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 281 毫秒
1.
The nature of the heme environment in methemalbumin, the Fe(III) protoporphyrin IX (heme)-human serum albumin (HSA) complex, was investigated by optical spectroscopy. Comparison of the optical spectra of methemalbumin, ferro-hemalbumin in the absence and presence of 2-methylimidazole, and their carbon monoxide derivatives with horseradish peroxidase (HRP) and its corresponding derivatives indicates that histidine is not present in the first coordination sphere of heme in methemalbumin and that the protein is devoid of a well-defined heme cavity. The complex exhibits peroxidase activity by catalyzing oxidation of 2,2'-azinobis(3-ethylbenzthiazoline-6-sulfonate) by hydrogen peroxide. Its activity ( K(M)=433 microM, molar catalytic activity=0.33 s(-1)), however, is considerably lower compared to HRP, indicating differences in the heme environments. Fluorescence intensity decays of Trp214 in HSA and methemalbumin, best fitted to a three-exponential model, gave the lifetimes 7.03 ns (30%), 3.17 ns (38%), and 0.68 ns (32%) for HSA and 8.04 ns (1.7%), 2.42 ns (19.7%), and 0.64 ns (78.6%) for methemalbumin. These lifetime values were further confirmed by a model-independent maximum entropy method. Similarity in the lifetimes and variations in the amplitudes suggest that while conformational heterogeneity of HSA is unperturbed on heme binding, redistribution of the populations of the three conformations occurs and the sub-state associated with the shortest lifetime dominates the total population by approximately 80%. Decay associated spectra (DAS) indicate that the observed lifetime variation with wavelength is predominantly due to ground state heterogeneity, though solvent dipolar relaxation also contributes. Time-resolved fluorescence anisotropy measurements of the Trp214 residue yielded information on motion within the protein together with the whole protein molecule. The binding of heme did not affect the rotational correlation time of the albumin molecule (approximately 20 ns). However, the motion of tryptophan within the protein matrix increased by a factor of approximately 3 (0.46 ns to 0.15 ns). This indicates that while the overall hydrodynamic volume of the albumin molecule remained the same, tryptophan underwent a more rapid internal rotation because of the efficient energy transfer to the bound heme. Optical studies, analysis of lifetime measurements, DAS, and anisotropy measurements together suggest that heme binds to a surface residue. The rapid internal motion of Trp214 during its excited state lifetime for the approximately 80% populated conformer of methemalbumin allows the orientation factor, kappa(2), to approach the average value of 2/3. From the time-resolved fluorescence measurements and the energy transfer calculations on methemalbumin, a Trp214-heme distance of 22 A was deduced.  相似文献   

2.
Synthetic polymer-bound hemin (iron(III) protoporphyrin IX) derivatives were effectively reduced by ferredoxin and ferredoxin-NADP reductase system. The resultant polymer-bound heme (iron(II) protoporphyrin IX) derivatives formed oxygen adducts with a lifetime of ca. 1 hr in aqueous solution at -30 degrees C. The reduction rate is discussed in terms of the structure of the hemin derivatives.  相似文献   

3.
The intraerythrocytic Plasmodium falciparum parasite converts most of host hemoglobin heme into a nontoxic heme crystal. Erythrocyte zinc protoporphyrin IX, normally present at 0.5 microM, which is a ratio of 1:40,000 hemes, can elevate 10-fold in some of the anemias associated with malaria disease protection. This work examines a binding mechanism for zinc protoporphyrin IX inhibition of heme crystallization similar to the antimalarial quinolines. Zinc protoporphyrin IX neither forms crystals alone nor extends on preformed heme crystals. Inhibition of both seed heme crystal formation and crystal extension occurs with an inhibitory concentration (IC)50 of 5 microM. Field emission in-lens scanning electron microscopy depicts the transition and inhibition of heme monomer aggregates to heme crystals with and without seeding of preformed hemozoin templates. In vitro zinc protoporphyrin IX, like the quinolines, binds to heme crystals in a saturable, specific, pH, and time-dependent manner. The ratio at saturation is approximately 1 zinc protoporphyrin IX per 250 hemes of the crystal. Unlike the quinolines, zinc protoporphyrin IX binds measurably in the absence of heme. Isolated ring and trophozoite stage parasites have an elevated zinc protoporphyrin IX to heme ratio 6 to 10 times that in the erythrocyte cytosol, which also corresponds to elevated ratios found in heme crystals purified from Plasmodium parasites. This work implicates protection from malaria by a mechanism where elevated zinc protoporphyrin IX in anemic erythrocytes binds to heme crystals to inhibit further crystallization. In endemic malaria areas, severe iron deficiency anemia should be treated with antimalarials along with iron replenishment.  相似文献   

4.
The porphyrin and tryptophan fluorescence of sperm whale apomyoglobin complexed with protoporphyrin IX has been studied in the pH range 2-13. It has been shown that the fluorescence and absorption spectra of protoporphyrin incorporated into the heme crevice remain constant in the pH range 5.5-10.8 but change significantly at pH less than 5.5 and pH greater than 10.8, due to the acid and alkaline denaturation, respectively, of the complex accompanied by dissociation of protoporphyrin IX. At the same pH ranges, the quantum yield of tryptophanyl fluorescence increases sharply as a result of removal of protoporphyrin, acting as a quencher, from the complex. Other parameters of tryptophanyl fluorescence (maximum position, halfwidth and spectrum shape) change in the alkaline region as well. In the acidic pH range, these parameters change only at pH less than 4.3, indicating that the Trp surroundings are more stable to denaturation than the heme crevice region. Between pH 5.5 and 10.9, where the complex of apomyoglobin with protoporphyrin IX is in its native state, the main parameters of tryptophan fluorescence remain unchanged except for the ratio I325/I350 which diminishes at pH greater than 9.5. Its alteration precedes the alkaline denaturation of the complex and can be explained by a local conformational change induced by the break of the 'salt bridges' essential for the maintenance of the native Mb structure in the N-terminal region. The fluorescence data obtained for apomyoglobin, myoglobin and the complex between protoporphyrin IX and apomyoglobin enable one to compare their structures and to evaluate the role of the porphyrin macrocycle and the iron atom in the formation of the native myoglobin structure and its functioning.  相似文献   

5.
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   

6.
The hemin moiety of horseradish peroxidase (donor:hydrogen-peroxide oxidoreductase, EC 1.11.1.7) was removed and the apoprotein reconstituted with the fluorescent protoporphyrin IX. Steady-state and time-resolved fluorescence properties of the HRP(desFe) adduct were examined; the multifrequency phase and modulation method was utilized for lifetime and dynamic polarization studies. The emission spectrum of HRP(desFe) had maxima at 633 and 696 nm. The lifetime of this emission was characterized by a single exponential decay of 16.87 ns at 22 degrees C. Debye rotational relaxation times for HRP(desFe) were determined using both static (Perrin plot) and dynamic (differential phase and modulation fluorometry) methods; these two approaches gave values of 96 and 86 ns, respectively. A spherical protein of HRP's molecular weight and partial specific volume would be expected to have a Debye rotational relaxation time, at 22 degrees C, in the range of 50 to 60 ns, depending upon the extent of hydration. Hence our results indicate that HRP(desFe) is asymmetric; the global rotational relaxation times observed are consistent with those of a prolate ellipsoid with an axial ratio of 3:1.  相似文献   

7.
Porphyromonas gingivalis acquires heme through an outer-membrane heme transporter HmuR and heme-binding hemophore-like lipoprotein HmuY. Here, we compare binding of iron(III) mesoporphyrin IX (mesoheme) and iron(III) deuteroporphyrin IX (deuteroheme) to HmuY with that of iron(III) protoporphyrin IX (protoheme) and protoporphyrin IX (PPIX) using spectroscopic methods. In contrast to PPIX, mesoheme and deuteroheme enter the HmuY heme cavity and are coordinated by His134 and His166 residues in a fully analogous way to protoheme binding. However, in the case of deuteroheme two forms of HmuY–iron porphyrin complex were observed differing by a 180° rotation of porphyrin about the α-γ-meso-carbon axis. Since the use of porphyrins either as active photosensitizers or in combination with antibiotics may have therapeutic value for controlling bacterial growth in vivo, it is important to compare the binding of heme derivatives to HmuY.  相似文献   

8.
Protoporphyrin IX is an activator of soluble guanylate cyclase (sGC), but its role as an endogenous regulator of vascular function through cGMP has not been previously reported. In this study we examined whether the heme precursor delta-aminolevulinic acid (ALA) could regulate vascular force through promoting protoporphyrin IX-elicited activation of sGC. Exposure of endothelium-denuded bovine pulmonary arteries (BPA) in organoid culture to increasing concentrations of the heme precursor ALA caused a concentration-dependent increase in BPA epifluorescence, consistent with increased tissue protoporphyrin IX levels, associated with decreased force generation to increasing concentrations of serotonin. The force-depressing actions of 0.1 mM ALA were associated with increased cGMP-associated vasodilator-stimulated phosphoprotein (VASP) phosphorylation and increased sGC activity in homogenates of BPA cultured with ALA. Increasing iron availability with 0.1 mM FeSO(4) inhibited the decrease in contraction to serotonin and increase in sGC activity caused by ALA, associated with decreased protoporphyrin IX and increased heme. Chelating endogenous iron with 0.1 mM deferoxamine increased the detection of protoporphyrin IX and force depressing activity of 10 microM ALA. The inhibition of sGC activation with the heme oxidant 10 muM 1H-[1,2,4]oxadiazolo[4,3-a]quinoxalin-1-one (ODQ) attenuated the force depressing actions of an NO donor without altering the actions of ALA. Thus control of endogenous formation of protoporphyrin IX from ALA by the availability of iron is potentially a novel physiological mechanism of controlling vascular function through regulating the activity of sGC.  相似文献   

9.
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   

10.
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   

11.
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   

12.
For quantitative estimation of coproporphyrin III and protoporphyrin IX from their mixture, a sensitive spectrofluorometric method was developed. At room temperature, coproporphyrin III fluoresces in neutral or alkaline pH at 622 nm having substantial fluorescence at 632 nm where protoporphyrin IX also fluoresces maximally. Similarly, protoporphyrin IX also has substantial fluorescence at 622 nm. Therefore, while estimating protoporphyrin IX (E400 F632) or coproporphyrin III (E400 F622) concentratton, it is essential to correct for the fluorescence due to coproporphyrin III at 632 nm and protoporphyrin IX at 622 nm. This was done by formulating equations from appropriate constants derived from pure samples of coproporphyrin III and protoporphyrin IX. As law as 1 pmole of coproporphyrin III or protoporphyrin IX could be estimated from their mixture by using the spectrofluorometric method.  相似文献   

13.
Jacobs JM  Jacobs NJ 《Plant physiology》1993,101(4):1181-1187
We have investigated the formation of porphyrin intermediates by isolated barley (Hordeum vulgare) plastids incubated for 40 min with the porphyrin precursor 5-aminolevulinate and in the presence and absence of a diphenylether herbicide that blocks protoporphyrinogen oxidase, the enzyme in chlorophyll and heme synthesis that oxidizes protoporphyrinogen IX to protoporphyrin IX. In the absence of herbicide, about 50% of the protoporphyrin IX formed was found in the extraplastidic medium, which was separated from intact plastids by centrifugation at the end of the incubation period. In contrast, uroporphyrinogen, an earlier intermediate, and magnesium protoporphyrin IX, a later intermediate, were located mainly within the plastid. When the incubation was carried out in the presence of a herbicide that inhibits protoporphyrinogen oxidase, protoporphyrin IX formation by the plastids was completely abolished, but large amounts of protoporphyrinogen accumulated in the extraplastidic medium. To detect extraplastidic protoporphyrinogen, it was necessary to first oxidize it to protoporphyrin IX with the use of a herbicide-resistant protoporphyrinogen oxidase enzyme present in Escherichia coli membranes. Protoporphyrinogen is not detected by some commonly used methods for porphyrin analysis unless it is first oxidized to protoporphyrin IX. Protoporphyrin IX and protoporphyrinogen found outside the plastid did not arise from plastid lysis, because the percentage of plastid lysis, measured with a stromal marker enzyme, was far less than the percentage of these porphyrins in the extraplastidic fraction. These findings suggest that of the tetrapyrrolic intermediates synthesized by the plastids, protoporphyrinogen and protoporphyrin IX, are the most likely to be exported from the plastid to the cytoplasm. These results help explain the extraplastidic accumulation of protoporphyrin IX in plants treated with photobleaching herbicides. In addition, these findings suggest that plastids may export protoporphyrinogen or protoporphyrin IX for mitochondrial heme synthesis.  相似文献   

14.
Unlike pathogenic fungi, the budding yeast Saccharomyces cerevisiae is not efficient at using heme as a nutritional source of iron. Here we report that for this yeast, heme uptake is induced under conditions of heme starvation. Heme synthesis requires oxygen, and yeast grown anaerobically exhibited an increased uptake of hemin. Similarly, a strain lacking aminolevulinate synthase exhibited a sixfold increase in hemin uptake when grown without 2-aminolevulinic acid. We used microarray analysis of cells grown under reduced oxygen tension or reduced intracellular heme conditions to identify candidate genes involved in heme uptake. Surprisingly, overexpression of PUG1 (protoporphyrin uptake gene 1) resulted in reduced utilization of exogenous heme by a heme-deficient strain and, conversely, increased the utilization of protoporphyrin IX. Pug1p was localized to the plasma membrane by indirect immunofluorescence and subcellular fractionation. Strains overexpressing PUG1 exhibited decreased accumulation of [(55)Fe]hemin but increased accumulation of protoporphyrin IX compared to the wild-type strain. To measure the effect of PUG1 overexpression on intracellular heme pools, we used a CYC1-lacZ reporter, which is activated in the presence of heme, and we monitored the activity of a heme-containing metalloreductase, Fre1p, expressed from a constitutive promoter. The data from these experiments were consistent with a role for Pug1p in inducible protoporphyrin IX influx and heme efflux.  相似文献   

15.
A continuous spectrofluorimetric assay for determining ferrochelatase activity has been developed using the physiological substrates ferrous iron and protoporphyrin IX under strictly anaerobic conditions. In contrast to heme, the product of the ferrochelatase-catalyzed reaction, protoporphyrin IX is fluorescent, and therefore the progress of the reaction can be monitored by following the decrease in protoporphyrin fluorescence intensity (with excitation and emission wavelengths at 505 and 635 nm, respectively). This continuous fluorimetric assay detects activities as low as 0.01 nmol porphyrin consumed min(-1), representing an increase in sensitivity of up to two orders of magnitude over the currently used, discontinuous assays. The determination of the steady-state kinetic parameters of ferrochelatase yielded K(m)(PPIX)=1.4+/-0.2 microM, K(m)(Fe(2+))=1.9+/-0.3 microM, and k(cat)=4.0+/-0.3 min(-1). In addition to its applicability for acquisition of kinetic data to characterize ferrochelatase and recombinant variants, this new method should permit detection of low concentrations of ferrochelatase in biological samples.  相似文献   

16.
Horseradish apoperoxidase (apoHRP) was reconstituted with various porphyrin derivatives, e.g., ferric, cupric, manganese, and zinc protoporphyrin IX, metal-free protoporphyrin IX, hematoporphyrin IX and deuteroporphyrin IX. The visible absorption spectra of these porphyrin-apoHRP complexes were examined. The time required for maximum development of the new Soret peak after reconstitution was used to measure the rate of porphyrin-apoHRP reconstitution. All of the four metal-protoporphyrins reconstituted with apoHRP at the same rate as metal-free protoporphyrin IX, whereas, for the metal-free porphyrins, the rates of reconstitution were in the order of deuteroporphyrin IX > hematoporphyrin IX > protoporphyrin IX. The porphyrins on the reconstituted porphyrin-apoHRP complexes were used as localized photosensitizers for photodynamic studies. No amino acid residues were oxidized on illumination of the ferric, cupric and manganese protoporphyrin IX-apoHRP complexes due to the paramagnetic properties of these metal ions. With diamagnetic zinc ion, two histidine and one methionine residues were oxidized which was the same as in the protoporphyrin IX- and hematoporphyrin IX-apoHRP complexes. However, only one histidine was destroyed on illumination of the deuteroporphyrin IX-apoHRP complex. The results confirmed the resistance of horseradish peroxidase to photodynamic action and suggested the involvement of at least one histidine residue in the heme environment of horseradish peroxidase.  相似文献   

17.
Heme binding to murine erythroleukemia cells. Evidence for a heme receptor   总被引:7,自引:0,他引:7  
Friend virus transformed murine erythroleukemia (MEL) cells are known to take up heme from the surrounding medium and to incorporate it into newly synthesized hemoglobin (Granick, J. L., and Sassa, S. (1978) J. Biol. Chem. 253, 5402-5406), but the mechanism of its uptake is unknown. We hypothesized the existence of a specific receptor for heme in the plasma membrane. Using [55Fe]heme, we examined the characteristics of its interaction with MEL cells at 4 degrees C. [55Fe]heme binding reached equilibrium within 4 h, was 80% dissociable by 16 h, and was independent of pH over the range 7.0-8.2. Specific heme binding was linear with cell number, and competitive binding studies with various heme analogues, such as free protoporphyrin IX, metal-substituted protoporphyrin IX, Fe-mesoporphyrin IX, and Fe-deuteroporphyrin IX, revealed significant stereospecificity for Fe-protoporphyrin IX. The dissociation constant of the interaction was 0.03 nM-1 with no evidence of cooperativity or multiple classes of sites. The average number of sites/cell was approximately 10,300. Reduction of binding following preincubation with trypsin, in conjunction with the above data, suggests that this cell type may display a receptor for heme which is comprised, as least in part, of protein.  相似文献   

18.
Achieving a maximal safe extent of resection during brain tumor surgery is the goal for improved patient prognosis. Fluorescence‐guided neurosurgery using 5‐aminolevulinic acid (5‐ALA) induced protoporphyrin IX has thereby become a valuable tool enabling a high frequency of complete resections and a prolonged progression‐free survival in glioblastoma patients. We present a widefield fluorescence lifetime imaging device with 250 mm working distance, working under similar conditions such as surgical microscopes based on a time‐of‐flight dual tap CMOS camera. In contrast to intensity‐based fluorescence imaging, our method is invariant to light scattering and absorption while being sensitive to the molecular composition of the tissue. We evaluate the feasibility of lifetime imaging of protoporphyrin IX using our system to analyze brain tumor phantoms and fresh 5‐ALA‐labeled human tissue samples. The results demonstrate the potential of our lifetime sensing device to go beyond the limitation of current intensity‐based fluorescence‐guided neurosurgery.   相似文献   

19.
Mouse Friend virus-transformed erythroleukemia cells in culture undergo erythroid differentiation when treated with a variety of compounds including iron protoporphyrin IX, i.e. hemin. Exogenous hemin is not only incorporated into hemoglobin in these cells but also stimulates heme biosynthesis (Granick, J. L., and Sassa, S. (1978) J. Biol. Chem. 253, 5402-5406). In this study, we examined whether metalloporphyrins other than hemin can also induce differentiation, and if so, whether they can also be incorporated into hemoglobin. Among eight metalloporphyrins examined in culture of these cells, i.e. Co, Mn, Cu, Mg, Ni, Zn, Sn, and Cd protoporphyrin IX, only Co protoporphyrin (10(-4) M) was found to significantly increase the biosynthesis of heme and hemoglobin. In contrast to hemin-mediated induction of erythroid differentiation, Co protoporphyrin was not incorporated into hemoglobin in Friend cells. These data indicate that Co protoporphyrin induces the formation of heme and hemoglobin in Friend cells and that these increases are due to the enhancement of heme biosynthetic activity.  相似文献   

20.
We describe a fluorometric assay for heme synthetase, the enzyme that is genetically deficient in erythropoietic protoporphyria. The method, which can readily detect activity in 1 microliter of packed human lymphocytes, is based on the formation of zinc protoheme from protoporphyrin IX. That zinc chelatase and ferrochelatase activities reside in the same enzyme was shown by the competitive action of ferrous ions and the inhibitory effects of N-methyl protoporphyrin (a specific inhibitor of heme synthetase) on zinc chelatase. The Km for zinc was 11 micrograms and that for protoporphyrin IX was 6 microM. The Ki fro ferrous ions was 14 microM. Zinc chelatase was reduced to 15.3% of the mean control activity in lymphocytes obtained from patients with protoporphyria, thus confirming the defect of heme biosynthesis in this disorder. The assay should prove to be useful for determining heme synthetase in tissues with low specific activity and to investigate further the enzymatic defect in protoporphyria.  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