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 31 毫秒
1.
A highly sensitive spectrofluorometric method for quantitative estimation of certain precursors of chlorophyll biosynthesis from the mixtures of plant tetrapyrroles having overlapping fluorescence emission spectra is developed. At room temperature (293 degrees K) protoporphyrin IX is monitored from its emission maximum, 633 nm, when excited at 400 nm (E400/F633). Protochlorophyllide is estimated at 638 nm, while being excited at 440 nm (E440/F638). Mg-protoporphyrin+Mg-protoporphyrin monoester pool has emission around 589-592 nm. Therefore the integration value of the emission band that extends from 580 to 610 nm is taken to calibrate its concentration. This spectrofluorometric method designed for the determination of protoporphyrin IX, esterified and nonesterified Mg-protoporphyrin pool, and protochlorophyllide is far superior to available spectrophotometric methods and estimates as low as 1 nM concentration of plant pigments. As minute quantities of individual pigments can be quantitatively analyzed from their mixtures, this method eliminates analytical uncertainties due to recovery losses caused by chromatography. However, only dilute samples can be estimated by this spectrofluorometric method as the quantitative relation between fluorescence and concentration deviates from linearity at high, i.e., above 150 nM, concentrations of pigment to be quantified.  相似文献   

2.
Yeast Saccharomyces cerevisiae accumulate in the presence of chelator 2,2'-dipyridyl, two fractions of photosensitizer protoporphyrin IX, which fluoresce with maxima at 625 and 635 nm. The two fractions were found to differ from each other in the dynamics and character of fluorescence photobleaching. In contrast to the protoporphyrin IX that fluoresces with a maximum at 625 nm, the protoporphyrin IX fraction that fluoresces with a maximum at 635 nm is more photolable; in addition, upon photobleaching of its fluorescence, a new maximum at 675 nm appears.  相似文献   

3.
A sensitive method using HPLC with fluorescence detection has been established for the measurement of porphyrins in biological materials. The assay recoveries were 88.0+/-1.8% for protoporphyrin IX in the blood, and ranged from 98.3+/-2.7% to 111.1+/-7.4% for various porphyrins in the urine. This method was employed to investigate the altered porphyrin profiles in rats after a single dose of various arsenicals including soluble sodium arsenate and sodium arsenite, and the relatively insoluble calcium arsenite, calcium arsenate and arsenic-contaminated soils at dose rates of 5 mg/kg or 0.5 mg/kg body weight. Porphyrin concentrations increased within 2448 hr after the arsenic treatment in blood and urine. Protoporphyrin IX is the predominant porphyrin in the blood. In rats administered 5 mg As(III)/kg body weight, protoporphyrin IX concentration elevated to 123% of the control values in rats, 24 hr after the treatment. Higher increases were recorded in the urinary protoporphyrin IX (253% at 24 hr; 397% on day 2), uroporphyrin (121% at 24 hr; 208% on day 2) and coproporphyrin III (391% at 24 hr; 304% on day 2), while there was no significant increase (109% on day 3) observed in the urinary coproporphyrin I excretion. In rats administered 5 mg As(V)/kg, urinary excretion of protoporphyrin LX, uroporphyrin, coproporphyrin III and coproporphyrin I elevated to the maximum levels by 48 hr with the corresponding percentage values compared to the control being 177%, 158%, 224% and 143%, respectively. In rats dosed with 5 mg As(III)/kg, the increases (expressed as % of the control values) of protoporphyrin IX in the blood were in the order: sodium arsenite (144%) > sodium arsenate (125%) > calcium arsenite (123%) > calcium arsenate. In contrast, there was no significant increase of protoporphyrin IX, when the six arsenic-contaminated cattle dip soils and nine copper chrome arsenate (CCA-contaminated) soils were administered to the rats. Probable explanations are discussed.  相似文献   

4.
The iron chelator 2,2'-dipyridyl (0.2 mM) more than fourfold increased the concentration of protoporphyrin IX and also of its zinc-containing complex in mitochondria of the yeast Saccharomyces cerevisiae. Protoporphyrin IX and a chlorine derivative of protoporphyrin IX which fluoresces at 670-675 nm were found in isolated plasma membranes of the yeast grown in the presence of 0.2 mM 2,2'-dipyridyl. The accumulation of endogenous porphyrins resulted in intensification of lipid photoperoxidation in mitochondria and plasma membranes and in a dramatically increased sensitivity of the cells to visible light (400-600 nm). The relative contribution of photodestruction of subcellular structures to photoinduced cell inactivation is discussed.  相似文献   

5.
Methods developed for porphyrin analysis have low recoveries and/or poor precision for the less polar protoporphyrin IX. We describe a simple method of analysis of porphyrins and biliverdin in bile and excreta of birds based on extraction with HCl 3N: acetonitrile and HPLC/UV analyses. Recoveries were good for protoporphyrin IX and other porphyrins (>79%). Applications of this method showed that porphyrins and biliverdin in birds excreta are mainly of biliary-fecal origin rather than urinary origin. Biliverdin and protoporphyrin IX increased proportionately more than the rest of the porphyrins and coproporphyrin III increased more than coproporphyrin I in the bile of Pb-poisoned mallards.  相似文献   

6.
Fluorescence emission of free protoporphyrin IX (PPIX, em. approximately 626 nm), zinc protoporphyrin IX (ZPP, em. approximately 594 nm) and fluorescent heme degradation product (FHDP, em. approximately 466 nm) are identified and simultaneously detected in mouse and human red cell hemolysates, when excited at 365 nm. A novel method is established for comparing relative FHDP, PPIX and ZPP levels in hemolysates without performing red cell porphyrin extractions. The ZPP fluorescence directly measured in hemolysates (F(365/594)) correlates with the ZPP fluorescence obtained from acetone/water extraction (R(2) = 0.9515, P < 0.0001). The relative total porphyrin (ZPP and PPIX) fluorescence obtained from direct hemolysate fluorescence measurements also correlates with red blood cell total porphyrins determined by ethyl acetate extraction (Piomelli extraction, R(2) = 0.88, P < 0.0001). These fluorescent species serves as biomarkers for alterations in Hb synthesis and Hb stability.  相似文献   

7.
In the presence of Mg-GTP, the rat liver guanylate cyclase, in either intact membranes or trypsin solubilized form, was stimulated by protoporphyrin IX 6 to 10-fold. However, when Mn-GTP was the substrate, protoporphyrin IX activated the membrane-bound guanylate cyclase only 50%, in contrast to the marked activation reported for the cytosolic enzyme. Meso- and deuteroporphyrin IX, hematoporphyrin and coproporphyrin III also activated membrane guanylate cyclase while uroporphyrin III, and hemin had no effect. Basal, Mg2+-dependent activity exhibited two classes of catalytic sites with apparent Km values of 2 mM and 0.12 mM. Activation by protoporphyrin resulted in the disappearance of the low affinity sites. The activated enzyme exhibited Michaelis-Menten kinetics and no alteration in its requirement for excess Mg2+. These data indicate that, in the presence of Mg2+, a heme-like structure can interact with the membrane-bound guanylate cyclase and regulate its activity.  相似文献   

8.
Several naturally occurring porphyrins and porphyrins used in photodynamic therapy inhibit glutathione S-transferase isoenzymes either purified from rat liver or lung or in cytosol from normal and from cancerous (Morris 7288C hepatoma) liver. Although differences occur in the type and amount of transferases in normal and cancerous liver and in the liver of rats bearing an extrahepatic tumour, these enzymes are potential binding sites for porphyrins. Porphyrin structure is an important factor in determining the affinity of binding, as shown by the relative inhibitory effectiveness. Of the dicarboxylic porphyrins in the mixture used clinically, OO'-diacetylhaematoporphyrin and monohydroxyethylmonovinyldeuteroporphyrin are more effective inhibitors than haematoporphyrin and protoporphyrin IX. Of the naturally occurring porphyrins the order of effectiveness is protoporphyrin IX (dicarboxylic) greater than coproporphyrin (tetracarboxylic) greater than uroporphyrin (octacarboxylic) and type I greater than type III isomers of both uroporphyrin and coproporphyrin, and the synthetic tetra-meso-phenylporphinetetrasulphonate is a better inhibitor (apparent Ki = 250 nM) than coproporphyrin, which contains a comparable number of negative charges. In addition, iron-porphyrin chelates are more effective inhibitors of the transferases, with 25-fold decrease in Ki value, than the free porphyrins. These results indicate that one means whereby porphyrins accumulate in tissues is the occupation of intracellular binding sites, such as the transferases. Since porphyrins inhibit the activity of these important detoxifying enzymes, there will be metabolic consequences to the cell.  相似文献   

9.
The hemY gene of the Bacillus subtilis hemEHY operon is essential for protoheme IX biosynthesis. Two previously isolated hemY mutations were sequenced. Both mutations are deletions affecting the hemY reading frame, and they cause the accumulation of coproporphyrinogen III or coproporphyrin III in the growth medium and the accumulation of trace amounts of other porphyrinogens or porphyrins intracellularly. HemY was found to be a 53-kDa peripheral membrane-bound protein. In agreement with recent findings by Dailey et al. (J. Biol. Chem. 269:813-815, 1994) B. subtilis HemY protein synthesized in Escherichia coli oxidized coproporphyrinogen III and protoporphyrinogen IX to coproporphyrin and protoporphyrin, respectively. The protein is not a general porphyrinogen oxidase since it did not oxidize uroporphyrinogen III. The apparent specificity constant, kcat/Km, for HemY was found to be about 12-fold higher with coproporphyrinogen III as a substrate compared with protoporphyrinogen IX as a substrate. The protoporphyrinogen IX oxidase activity is consistent with the function of HemY in a late step of protoheme IX biosynthesis, i.e., HemY catalyzes the penultimate step of the pathway. However, the efficient coproporphyrinogen III to coproporphyrin oxidase activity is unexplained in the current view of protoheme IX biosynthesis.  相似文献   

10.
Alterations of the chemical structure of protoporphyrin IX markedly altered the activation of soluble guanylate cyclase purified from bovine lung. Hydrophobic side chains at positions 2 and 4 and vicinal propionic acid residues at positions 6 and 7 of the porphyrin ring (protoporphyrin IX, mesoporphyrin IX) were essential for maximal enzyme activation (Ka = 7-8 nM; Vmax = 6-8 mumol of cGMP/min/mg). Substitution of hydrophobic with polar groups (hematoporphyrin IX, coproporphyrin III), or with hydrogen atoms ( deuteroporphyrin IX), and methylation of propionate residues resulted in decreased enzyme stimulation. Stimulatory porphyrins increased the Vmax and the apparent affinities of enzyme for MgGTP and uncomplexed Mg2+. An open central core in the porphyrin ring was essential for enzyme activation. The pyrrolic nitrogen adduct, N-phenylprotoporphyrin IX, was inhibitory and competitive with protoporphyrin IX (KI = 73 nM). Similarly, metalloporphyrins inhibited enzymatic activity and ferro-protoporphyrin IX (KI = 350 nM), zinc-protoporphyrin IX (KI = 50 nM) and manganese-protoporphyrin IX (KI = 9 nM) were competitive with protoporphyrin IX. Inhibitory porphyrins and metalloporphyrins also prevented enzyme activation by S-nitroso-N- acetylpenicillamine and NO. Guanylate cyclase reconstituted with such porphyrins required higher concentrations of protoporphyrin IX for further activation and were not activated by NO. Thus, porphyrins, metalloporphyrins, and NO appeared to interact at a common binding site on guanylate cyclase. This common site is likely that which normally binds heme and, therefore, NO-heme when the heme-containing enzyme is exposed to NO. Thus, NO and nitroso compounds may react with enzyme-bound heme to generate a modified porphyrin which structurally resembles protoporphyrin IX in its interaction with guanylate cyclase.  相似文献   

11.
Coproporphyrinogen oxidase, the sixth enzyme in the biosynthetic heme pathway, catalyzes the oxidative decarboxylation of coproporphyrinogen III to protoporphyrinogen IX. A reversed-phase high pressure liquid chromatography method was developed to measure coproporphyrinogen oxidase enzymatic activity in rat liver. With this method, the separation, identification and quantification of coproporphyrin III (oxidized substrate) and protoporphyrin IX (oxidized product) present in the assays could be carried out with no need of derivatization and in less than 15 min. Rat and human liver coproporphyrinogen oxidase basal activities determined using this method were 0.41+/-0.05 nmol of protoporphyrin IX/h per mg of hepatic protein and 0.87+/-0.06 protoporphyrin IX/h per mg of hepatic protein, respectively. Kinetic studies showed that optimum pH for rat CPGox is 7.3, and that its activity is linear in the range of protein concentrations and incubation times assayed. The present paper describes a sensitive, specific and rapid fluorometric high performance liquid chromatography method to measure coproporphyrinogen oxidase, which could be applied to the diagnosis of human coproporphyria, and which is also suitable for the study of lead and other metal poisoning that produce alterations in this enzymatic activity.  相似文献   

12.
A temperature-sensitive respiration-deficient mutant of yeast lacks hemoproteins and accumulates coproporphyrin III when cultivated at elevated temperatures. Cells grown at 20 C respired normally and contained cytochromes a, b, and c. Cells grown at 35 C showed respiration-deficient mutant characters; they did not respire, lacked cytochromes, and accumulated coproporphyrin III. Addition of protoporphyrin IX or protohemin IX to the culture medium restored the respiratory activity of this mutant during growth at 35 C. The activities of various enzymes, including succinate-2,6-dichlorophenol indophenol (DCPIP), reduced nicotinamide adenine dinucleotide (NADH(2))-DCPIP, succinate-cytochrome c, and NADH(2)-cytochrome c oxidoreductase, and cytochrome oxidase, and the cytochrome c content of cells cultured in various conditions were determined. Changes in the number and structure of mitochondria were associated with changes in respiratory activity.  相似文献   

13.
Cain G. D. and Bassow F. 1976. Porphyrins in the perienteric fluid of Ascaris lumbricoides. International Journal for Parasitology6: 79–82. Porphyrins in the perienteric fluid of adult female A. lumbricoides were esterified in methanolic H2SO4, extracted in chloroform, separated by thin-layer chromatography, and identified spectrophotometrically before and after conversion to their zinc and copper chelates. Protoporphyrin IX was the major component, comprising 95·4% of the total; the remaining 4·6 % was coproporphyrin III. Uroporphyrin was not detected; no porphyrins were recovered from other worm tissues. Fluid from worms with light and dark colored guts varied in protoporphyrin content from 0·58 to 4·08 nmoles/ml, respectively, but fluid from both groups contained similar molar ratios of protoporphyrin, coproporphyrin and heme.  相似文献   

14.
Porphyrin-Accumulating Mutants of Escherichia coli   总被引:17,自引:9,他引:8       下载免费PDF全文
Four mutants (pop-1, pop-6, pop-10, and pop-14) which accumulate a red water-insoluble pigment were obtained in Escherichia coli K-12 AB1621. For each mutant, the red pigment was shown to be protoporphyrin IX, a late precursor of heme. Mutagenic treatment of mutant pop-1 yielded a secondary mutant, pop-1 sec-20, which accumulated a brown water-soluble pigment. The brown pigment was shown to be coproporphyrin III. Mutant pop-1 resembled the parental strain in its cytochrome absorption spectrum, catalase activity, and ability to grow on nonfermentable carbon and energy sources; therefore, its ability to produce and utilize heme was unimpaired. Judged on the same criteria, the secondary mutant, pop-1 sec-20, was partially heme and respiratory deficient. Growth in anaerobic conditions decreased by 25% the accumulation of protoporphyrin by pop-1; under the same conditions, pop-1 sec-20 did not accumulate coproporphyrin or coproporphyrinogen. The mutations causing protoporphyrin accumulation in all four pop mutants were found to map in the lac to purE (10-13 min) region of the E. coli chromosome. In the case of mutant pop-1, the mutation was shown to be strongly linked to the tsx locus (12 min). In mutant pop-1 sec-20, the second mutation causing coproporphyrin accumulation was co-transducible with the gal locus at a frequency of 88 to 96%. The mechanism of porphyrin accumulation by the mutants is discussed.  相似文献   

15.
For the first time, the enzymatic one-electron oxidation of several naturally occurring and synthetic water-soluble porphyrins by peroxidases was investigated by ESR and optical spectroscopy. The ESR spectra of the free radical metabolites of the porphyrins were singlets (g = 2.0024, delta H = 2-3 G), which we assigned to their respective porphyrin pi-cation free radicals. Several porphyrins were investigated and ranked by the intensity of their ESR spectra (coproporphyrin III greater than coproporphyrin I greater than deuteroporphyrin IX greater than mesoporphyrin IX greater than Photofrin II greater than protoporphyrin IX greater than uroporphyrin I greater than uroporphyrin III greater than hematoporphyrin IX). The porphyrins were oxidized by several peroxidases (horseradish peroxidase, lactoperoxidase, and myeloperoxidase), yielding the same type of ESR spectra. From these results, we conclude that porphyrins are substrates for peroxidases. The changes in the visible absorbance spectra of the porphyrins during enzymatic oxidation were monitored. The two-electron oxidation product, which was assigned to the dihydroxyporphyrin, was detected as an intermediate of the oxidation process. The optical spectrum of the porphyrin pi-cation free radical was not detected, probably due to its low steady-state concentration.  相似文献   

16.
We describe fluorometric assays for two enzymes of the heme pathway, coproporphyrinogen oxidase and protoporphyrinogen oxidase. Both assays are based on measurement of protoporphyrin IX fluorescence generated from coproporphyrinogen III by the two consecutive reactions catalyzed by coproporphyrinogen oxidase and protoporphyrinogen oxidase. Both enzymatic activities are measured by recording protoporphyrin IX fluorescence increase in air-saturated buffer in the presence of EDTA (to inhibit ferrochelatase that can further metabolize protoporphyrin IX) and in the presence of dithiothreitol (that prevents nonenzymatic oxidation of porphyrinogens to porphyrins). Coproporphyrinogen oxidase (limiting) activity is measured in the presence of a large excess of protoporphyrinogen oxidase provided by yeast mitochondrial membranes isolated from commercial baker's yeast. These membranes are easy to prepare and are stable for at least 1 year when kept at -80 degrees C. Moreover they ensure maximum fluorescence of the generated protoporphyrin (solubilization effect), avoiding use of a detergent in the incubation medium. The fluorometric protoporphyrinogen oxidase two-step assay is closely related to that already described (J.-M. Camadro, D. Urban-Grimal, and P. Labbe, 1982, Biochem. Biophys. Res. Commun. 106, 724-730). Protoporphyrinogen is enzymatically generated from coproporphyrinogen by partially purified yeast coproporphyrinogen oxidase. The protoporphyrinogen oxidase reaction is then initiated by addition of the membrane fraction to be tested. However, when very low amounts of membrane are used, low amounts of Tween 80 (less than 1 mg/ml) have to be added to the incubation mixture to solubilize protoporphyrin IX in order to ensure optimal fluorescence intensity. This detergent has no effect on the rate of the enzymatic reaction when used at concentrations less than 2 mg/ml. Activities ranging from 0.1 to 4-5 nmol protoporphyrin formed per hour per assay are easily and reproducibly measured in less than 30 min.  相似文献   

17.
The action of porphyrins, uroporphyrin I and III (URO I and URO III), pentacarboxylic porphyrin I (PENTA I), coproporphyrin I and III (COPRO I and COPRO III), protoporphyrin IX (PROTO IX) and mesoporphyrin (MESO), on the activity of human erythrocytes delta-aminolevulinic acid dehydratase, porphobilinogenase, deaminase and uroporphyrinogen decarboxylase in the dark and under UV light was investigated. Both photoinactivation and light-independent inactivation was found in all four enzymes using URO I as sensitizer. URO III had a similar action as URO I on porphobilinogenase and deaminase and PROTO IX exerted equal effect as URO I on delta-aminolevulinic acid dehydratase and uroporphyrinogen decarboxylase. Photodynamic efficiency of the porphyrins was dependent on their molecular structure. Selective photodecomposition of enzymes by URO I, greater specificity of tumor uptake by URO I and enhanced porphyrin synthesis by tumors from delta-aminolevulic acid, with predominant formation of URO I, underline the possibility of using URO I in detection of malignant cells and photodynamic therapy.  相似文献   

18.
An investigation of the behavior of protoporphyrin IX, deuteroporphyrin IX, haematoporphyrin IX and coproporphyrin III in aqueous solution revealed extensive and complex aggregation processes. Protoporphyrin appears to be highly aggregated under all conditions studied. At concentrations below 4 muM, aggregation of deutero-, haemato- and coproporphyrin is probably restricted to dimerization. At approx. 4muM each of these three porphyrins exhibits sharp changes in spectra consistent with a "micellization" process to form large aggregates of unknown size. This critical concentration increases with increasing temperature and pH, but is not very sensitive to variation in ionic strength. Temperature-jump kinetic studies on deuteroporphyrin also imply an initial dimerization process, the rate constants for which are comparable with those for various synthetic porphyrins, followed by a further extensive aggragation. The ability of a particular porphyrin to dimerize appears to parallel that of the corresponding iron(III) complexes (ferrihaems), although it is thought that ferrihaems do not exhibit further aggregation under these conditions.  相似文献   

19.
Due to the potential special position of protoporphyrin IX in the evolution of photosynthesis, the absorption and fluorescence characteristics of this pigment and its complexes with human serum albumin (HSA) and basic proteinoid have been studied in parallel with their photochemical activity. The most significant change in the absorption spectrum of PP IX was the appearance of a new maximum at 455 (or 461) nm in the presence of HSA or proteinoid respectively. Some changes in the physicochemical properties of PP IX in different microenvironments have been detected by changes in fluorescence emission and excitation spectra (intensity, quantum yields, position of maxima). The increase of fluorescence quantum yield resulting from the formation of PP IX complexes with HSA or proteinoid correlates with the increase of their photochemical activity. Results obtained are discussed from the point of view of the early evolution of the photosynthetic apparatus.  相似文献   

20.
During heme biosynthesis in Escherichia coli two structurally unrelated enzymes, one oxygen-dependent (HemF) and one oxygen-independent (HemN), are able to catalyze the oxidative decarboxylation of coproporphyrinogen III to form protoporphyrinogen IX. Oxygen-dependent coproporphyrinogen III oxidase was produced by overexpression of the E. coli hemF in E. coli and purified to apparent homogeneity. The dimeric enzyme showed a Km value of 2.6 microm for coproporphyrinogen III with a kcat value of 0.17 min-1 at its optimal pH of 6. HemF does not utilize protoporphyrinogen IX or coproporphyrin III as substrates and is inhibited by protoporphyrin IX. Molecular oxygen is essential for the enzymatic reaction. Single turnover experiments with oxygen-loaded HemF under anaerobic conditions demonstrated electron acceptor function for oxygen during the oxidative decarboxylation reaction with the concomitant formation of H2O2. Metal chelator treatment inactivated E. coli HemF. Only the addition of manganese fully restored coproporphyrinogen III oxidase activity. Evidence for the involvement of four highly conserved histidine residues (His-96, His-106, His-145, and His-175) in manganese coordination was obtained. One catalytically important tryptophan residue was localized in position 274. None of the tested highly conserved cysteine (Cys-167), tyrosine (Tyr-135, Tyr-160, Tyr-170, Tyr-213, Tyr-240, and Tyr-276), and tryptophan residues (Trp-36, Trp-123, Trp-166, and Trp-298) were found important for HemF activity. Moreover, mutation of a potential nucleotide binding motif (GGGXXTP) did not affect HemF activity. Two alternative routes for HemF-mediated catalysis, one metal-dependent, the other metal-independent, are proposed.  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