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 93 毫秒
1.
The substrate specificity of the different forms of rat liver biliverdin reductase was examined using synthetic biliverdins. Biliverdins carrying methyl, ethyl and one propionate residue in their structure were not substrates of biliverdin reductase. Biliverdins with one propionate and one acetate residue or with two acetate residues were not reduced by the enzyme either. The presence of two propionates in the biliverdin structure gave a biliverdin with substrate activity. Increasing the number of propionates to four, as in coprobiliverdins, did not affect substrate activity, while the octaacid urobiliverdins were also good substrates of the enzymes. The beta isomer of urobiliverdin III and coprobiliverdin III were reduced at much higher rates by molecular form 3 of the enzyme as compared to molecular form 1, a fact which had already been observed with the beta isomer of biliverdins IX, XIII and hematobiliverdin. All the biliverdins mentioned above were readily reduced to bilirubins by sodium borohydride. The purified molecular forms 1 and 3 displayed sigmoidal kinetics with most of the biliverdins tested. The data were analyzed by nonlinear regression in a microcomputer and it was found that they fitted a model of a moderate cooperative dimer where both ES and ES2 are catalytically active. The Vm, Ks and the Hill numbers, nH, for biliverdin IX alpha and beta, hematobiliverdin IX alpha and beta, and several synthetic biliverdin isomers are given. Molecular form 2 showed classical Michaelian kinetics.  相似文献   

2.
Biliverdin IXbeta reductase (BVR-B) catalyzes the pyridine nucleotide-dependent production of bilirubin-IXbeta, the major heme catabolite during early fetal development. BVR-B displays a preference for biliverdin isomers without propionates straddling the C10 position, in contrast to biliverdin IXalpha reductase (BVR-A), the major form of BVR in adult human liver. In addition to its tetrapyrrole clearance role in the fetus, BVR-B has flavin and ferric reductase activities in the adult. We have solved the structure of human BVR-B in complex with NADP+ at 1.15 A resolution. Human BVR-B is a monomer displaying an alpha/beta dinucleotide binding fold. The structures of ternary complexes with mesobiliverdin IValpha, biliverdin IXalpha, FMN and lumichrome show that human BVR-B has a single substrate binding site, to which substrates and inhibitors bind primarily through hydrophobic interactions, explaining its broad specificity. The reducible atom of both biliverdin and flavin substrates lies above the reactive C4 of the cofactor, an appropriate position for direct hydride transfer. BVR-B discriminates against the biliverdin IXalpha isomer through steric hindrance at the bilatriene side chain binding pockets. The structure also explains the enzyme's preference for NADP(H) and its B-face stereospecificity.  相似文献   

3.
Amorphous isomerically pure biliverdin IX alpha is readily prepared in more than 70% yield by dehydrogenation of bilirubin with 2,3-dichloro-5,6-dicyanobenzoquinone in dimethyl sulphoxide under carefully controlled conditions. Crystalline biliverdin IX alpha and amorphous [14C]biliverdin can be obtained similarly in more than 40+ yield. The pure crystalline pigment was characterized by elemental analysis, methylation, chemical and enzymic reduction to bilirubin, i.r.- and u.v.-visible-absorption spectroscopy, n.m.r. spectroscopy and field-desorption mass spectrometry, and its solubility was determined. Under certain conditions, dehydrogenation, gave biliverdin contaminated with III alpha and XIII alpha isomers as a result of disproporationation of bilirubin. Formation of non-IX alpha isomers depends on the concentrations of the reagents and the order in which they are mixed, and occurs under neutral anaerobic conditions. Free-radical reactions probably are responsible, suggesting that the first step in the deydrogenation of bilirubin with 2,3-dichloro-5,6-dicyanobenzoquinone in dimethyl sulphoxide is formation of a bilirubin cation radical, rather than hydride ion abstraction.  相似文献   

4.
Neutral solvent systems were developed to isolate the alpha, beta, gamma, and delta isomers of biliverdin IX dimethyl ester by TLC. The individual free acids of biliverdin IX were obtained by saponification of the corresponding dimethyl esters. The bilirubin IX isomers were prepared by reducing the corresponding biliverdin IX isomers with NaBH3CN. Starting from a pure biliverdin IX dimethyl ester, the corresponding free acid of biliverdin IX or bilirubin IX was available within 3-4 h. Preparation of spectrally pure bile pigment required final TLC on acid-cleaned neutral TLC plates. The absorption spectra of the free acids and dimethyl esters of biliverdin IX in methanol showed a broad band at about 650 nm and a sharp band at about 375 nm. The long-wave-length band was extremely sensitive to the presence of strong acid. A 10-fold molar excess of HCl caused a 35- to 50-nm shift of the absorption maximum to longer wavelengths and near doubling of the maximum absorption. The molar absorption coefficients of biliverdins were identical for each free acid and dimethyl ester pair. In each case, Beer's law was followed in both methanol and acidified methanol. Methanol also proved to be a suitable solvent for spectroscopic determination of the non-alpha isomers of bilirubin IX. The wavelength of maximum absorption and molar absorption coefficient of each dipyrrolic ethyl anthranilate azo pigment derived from the various bilirubin IX isomers are also reported.  相似文献   

5.
Etiolated Avena seedlings grown in the presence of 4-amino-5-hexynoic acid, an inhibitor of 5-aminolevulinic acid synthesis in plants, contain less than 10% of the spectrally detectable levels of phytochrome found in untreated seedlings (Elich, T.D., and Lagarias, J.C. (1988) Plant Physiol. 88, 747-751). In this study, incubation of explants from such seedlings with [14C]biliverdin IX alpha led to rapid covalent incorporation of radiolabel into a single 124-kDa polypeptide in soluble protein extracts. Immunoprecipitation experiments confirmed that this protein was phytochrome. Parallel experiments were performed with four unlabeled linear tetrapyrroles, the naturally occurring biliverdin IX alpha isomer, two non-natural isomers, biliverdin XIII alpha and biliverdin III alpha, and phycocyanobilin-the cleaved prosthetic group of the light-harvesting antenna protein C-phycocyanin. In all cases, except for the III alpha isomer of biliverdin, a time-dependent recovery of photoreversible phytochrome was observed. The newly formed phytochrome obtained after incubation with biliverdin IX alpha exhibited spectral characteristics identical with those of the native protein. In contrast, the spectral properties of phytochromes formed during incubation with biliverdin XIII alpha and phycocyanobilin differed significantly from those of the native chromoprotein. These results indicate that biliverdin IX alpha is an intermediate in the biosynthesis of the phytochrome chromophore and that phytochromes with prosthetic groups derived from bilatrienes having non-natural D-ring substituents are photochromic.  相似文献   

6.
The substrate specificity of rat liver biliverdin reductase was probed using helical and extended biliverdins. The former were the ZZZ-all-syn biliverdins IX alpha and IX gamma, and the latter were the 5Z-syn, 10Z-syn, 15Z-anti; 5Z-anti, 10E-anti, 15E-anti biliverdins. It was found that the reduction rates of the biliverdins increased with the progressive stretching of their conformations. The most extended biliverdin was reduced at a higher rate than biliverdin IX alpha. The chemical reduction rates to bilirubins followed a similar pattern. Nucleophilic addition of 2-mercaptoethanol to the C10 methine was also favored in the extended biliverdins.  相似文献   

7.
Oxidative cleavage of hematohemin IX in pyridine solution in the presence of ascorbic acid (coupled oxidation), followed by esterification of the products with boron trifluoride/methanol produced the four possible hematobiliverdin dimethyl esters in 11.1% overall yield. Transetherifications took place simultaneously with the esterification reaction and resulted in the formation of the dimethyl ester of hematobiliverdin IX gamma 8a,13a-dimethyl ether (1.8%), the dimethyl ester of hematobiliverdin IX beta 13a,18a-dimethyl ether (1.9%), the dimethyl ester of hematobiliverdin IX delta 8a-monomethyl ether (1.4%), and the dimethyl ester of hematobiliverdin IX alpha 18a-monomethyl ether (0.4%). The latter was the sole product obtained after the enzymatic oxidation of hematohemin with heme oxygenase, after esterification of the reaction product with boron trifluoride/methanol. When the esterification step was omitted hematobiliverdin IX alpha was obtained from the enzymatic oxidation. The structures of the hematobiliverdin derivatives were secured by their NMR and mass spectra data. Saponification of the dimethyl esters afforded the hematobiliverdin methyl ethers, which were excellent substrates of biliverdin reductase and were readily reduced to the corresponding bilirubins. Hematobiliverdin IX alpha was also a good substrate of biliverdin reductase. It is concluded that the enzymatic oxidation of hematohemin IX by heme oxygenase is alpha-selective, while biliverdin reductase shows no selectivity in the reduction of the four hematobiliverdin isomers.  相似文献   

8.
Biliverdin reductase was purified from pig spleen soluble fraction to a purity of more than 90% as judged by sodium dodecyl sulfate-polyacrylamide gel electrophoresis. The enzyme was a monomer protein with a molecular weight of about 34,000. Its isoelectric point was at 6.1-6.2. The enzyme was strictly specific to biliverdin and no other oxiodoreductase activities could be detected in the purified enzyme preparation. The purified enzyme could utilize both NADPH and NADH as electron donors for the reduction of biliverdin. However, there were considerable differences in the kinetic properties of the NADPH-dependent and the NADH-dependent biliverdin reductase activities: Km for NADPH was below 5 microM while that for NADH was 1.5-2 mM; the pH optimum of the reaction with NADPH was 8.5 whereas that of the reaction with NADH was 6.9; Km for biliverdin in the NADPH system was 0.3 microM whereas that in the NADH system was 1-2 microM. In addition, both the NADPH-dependent and NADH-dependent activities were inhibited by excess biliverdin, but this inhibition was far more pronounced in the NADPH system than in the NADH system. IX alpha-biliverdin was the most effective substrate among the four biliverdin isomers, and the dimethylester of IX alpha-biliverdin could not serve as a substrate. Biliverdin reductase was also purified about 300-fold from rat liver soluble fraction. The hepatic enzyme was also a monomer protein with a molecular weight of 34,000 and showed properties quite similar to those of the splenic enzyme as regards the biliverdin reductase reaction. The isoelectric point of the hepatic enzyme, however, was about 5.4. It was assumed that NADPH rather than NADH is the physiological electron donor in the intracellular reduction of IX alpha-biliverdin. The stimulatory effects of bovine and human serum albumins on the biliverdin reductase reactions were also examined.  相似文献   

9.
We report an HPLC method for separating the four regioisomers of verdoheme formed in the coupled oxidation of hemin with oxygen and ascorbate in aqueous pyridine. The reversed-phase ion-pair system uses hexafluoroacetone and pyridine as ion-pair agents. The regiochemistry of the separated isomers was established both by HPLC of the corresponding biliverdin IX derivatives and by 1H NMR of each isomer. Optical spectra of the pyridine verdohemochrome isomers were similar to each other, but showed differences in the absorption maxima in the red region, which appear at 680, 663, 648 and 660 nm for the alpha, beta, gamma, and delta-isomers, respectively. Each of the four isomers was incorporated anaerobically into heme oxygenase-1, yielding the corresponding verdoheme-enzyme complex. The ferrous forms had absorption maxima at 690, 667, 655, and 663 nm, and their CO-bound forms had maxima at 638, 624, 616, and 626 nm for alpha, beta, gamma, and delta-isomer, respectively. Addition of ferricyanide to the alpha-verdoheme-heme oxygenase complex brought about a ferric low-spin heme-like signal, which is identical with the ferric alpha-verdoheme complexed with the heme oxygenase that was observed in the heme oxygenase reaction.  相似文献   

10.
The amount and isomeric composition of urinary biliverdin in rabbits were analysed by h.p.l.c. Physiological values were maintained after the injection of haemin. On the other hand, when haemoglobins from several mammalian species were injected into rabbits, the excretion of biliverdin-IX alpha and biliverdin-IX beta were increased 6-18-fold and 32-66-fold respectively over physiological excretion. Injection of myoglobin resulted in a 44-fold increase in excretion of the IX alpha-isomer. Coupled oxidation with ascorbate of haemoglobin and myoglobin by oxygen produced mainly the IX alpha- and IX beta-isomers from haemoglobin and the IX alpha-isomer from myoglobin. The destruction of part of the haem from injected haemoproteins by non-enzymic chemical degradation would account for the observed respective increases in the excretion of biliverdin isomers. The excretion of biliverdin isomers after the injection of phenylhydrazine into rabbits was similar to that after the injection of haemoglobin.  相似文献   

11.
The unicellular red alga, Cyanidium caldarium, synthesizes phycocyanobilin from protoheme via biliverdin IX alpha. In vitro transformation of protoheme to biliverdin IX alpha and biliverdin IX alpha to phycobilins were previously shown to require NADPH, ferredoxin, and ferredoxin-NADP+ reductase, as well as specific heme oxygenase and phycobilin formation enzymes. The role of NADPH in these reactions was investigated in this study. The C. caldarium enzymatic activities that catalyze biliverdin IX alpha formation from protoheme, and phycobilin formation from biliverdin IX alpha, were partially purified by differential (NH4)2SO4 precipitation. The enzyme fractions, when supplemented with a light-driven ferredoxin-reducing photosystem I fraction derived from spinach leaves, catalyzed light-dependent transformation of protoheme to biliverdin IX alpha and biliverdin IX alpha to phycobilins, with or without the addition of NADPH and ferredoxin-NADP+ reductase. In the dark, neither reaction occurred unless NADPH and ferredoxin-NADP+ reductase were supplied. These results indicate that the only role of NADPH in both reactions of phycobilin biosynthesis, in vitro, is to reduce ferredoxin via ferredoxin-NADP+ reductase and that reduced ferredoxin can directly supply the electrons needed to drive both steps in the transformation of protoheme to phycocyanobilin.  相似文献   

12.
The liver, kidney and spleen of the mouse and rat and the kidney and spleen of the ox express a monomeric form of biliverdin reductase (Mr 34,000), which in the case of the ox kidney enzyme exists in two forms (pI 5.4 and 5.2) that are probably charge isomers. The livers of the mouse and rats express, in addition, a protein (Mr 46,000) that cross-reacts with antibodies raised against the ox kidney enzyme and may be related to form 2 described by Frydman, Tomaro, Awruch & Frydman [(1983) Biochim. Biophys. Acta 759, 257-263]. Higher-Mr forms appear to exist in the guinea pig and hamster. The ox kidney enzyme has three thiol groups, of which two are accessible to 5,5'-dithiobis-(2-nitrobenzoate) in the native enzyme. Immunocytochemical analysis reveals that biliverdin reductase is localized in proximal tubules of the inner cortex of the rat kidney. Biliverdin reductase antiserum also stains proximal tubules in human and ox kidney. The staining of podocytes in glomeruli of ox kidney with antiserum to aldose reductase is particularly prominent. The localization of biliverdin reductase in the inner cortical zone of rat kidney is similar to that described for glutathione S-transferase YfYf, and it is suggested that one function of this 'intracellular binding protein' may be to maintain a low free concentration of biliverdin to allow biliverdin reductase to operate efficiently.  相似文献   

13.
Rat liver biliverdin reductase was purified from control and bromobenzene-treated rats and was designated as C-BVR-T and B-BVR-T, respectively. Sodium dodecyl sulfate-polyacrylamide gel electrophoresis revealed the existence of two molecular weight variants (30,100 and 29,800) in C-BVR-T but only one form (30,100) in B-BVR-T. Western immunoblotting confirmed that both molecular weight variants were biliverdin reductase. Nondenaturing electrophoresis separated C-BVR-T and B-BVR-T preparations into groups of four variants, designated as BVR ND1 to ND4. However, the C-BVR-T preparation contained three major forms (BVR ND1, ND2, and ND3) while the B-BVR-T preparation contained two major forms (BVR ND2 and ND3). In vitro treatment of biliverdin reductase preparations with either bromobenzene or dithiothreitol did not interconvert the variants of the enzyme. QAE-Sepharose anion-exchange chromatography was used to isolate the ND2 and ND3 variants for physiochemical analysis. The amino acid composition of the variants was rather similar except for their Tyr content. Also, the peptide maps were similar except for a series of moderately early chromatographic peaks. These findings implied secondary modifications to the protein rather than substantial differences in primary structure. The pH-dependent cofactor requirements for enzyme activity were examined. Both variants exhibited 2 pH optima that were cofactor dependent; maximum activity with NADPH and NADH was observed at pH 8.5 and 6.7, respectively. However, both variants exhibited a higher catalytic rate with NADH than with NADPH at their pH optima. Furthermore, BVR ND3 exhibited a higher catalytic rate than BVR ND2 with either cofactor throughout the pH range 6.5-9.  相似文献   

14.
Rat liver biliverdin reductase exists in two molecular forms. The major one (molecular form 1) is transformed, under conditions of oxidative stress into another molecular form (molecular form 3) which is an S-S bridged dimer of form 1. The chemical modifications of the thiol, arginine and lysine residues of molecular form 1 which resulted in an inhibition of its catalytic activity did not affect the activity of molecular form 3. Rabbit polyclonal antibodies raised against form 1 did not recognize form 3. This lack of recognition persisted even when the dimer (form 3) was denatured with SDS or urea under non-reductive conditions. Reduction of form 3 with reduced thioredoxin gave the monomeric form 1, which was fully recognized by the antibodies. The latter recognized the biliverdin reductases from rat spleen and kidney to the same extent as they did with form 1. Molecular form 1 was completely inhibited by the addition of the antibodies. This inhibition was prevented by preincubation of the enzyme with either the substrate (biliverdin) or the cosubstrate (NADPH). Preincubation with the latter or with NADP+ (but not with bilirubin) strongly impaired the recognition of form 1 by the antibodies. Modification of the lysine or arginine residues of form 1 which were involved in substrate binding, impaired the interaction of the enzyme with the antibodies. The antisera blocked the enzymatic conversion of form 1 to form 3, but alkylation of the thiol residue involved in this dimerization had no effect on the interaction of form 1 with the antibodies. The lack of recognition of form 3 by the antibodies suggest that the antigenic site of the former becomes buried upon dimerization.  相似文献   

15.
Biliverdin and bilirubin mono- and di-beta-glucuronides were prepared by nucleophilic substitution of the 1-O-mesyl derivative of alpha-ethoxyethyl-protected glucuronic acid (compound II) with the tetrabutylammonium salts of biliverdin and bilirubin. Removal of the acetal-protecting groups by mild acid treatment yielded biliverdin glucuronides, which were reduced to bilirubin glucuronides. Depending on reaction conditions the pure beta-anomers or mixtures highly enriched in the beta-anomers were obtained. The biliverdin and bilirubin glucuronides were identical with pigments derived from bile. They were characterized as the IX alpha isomers and the beta-anomers by alkaline hydrolysis, n.m.r. spectroscopy, hydrolysis with beta-glucuronidase and conversion into dipyrrolic azopigments. Model reactions of the 1-O-mesylate (II) with other nucleophiles also were performed, i.e. the acetate anion and various alcohols.  相似文献   

16.
An enzyme extract from the phycocyanin-containing unicellular rhodophyte, Cyanidium caldarium, reductively transforms biliverdin IX alpha to phycocyanobilin, the chromophore of phycocyanin, in the presence of NADPH. Unpurified cell extract forms both 3(E)-phycocyanobilin, which is identical to the major pigment that is released from phycocyanin by methanolysis, and 3(Z)-phycocyanobilin, which is obtained as a minor methanolysis product. After removal of low molecular weight material from the cell extract, only 3(Z)-phycocyanobilin is formed. 3(E)-Phycocyanobilin formation from biliverdin IX alpha, and the ability to isomerize 3(Z)-phycocyanobilin to 3(E)-phycocyanobilin, are reconstituted by the addition of glutathione to the incubation mixture. Partially purified protein fractions derived from the initial enzyme extract form 3(Z)-phycocyanobilin plus two additional, violet colored bilins, upon incubation with NADPH and biliverdin IX alpha. Further purified protein fractions produce only the violet colored bilins from biliverdin IX alpha. One of these bilins was identified as 3(Z)-phycoerythrobilin by comparative spectrophotometry, reverse-phase high pressure liquid chromatography, and 1H NMR spectroscopy. A C. caldarium protein fraction catalyzes the conversion of 3(Z)-phycoerythrobilin to 3(Z)-phycocyanobilin. This fraction also catalyzes the conversion of 3(E)-phycoerythrobilin to 3(E)-phycocyanobilin. The conversion of phycoerythrobilins to phycocyanobilins requires neither biliverdin nor NADPH. The synthesis of phycoerythrobilin and its conversion to phycocyanobilin by extracts of C. caldarium, a species that does not contain phycoerythrin, indicates that phycoerythrobilin is a biosynthetic precursor to phycocyanobilin. The enzymatic conversion of the ethylidine group from the Z to the E configuration suggests that the E-isomer is the precursor to the protein-bound chromophore.  相似文献   

17.
Cell-free extract of the unicellular rhodophyte, Cyanidium caldarium catalyzes enzymatic reduction of biliverdin IX alpha to phycocyanobilin, the chromophore of the light-harvesting phycobiliprotein, phycocyanin. The enzyme activity is soluble, and the required reductant is NADPH. The extract has been separated into three protein fractions, all of which are required to reconstitute biliverdin reduction. One fraction contains ferredoxin, which was identified by its absorption spectrum. This fraction could be replaced with commercial ferredoxin derived from spinach or the red alga, Porphyra umbilicalis. The second protein fraction contains ferredoxin-NADP+ reductase, which was identified by the ability to catalyze ferredoxin-dependent reduction of cytochrome c in the presence of NADPH. This fraction could be replaced with commercial spinach ferredoxin-NADP+ reductase. These two components appear to be identical to previously described components of the algal heme oxygenase system that catalyzes biliverdin IX alpha formation from protoheme in C. caldarium extracts. The third protein fraction, in the presence of the first two (or their commercial counterparts) plus NADPH, catalyzes the reduction of biliverdin IX alpha to phycocyanobilin. The results indicate that the transformation of biliverdin to phycocyanobilin catalyzed by C. caldarium extracts is a ferredoxin-linked reduction process. The results also suggest the possibility that heme oxygenation and biliverdin reduction may occur in C. caldarium on associated enzyme systems.  相似文献   

18.
Rat-liver biliverdin reductase exists in two molecular forms. The major form 1 has a molecular mass of 34 kDa, while the minor form 2 has a molecular mass of 56 kDa. Form 1 was converted into a second major form (form 3) with a molecular mass of 68 kDa by a NAD+-dependent peroxisomal dehydrogenase which was induced under conditions of oxidative stress [Frydman, R. B., Tomaro, M. L., Awruch, J. & Frydman, B. (1984) Biochem. Biophys. Res. Commun. 121, 249]. Molecular form 1 from rat kidney was not affected by the dehydrogenase, and a structural explanation for this difference was therefore sought. Both form 1 biliverdin reductases, isolated from rat liver and kidney, were purified to homogeneity using affinity chromatography, FPLC and HPLC techniques. The homogeneous enzymes were found to be identical when compared by their HPLC retention times, amino acid compositions and electrophoretic behaviour on polyacrylamide gels under non-denaturing conditions and on SDS/polyacrylamide gels. On HPLC analysis the peptides resulting from the CNBr cleavage were found to be the same for both enzymes, when either the native enzymes or their thioethylpyridine derivatives were compared. When the HPLC fingerprints of the tryptic digests were compared, they were found to be very similar, except for a peptide eluting at 31.60 min in the liver digest and at 23.60 min in the kidney digest. When the enzyme from both origins was alkylated with 4-dimethylaminoazobenzene-4'-iodoacetamide and then digested with trypsin, the HPLC fingerprints of the alkylated cysteine-carrying peptides were almost identical, except for a peptide with a retention time of 19.03 min in the liver digest and of 18.19 min in the kidney digest. The liver reductase was not amenable to Edman degradation suggesting a block at the NH2-terminus; in the kidney enzyme, however, it was free and an NH2-terminal sequence of 12 amino acids could be determined. The liver enzyme was found to be more sensitive toward p-hydroxymercuriphenyl sulfonate than the kidney enzyme.  相似文献   

19.
Utilizing an in vitro coupled assay system, we show that isolated plastids from cucumber cotyledons convert the linear tetrapyrrole biliverdin IX alpha to the free phytochrome chromophore, phytochromobilin, which assembles with oat apophytochrome to yield photoactive holoprotein. The spectral properties of this synthetic phytochrome are indistinguishable from those of the natural photoreceptor. The plastid-dependent biliverdin conversion activity is strongly stimulated by both NADPH and ATP. Substitution of the nonnatural XIII alpha isomer of biliverdin for the IX alpha isomer affords a synthetic holophytochrome adduct with blue-shifted difference spectra. These results, together with experiments using boiled plastids, indicate that phytochromobilin synthesis from biliverdin is enzyme-mediated. Experiments where NADPH (and ATP) levels in intact developing chloroplasts are manipulated by feeding the metabolites 3-phosphoglycerate, dihydroxyacetone phosphate, and glucose 6-phosphate or by illumination with white light, support the hypothesis that the enzyme that accomplishes this conversion, phytochromobilin synthase, is plastid-localized. It is therefore likely that all of the enzymes of the phytochrome chromophore biosynthetic pathway reside in the plastid.  相似文献   

20.
A partially purified protein fraction from the phycocyanin-containing unicellular rhodophyte, Cyanidium caldarium, reductively transforms biliverdin IX alpha to a violet colored bilin in the presence of NADPH, ferredoxin, and ferredoxin-NADP+ reductase. This bilin has a violin-like absorption spectrum with maxima at 335 and 560 nm in methanolic HCl and at 337, 567, and 603-604 nm in CHCl3. The bilin has been determined to be 15,16-dihydrobiliverdin IX alpha by comparative spectrophotometry and 1H NMR spectroscopy. This product of biliverdin IX alpha reduction is converted enzymatically to phycobilins by further reduction. A general biosynthetic pathway is proposed which accounts for the formation of the phycobilins from biliverdin IX alpha by a two-step reduction process followed by isomerization.  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