structure, which bears the inactivating 11-cis-retinal bound by a protonated Schiff-base to Lys296 in TM7. Light-induced 11-cis-/all-trans-isomerization leads to the Schiff-base deprotonated active Meta II intermediate. With Meta II decay, the Schiff-base bond is hydrolyzed, all-trans-retinal is released from the pocket, and the apoprotein opsin reloaded with new 11-cis-retinal. The crystal structure of opsin in its active Ops* conformation provides the basis for computational modeling of retinal release and uptake. The ligand-free 7TM bundle of opsin opens into the hydrophobic membrane layer through openings A (between TM1 and 7), and B (between TM5 and 6), respectively. Using skeleton search and molecular docking, we find a continuous channel through the protein that connects these two openings and comprises in its central part the retinal binding pocket. The channel traverses the receptor over a distance of ca. 70 Å and is between 11.6 and 3.2 Å wide. Both openings are lined with aromatic residues, while the central part is highly polar. Four constrictions within the channel are so narrow that they must stretch to allow passage of the retinal β-ionone-ring. Constrictions are at openings A and B, respectively, and at Trp265 and Lys296 within the retinal pocket. The lysine enforces a 90° elbow-like kink in the channel which limits retinal passage. With a favorable Lys side chain conformation, 11-cis-retinal can take the turn, whereas passage of the all-trans isomer would require more global conformational changes. We discuss possible scenarios for the uptake of 11-cis- and release of all-trans-retinal. If the uptake gate of 11-cis-retinal is assigned to opening B, all-trans is likely to leave through the same gate. The unidirectional passage proposed previously requires uptake of 11-cis-retinal through A and release of photolyzed all-trans-retinal through B.  相似文献   

11.
《Inorganica chimica acta》1988,148(2):209-213
cis-[Co(ox)(tetb]ClO4 is readily isolated from the reaction between tetb (tetb = rac-Me6-cyclam = 5,5,7(R,S),12,12,14(R,S)-hexamethyl-1,4,8,11-tetraazacyclotetradecane) and K3Co(ox)3·3H2O in aqueous solution. Removal of the coordinated oxalato ligand by acid hydrolysis (with HCl/HClO4) results in the formation of α-trans-(RRRR,SSSS), [CoCl2{(SSeq,RReq)-tetb}]ClO4 with both six-membered rings in the twist conformation. The (RR)-(+)-tartrate (−2) ion coordinates enantioselectively with this isomer to give (–)-cis-[Co2{μ-(RR)-(+)-(tart)}{(RR)-tetb}2](ClO4)2 and this, in turn, reacts with HCl/HClO4 to give (–)-α-trans-(1S,4S,8S,11S)-[CoCl12{7Req,14Req)-tetb}]ClO4. The absolute configuration of the tetraamine obtained in the resolution procedure was established by synthesizing the α-trans-dichloro isomer using tetb of known absolute configuration.  相似文献   

12.
The mechanism of retinol isomerization in the vertebrate retina visual cycle remains controversial. Does the isomerase enzyme RPE65 operate via nucleophilic addition at C11 of the all-trans substrate, or via a carbocation mechanism? To determine this, we modeled the RPE65 substrate cleft to identify residues interacting with substrate and/or intermediate. We find that wild-type RPE65 in vitro produces 13-cis and 11-cis isomers equally robustly. All Tyr-239 mutations abolish activity. Trp-331 mutations reduce activity (W331Y to ∼75% of wild type, W331F to ∼50%, and W331L and W331Q to 0%) establishing a requirement for aromaticity, consistent with cation-π carbocation stabilization. Two cleft residues modulate isomerization specificity: Thr-147 is important, because replacement by Ser increases 11-cis relative to 13-cis by 40% compared with wild type. Phe-103 mutations are opposite in action: F103L and F103I dramatically reduce 11-cis synthesis relative to 13-cis synthesis compared with wild type. Thr-147 and Phe-103 thus may be pivotal in controlling RPE65 specificity. Also, mutations affecting RPE65 activity coordinately depress 11-cis and 13-cis isomer production but diverge as 11-cis decreases to zero, whereas 13-cis reaches a plateau consistent with thermal isomerization. Lastly, experiments using labeled retinol showed exchange at 13-cis-retinol C15 oxygen, thus confirming enzymatic isomerization for both isomers. Thus, RPE65 is not inherently 11-cis-specific and can produce both 11- and 13-cis isomers, supporting a carbocation (or radical cation) mechanism for isomerization. Specific visual cycle selectivity for 11-cis isomers instead resides downstream, attributable to mass action by CRALBP, retinol dehydrogenase 5, and high affinity of opsin apoproteins for 11-cis-retinal.  相似文献   

13.
AIM: To investigate the interaction of reconstituted rhodopsin, 9-cis-retinal-rhodopsin and 13-cis-retinal-rhodopsin with transducin, rhodopsin kinase and arrestin-1. METHODS: Rod outer segments(ROS) were isolated from bovine retinas. Following bleaching of ROS membranes with hydroxylamine, rhodopsin and rhodopsin analogues were generated with the different retinal isomers and the concentration of the reconstituted pigments was calculated from their UV/visible absorption spectra. Transducin and arrestin-1 were purified to homogeneity by column chromatography, and an enriched-fraction of rhodopsin kinase was obtainedby extracting freshly prepared ROS in the dark. The guanine nucleotide binding activity of transducin was determined by Millipore filtration using β,γ-imido-(3H)-guanosine 5'-triphosphate. Recognition of the reconstituted pigments by rhodopsin kinase was determined by autoradiography following incubation of ROS membranes containing the various regenerated pigments with partially purified rhodopsin kinase in the presence of(γ-32P) ATP. Binding of arrestin-1 to the various pigments in ROS membranes was determined by a sedimentation assay analyzed by sodium dodecyl sulphatepolyacrylamide gel electrophoresis. RESULTS: Reconstituted rhodopsin and rhodopsin analogues containing 9-cis-retinal and 13-cis-retinal rendered an absorption spectrum showing a maximum peak at 498 nm, 486 nm and about 467 nm, respectively, in the dark; which was shifted to 380 nm, 404 nm and about 425 nm, respectively, after illumination. The percentage of reconstitution of rhodopsin and the rhodopsin analogues containing 9-cis-retinal and 13-cis-retinal was estimated to be 88%, 81% and 24%, respectively. Although only residual activation of transducin was observed in the dark when reconstituted rhodopsin and 9-cis-retinal-rhodopsin was used, the rhodopsin analogue containing the 13-cis isomer of retinal was capable of activating transducin independently of light. Moreover, only a basal amount of the reconstituted rhodopsin and 9-cis-retinal-rhodopsin was phosphorylated by rhodopsin kinase in the dark, whereas the pigment containing the 13-cis-retinal was highly phosphorylated by rhodopsin kinase even in the dark. In addition, arrestin-1 was incubated with rhodopsin, 9-cis-retinal-rhodopsin or 13-cis-retinal-rhodopsin. Experiments were performed using both phosphorylated and non-phosphorylated regenerated pigments. Basal amounts of arrestin-1 interacted with rhodopsin, 9-cis-retinal-rhodopsin and 13-cis-retinal-rhodopsin under dark and light conditions. Residual arrestin-1 was also recognized by the phosphorylated rhodopsin and phosphorylated 9-cis-retinal-rhodopsin in the dark. However, arrestin-1 was recognized by phosphorylated 13-cis-retinal-rhodopsin in the dark. As expected, all reformed pigments were capable of activating transducin and being phosphorylated by rhodopsin kinase in a lightdependent manner. Additionally, all reconstituted photolyzed and phosphorylated pigments were capable of interacting with arrestin-1. CONCLUSION: In the dark, the rhodopsin analogue containing the 13-cis isomer of retinal appears to fold in a pseudo-active conformation that mimics the active photointermediate of rhodopsin.  相似文献   

14.
1. Rhodesian copalwood (Guibourtia coleosperma) contains three diastereo-isomeric leuco-fisetinidins. These consist of the (−)-2,3-cis–3,4-cis (2R,3R,4R) and (−)-2,3-cis–3,4-trans (2R,3R,4S) 3′,4′,7-trihydroxyflavan-3,4-diols, and the third was shown to be a 2,3-trans–3,4-cis isomer by means of paper ionophoresis. 2. There occurrence in similar proportions as tannin precursors also in the tropical hardwoods G. tessmannii and G. demeusii implies a close taxonomic relationship between these, and with G. coleosperma. 3. Epimerization of the natural (−)-3′,4′,7- trihydroxy-2,3-trans-flavan-3,4-trans-diol affords a mixture from which the (−)-2,3-cis–3,4-cis isomer was separated readily, but the (−)-2,3-trans–3,4-cis isomer was obtained with difficulty. These were formed by epimerization of the (−)-2,3-trans–3,4-trans isomer at C-2 and C-4, and at C-4, respectively.  相似文献   

15.
The rhodopsin system of the squid   总被引:6,自引:19,他引:6  
Squid rhodopsin (λmax 493 mµ)—like vertebrate rhodopsins—contains a retinene chromophore linked to a protein, opsin. Light transforms rhodopsin to lumi- and metarhodopsin. However, whereas vertebrate metarhodopsin at physiological temperatures decomposes into retinene and opsin, squid metarhodopsin is stable. Light also converts squid metarhodopsin to rhodopsin. Rhodopsin is therefore regenerated from metarhodopsin in the light. Irradiation of rhodopsin or metarhodopsin produces a steady state by promoting the reactions, See PDF for Equation Squid rhodopsin contains neo-b (11-cis) retinene; metarhodopsin all-trans retinene. The interconversion of rhodopsin and metarhodopsin involves only the stereoisomerization of their chromophores. Squid metarhodopsin is a pH indicator, red (λmax 500 mµ) near neutrality, yellow (λmax 380 mµ) in alkaline solution. The two forms—acid and alkaline metarhodopsin—are interconverted according to the equation, Alkaline metarhodopsin + H+ acid metarhodopsin, with pK 7.7. In both forms, retinene is attached to opsin at the same site as in rhodopsin. However, metarhodopsin decomposes more readily than rhodopsin into retinene and opsin. The opsins apparently fit the shape of the neo-b chromophore. When light isomerizes the chromophore to the all-trans configuration, squid opsin accepts the all-trans chromophore, while vertebrate opsins do not and hence release all-trans retinene. Light triggers vision by affecting directly the shape of the retinene chromophore. This changes its relationship with opsin, so initiating a train of chemical reactions.  相似文献   

16.
SEVERAL laboratories1–6 have recently been concerned with the mechanism of the bathochromic shift of about 120 nm which results when 11-cis retinal (λ max 380 nm) combines with the protein opsin to form rhodopsin (λmax 498 nm). A red shift of up to 186 nm is involved in the formation of iodopsin from 11-cis retinal and cone opsin7,8. The active site of bovine rhodopsin consists of the 11-cis retinylidene chromophore attached to a primary amine group of the protein forming a Schiff-base linkage of the type shown in Fig. 1, Ia. On the basis of the chemical reactions of rhodopsin and its derivatives it has been suggested that an interaction between a protonated form of the chromophore (structure of the type Ib) and a lipophilic environment contributes11 to the red shift.  相似文献   

17.
Such (+)- and (?)-cis-cycloheximide isomers as isocyclohcximide (1a, 1b), α-epiisocycloheximide (2a, 2b) and neocycloheximide (3a, 3b) were synthesized by aldol condensation of (?)-(2R, 4R)- and (+)-(2S, 4S)-cis-2,4-dimethyl-1-cyclohexanone (5a, 5b). obtained by microbial resolution, with 4-(2-oxoethyl)-2,6-piperidinedione (7). The absolute configuration of the (?)-cis-ketone 5a was confirmed by chemical correlation with natural (2S, 4S, 6S, αR)-cycloheximide (4). The newly synthesized isomer, (?)-α-epiisocycloheximide (2b), showed strong antimicrobial activity against S. cerevisiae andP. oryzae close to that of natural cycloheximide (4).  相似文献   

18.
In the malaria mosquito Anopheles gambiae polymorphic chromosomal inversions may play an important role in adaptation to environmental variation. Recently, we used microarray-based divergence mapping combined with targeted resequencing to map nucleotide differentiation between alternative arrangements of the 2La inversion. Here, we applied the same technique to four different polymorphic inversions on the 2R chromosome of An. gambiae. Surprisingly, divergence was much lower between alternative arrangements for all 2R inversions when compared to the 2La inversion. For one of the rearrangements, 2Ru, we successfully mapped a very small region (∼100 kb) of elevated divergence. For the other three rearrangements, we did not identify any regions of significantly high divergence, despite ample independent evidence from natural populations of geographic clines and seasonal cycling, and stable heterotic polymorphisms in laboratory populations. If these inversions are the targets of selection as hypothesized, we suggest that divergence between rearrangements may have escaped detection due to retained ancestral polymorphism in the case of the youngest 2R rearrangements and to extensive gene flux in the older 2R inversion systems that segregate in both An. gambiae and its sibling species An. arabiensis.MORE than 70 years ago Dobzhansky and Sturtevant (1938) first discovered polymorphic inversion arrangements carried by various Drosophila pseudoobscura populations. After observing correlations between environmental conditions and inversion frequencies, Dobzhansky proposed that inversions are under strong selection due to their role in promoting local adaptation to the heterogeneous conditions a species encounters both spatially and temporally (Dobzhansky 1944, 1948; Powell 1997). More recent studies have implicated chromosomal inversions in the adaptation of a diversity of eukaryotes including humans (Coluzzi et al. 1979; Feder et al. 2003; Hoffmann et al. 2004; Stefansson et al. 2005). Long known to be common in dipteran insects, more recent HapMap data suggest that polymorphic inversions may be numerous in human populations and by extension other mammals (Bansal et al. 2007). Given their potential importance in facilitating adaptation, surprisingly little is known about the mechanism(s) or the genes responsible for maintaining inversion polymorphisms in natural populations.Gene exchange between inverted and standard arrangements, although reduced, can still occur through gene flux: the action of gene conversion and multiple crossovers in inversion heterozygotes (heterokaryotypes) (Chovnick 1973; Navarro et al. 1997; Schaeffer and Anderson 2005). Over time allelic variation unrelated to ecological adaptation should become homogenized between arrangements, while alleles which are under divergent selection pressures should remain in linkage disequilibrium with each other and with the inversion itself, leading to heightened differentiation between standard and inverted arrangements at and near the target loci. In principle, this process allows the identification of specific loci involved in adaptive divergence (Schaeffer et al. 2003; Schaeffer and Anderson 2005; Storz 2005). Consistent with this model, previous low-resolution studies of Drosophila inversions revealed heterogeneous patterns of nucleotide diversity relative to divergence, as well as the interspersion of regions of high and low genetic association potentially due to the interaction of selection and gene flux (Schaeffer et al. 2003; Kennington et al. 2006; but see Munte et al. 2005). The application of high-resolution tools flowing from completely sequenced genomes will facilitate the mapping of genes that are the targets of divergent natural selection within gene arrangements.Although Drosophila has been the favored model, the African malaria vector Anopheles gambiae sensu stricto also provides an excellent system for studying the maintenance of inversion polymorphisms, not only within a species but across speciation events of different ages in the An. gambiae sibling species complex. The nominal species An. gambiae s.s. (hereafter, An. gambiae) is synanthropic: almost exclusively biting humans, resting indoors, and exploiting anthropogenic larval habitats (Coluzzi 1999). This close association with humans, vital to making An. gambiae one of the most proficient vectors of malaria, is likely to have been facilitated by chromosomal inversions thought to confer adaptive benefits in heterogeneous climatic and ecological settings in Africa. Seven common polymorphic inversions exist on the second chromosome. Six of these are located on the right arm (2R): j, b, c, u, d, and k, while 2La is the only inversion on the left arm (Coluzzi et al. 2002). Facilitated by the sequenced reference genome (Holt et al. 2002), some of the breakpoints for these polymorphic inversions have been localized to small genomic regions (Sharakhov et al. 2006; Coulibaly et al. 2007; Sangare 2007). Most of these inversions appear to be the targets of strong selection. Five of the inversions (2La and 2Rb, -c, -d, and -u) are nonrandomly associated with degree of aridity; each cycles seasonally with rainfall, and all except 2Ru form stable geographic clines in frequency from mesic forest to xeric regions bordering the Sahara (Coluzzi et al. 1979; Toure et al. 1994, 1998; Powell et al. 1999). Inversion 2Rj is not clinal, but its distribution in Mali is consistent with adaptation to novel rockpool niches (Coluzzi et al. 1985; Manoukis et al. 2008).In the An. gambiae species complex, inversion polymorphisms can be maintained across the boundaries of emerging and even full species. An. gambiae and its sibling An. arabiensis, strictly sympatric throughout most of their extensive ranges in sub-Saharan Africa, differ by multiple fixed chromosomal rearrangements on the X but share three chromosome 2 inversions: 2La, fixed in An. arabiensis and polymorphic in An. gambiae; and 2Rb and -c, polymorphic in both species (Coluzzi et al. 1979, 2002). Moreover, these same inversions and all other common An. gambiae inversions with the exception of 2Rj are shared and polymorphic in two lineages apparently undergoing ecological speciation within An. gambiae—the assortatively mating M and S molecular forms (della Torre et al. 2002, 2005). Inversion frequencies are correlated with climatic and ecological conditions in parallel in both lineages (Costantini et al. 2009; Simard et al. 2009). Unlike the full species, the M and S incipient species are not distinguished by any fixed inversion differences. Indeed, genomewide divergence mapping between the M and S forms revealed that significant differentiation was confined to two small low-recombination regions adjacent to the centromeres of 2L and X which are distant from any inversions (Turner et al. 2005). Thus, in distinction to models of speciation invoking inversions as facilitating the persistence of hybridizing species (Noor et al. 2001; Rieseberg 2001; Ortiz-Barrientos et al. 2002; Navarro and Barton 2003), the An. gambiae data suggest that chromosome 2 inversions are not directly responsible for reproductive isolation. Instead, the same chromosome 2 inversion polymorphisms appear to confer similar ecological benefits, within and across species boundaries. A long-term research goal is to identify the mechanisms and the genes controlling these processes.Previously we conducted the first high-density genomic scan of divergence across a chromosomal inversion (2La) in An. gambiae (White et al. 2007). By hybridizing genomic DNA from S form mosquitoes homokaryotypic for alternate gene arrangements on chromosome 2L (2La or 2L+a) to oligonucleotide microarrays we were able to measure divergence across the 22-Mb inversion at nearly 14,000 markers. Differentiation in the rearranged region was significantly higher than in collinear portions of chromosome 2L. Between breakpoints the pattern of differentiation was heterogeneous: two genomic clusters of significantly higher divergence were identified near but not adjacent to the breakpoints. Directed resequencing within the S form confirmed these results and suggested that both clusters contained genes targeted by selection. Observed levels of linkage disequilibrium between the 2La breakpoints and markers in the clusters are highly unlikely under a neutral scenario, in light of known recombination rates and plausible estimates of the age of the inversion.The present study characterizes the patterns of genetic variation in polymorphic rearrangements on the opposite (right) arm of chromosome 2: 2Rj, -b, -c, and -u. With the goal of identifying candidate genes maintaining these inversions in natural populations, we applied microarray-based divergence mapping to measure differentiation between alternative 2R arrangements. Because three of four inversions have taxonomic distributions that span incipient and/or completed speciation events, we validated the microarray findings by targeted sequencing in multiple taxa: sympatric Malian populations of An. gambiae M and S forms, and the sibling species An. arabiensis.  相似文献   

19.
A new reversed phase-high performance liquid chromatography method has been developed to analyze the full complement of higher plant photosynthetic pigments (cis-neoxanthin, neoxanthin, violaxanthin, taraxanthin, anteraxanthin, lutein, zeaxanthin, cis-lutein, chlorophyll b, chlorophyll a, α- and β-carotene). The separation is carried out on a C18 column in about 10 minutes, using a single high-pressure pump and three different mobile phases in three isocratic steps. This method introduces a major improvement in higher plant photosynthetic pigment analysis, resolving in only 10 minutes all photosynthetic pigments while achieving good separation of lutein from its isomer zeaxanthin. Zeaxanthin is involved in the xanthophyll cycle, which recently has been proposed to play a significant role in the protection of the photosynthetic apparatus from photoinhibitory conditions (Demmig et al. [1987] Plant Physiol 84: 218-224).  相似文献   

20.
Bacteriorhodopsin monomer dispersed in a solution of the detergent L-1690 could maintain the specific interaction between retinal and protein in the pH range 9.0-0.0 at 25°C. λmax of the absorbance spectrum was 550 nm at pH 9.0, 556 nm at pH 5.5, 609 nm at pH 2.1 and 570 nm at pH 0.0. Increasing the NaCl concentration in the solution promoted formation of the 609 nm product at pH 5.0-3.0 and also its transition to the 570 nm product at pH 2.5-1.0. Retinal isomer analysis gave a ratio of 13-cis- to all-trans-retinal of 53 : 47 at pH 5.5. When the pH of the solution was reduced, the relative content of all-trans-retinal increased and the ratio of 13-cis- to all-trans-retinal was 14 : 86 at pH 0.0. Illumination of the solution at pH 7.2 yielded a product containing 9-cis-retinal or 9-cis, 13-cis-retinal, which may be due to a reaction other than the photoreaction cycle.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号