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.
The effects of hydrophobicity and charges of dicarboxylic porphyrins upon their interactions with membrane model systems are investigated. Four protonation steps are evidenced from fluorescence emission studies of hematoporphyrin IX and its more hydrophobic parent compound lacking of alcoholic chain, deuteroporphyrin IX. They are attributed to the successive protonations of the inner nitrogens of the porphyrin cycle (pK = 4.7 and 2.9 for hematoporphyrin and 4.4 and 2.7 for deuteroporphyrin) and successive deprotonations of propionic groups (pK ≈ 5.0 and 5.5 for hematoporphyrin and 5.4 and 6.0 for deuteroporphyrin). These porphyrins, as well as their dimethyl ester forms, are shown to incorporate as monomers into the hydrophobic bilayer of egg phosphatidylcholine small unilamellar vesicles, although the esterified forms are highly aggregated in aqueous solutions. In the case of the non-esterified forms, the incorporation of the porphyrins into the lipidic bilayer is reversible and strongly pH-dependent. A theoretical model is presented which takes into account the various protonation steps and the partition equilibria of the porphyrin between the vesicle lipidic phase and the water medium. The neutral form of the porphyrin (i.e., carboxylic groups protonated) presents the higher affinity, with constants of K ≈ 2 · 105 and K ≈ 6 · 106 M−1 (relative to lipid concentration) for hematoporphyrin and deuteroporphyrin, respectively. Protonation of one inner nitrogen leading to the monocationic form is sufficient to prevent incorporation into the hydrophobic bilayer. On the other hand, deprotonation of the peripheral propionic chains leading to anionic forms is less effective. These interactions between vesicles and porphyrins lead to shifts of the apparent pK of nitrogens and carboxylic groups, the latter one being now in the range of physiological pH. These results are discussed with regards to the hypothesis of a possible role of pH in the preferential uptake of porphyrins by tumors.  相似文献   

2.
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   

3.
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   

4.
Steady-state and stopped-flow measurements of the absorbance and fluorescence of aqueous solutions were performed to characterize the pH-dependent ionization and aggregation states of deuteroporphyrin. Porphyrin self-association promoted by neutralization of the carboxylic groups takes place within a few milliseconds impeding characterization of the monomer ionization states. Extrapolation at infinite dilution of the values obtained from steady-state measurements yielded the pKs of the carboxylic groups (6.6, 5.3) and inner nitrogens (4.1, 2.3). The kinetics of interactions of the porphyrin with unilamellar fluid state dioleoylphosphatidylcholine vesicles was examined in a large pH range, with focus on the entry step. From alkaline pH to a value of 6.5, the entrance rate is maximal (1.69 x 10(6) M(-1) s(-1) versus phospholipid concentration). It decreases to 2.07 x 10(5) M(-1) s(-1) at lower pH with an apparent pK of 5.39. This effect appears to be related to the formation of porphyrin dimer rather than to the protonation of inner nitrogen. In keeping with previous data, these results support the concept of a pH-mediated selectivity of carboxylic porphyrins for tumor. They also indicate that the propensity of these molecules to self-associate at low pH could yield to some retention in acidic intracellular vesicles of the endosome/lysosome compartment.  相似文献   

5.
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   

6.
We describe here a sensitive method for the purification and analysis of porphyrins present in hematoporphyrin derivative. Hematoporphyrin derivative is a solution containing a complex mixture of dicarboxylic porphyrins such as hematoporphyrin IX, monohydroxyethyl monovinyl deuteroporphyrin isomers, and protoporphyrin IX in addition to porphyrin aggregates of variable molecular sizes. This mixture is known for its ability to be selectively retained by tumor cells and for its cytotoxicity in the presence of light. In order to study the mechanisms of hematoporphyrin derivative uptake and its cellular metabolism, extraction methods are required that combine high recoveries with minimum changes of very labile components. Extraction with perchloric acid: methanol mixtures recovered only some 60% of the porphyrins taken up by tumor cells and artifactual fluorescent spots were seen on thin-layer chromatograms. Improved yields were obtained upon extraction with dimethyl sulfoxide or Triton X-100:4-(2-hydroxyethyl)-1-piperazineethanesulfonic acid (Hepes) buffer mixture, but the extracts were not suitable for reverse-phase thin-layer chromatography (RTLC). The procedure described here consists of extracting porphyrins from cultured tumor cells with a buffered detergent followed by sequential chromatography on DEAE-cellulose columns and on reverse-phase octadecylsilyl cartridges. Identification of the isolated free dicarboxylic porphyrins is conveniently done by RTLC.  相似文献   

7.
Steady-state and stopped-flow measurements of the absorbance and fluorescence of aqueous solutions were performed to characterize the pH-dependent ionization and aggregation states of deuteroporphyrin. Porphyrin self-association promoted by neutralization of the carboxylic groups takes place within a few milliseconds impeding characterization of the monomer ionization states. Extrapolation at infinite dilution of the values obtained from steady-state measurements yielded the pKs of the carboxylic groups (6.6, 5.3) and inner nitrogens (4.1, 2.3). The kinetics of interactions of the porphyrin with unilamellar fluid state dioleoylphosphatidylcholine vesicles was examined in a large pH range, with focus on the entry step. From alkaline pH to a value of 6.5, the entrance rate is maximal (1.69×106 M−1 s−1 versus phospholipid concentration). It decreases to 2.07×105 M−1 s−1 at lower pH with an apparent pK of 5.39. This effect appears to be related to the formation of porphyrin dimer rather than to the protonation of inner nitrogen. In keeping with previous data, these results support the concept of a pH-mediated selectivity of carboxylic porphyrins for tumor. They also indicate that the propensity of these molecules to self-associate at low pH could yield to some retention in acidic intracellular vesicles of the endosome/lysosome compartment.  相似文献   

8.
The influence of various porphyrins (deuteroporphyrin IX, mesoporphyrin IX, protoporphyrin IX, hematoporphyrin) and two related compounds (hemin, biliverdin) on the spontaneous proliferation of mouse spleen lymphocytes has been estimated in vitro by the 3H-thymidine uptake assay. It has been found that porphyrins (endogenous ligands for the mitochondrial benzodiazepine receptor) produce a concentration-dependent inhibition of 3H-thymidine incorporation into the DNA of these cells. Metalloporphyrin-hemin has been observed to evoke a weak inhibitory effect, in a high concentration (10(-4)M), whereas biliverdin, a porphyrins degradation product, was inactive in the same experimental conditions. Those findings indicate that endogenous porphyrins, presumably acting through the mitochondrial benzodiazepine receptor, could regulate the proliferation of mouse spleen lymphocytes in vitro.  相似文献   

9.
Three porphyrin systems have been characterised for use in two-photon fluorescence imaging of biological samples. We have determined the two-photon absorption cross sections (sigma(2)) of the di-cation, free-base and metallated forms of hematoporphyrin derivative (HpD), hematoporphyrin IX (Hp9) and a boronated protoporphyrin (BOPP) using the open-aperture Z-scan and the two-photon induced fluorescence (TPIF) techniques at an excitation wavelength of 800 nm. The insertion of either protons or a metal ion into the macrocycle is shown not to significantly influence the sigma(2) of the porphyrins. Two-photon time-resolved fluorescence images of C6 glioma cells transfected with a free-base form of the BOPP have been obtained as a function of the porphyrin concentration. These studies reveal a maximum useful porphyrin concentration for fluorescence imaging purposes of approximately 30 microg mL(-1).  相似文献   

10.
We have already reported in Balb C mouse transplantable mammary carcinoma, that uroporphyrin I and III are superior as tumour localizers when compared to hematoporphyrin derivative and a derivative thereof, photofrin II. This study compares the binding of porphyrins to proteins which may be found in tumour cells or stroma to investigate whether there is a common binding determinant. Coproporphyrin III and deuteroporphyrin IX which are non-tumour localizing porphyrins, were also part of the comparative study. The interaction of these porphyrins with acid soluble collagen and acid insoluble collagen, elastin, and fibrin was evaluated, and the binding of uroporphyrin isomers I and III and deuteroporphyrin IX to gelatin and fibrinogen, was also determined. The results suggest that collagen, especially the acid soluble form, and gelatin preferentially bind the four porphyrins which localize in mammary carcinoma tissue. The well reported observations that malignant epithelial cells, including breast cancer, produce collagen and contain a rate-limiting enzyme in collagen biosynthesis would support the notion that de novo synthesis of this protein may in part govern the tumour uptake and retention of porphyrins. Elastin, fibrinogen and fibrin showed non-discriminant binding to the porphyrins under study.  相似文献   

11.
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   

12.
A comparison of the photosensitizing ability of a variety of porphyrins for photohaemolysis gives the following order of activity: protoporphyrin greater than deuteroporphyrin, mesoporphyrin, haematoporphyrin dimethyl ester much greater than haematoporphyrin diacetate, haematoporphyrin greater than haematoporphyrin monoacetate, coproporphyrin III, haematoporphyrin derivative, coproporphyrin III tetramethyl ester greater than uroporphyrin I, meso-tetra-(N-methyl-4-pyridinium)porphyrin tetratoluene-p-sulphonate, meso-tetra-(p-carboxyphenyl)porphyrin, protoporphyrin dimethyl ester, meso-tetra-(p-hydroxy-sulphonylphenyl)porphyrin tetrasodium salt, uroporphyrin III, deuteroporphyrin-3,8-disulphonic acid and protohaemin. The results for the metal-free porphyrins are rationalized in terms of solubility and partition properties, and a model is proposed for the incorporation of amphipathic porphyrins into the membrane lipid bilayer. Experiments with erythrocytes from patients with erythropoeitic protoporphyria and with normal erythrocytes to which porphyrin was added in a deuterium oxide medium do not lead to an increase in the rate of photohaemolysis. A possible explanation for this somewhat surprising observation is outlined.  相似文献   

13.
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   

14.
The photosensitized luminescence of singlet molecular oxygen has been studied in aqueous and alcoholic solutions of hematoporphyrin IX (HP) and di- and oligomeric components of "hematoporphyrin derivative" (photofrin II) which is known to be used as a drug in photodynamic tumor therapy. The quantum yields of 1O2 generation (gamma delta) by these compounds have been determined. It was found that the highest gamma delta values are characteristic of alcoholic and micellar detergent aqueous solutions. In detergent-free aqueous solutions containing mainly associated porphyrin molecules, gamma delta is much lower (5-30%), polymeric photofrin components being considerably less active than HP. Both localization of porphyrins in hydrophobic loci and high photosensitizing activity in lipid phase are supposed to play the key role in tumor photodestruction.  相似文献   

15.
Previous genetic and biochemical studies have confirmed that hemoglobin and hemin utilization in Porphyromonas gingivalis is mediated by the outer membrane hemoglobin and heme receptor HmuR, as well as gingipain K (Kgp), a lysine-specific cysteine protease, and gingipain R1 (HRgpA), one of two arginine-specific cysteine proteases. In this study we report on the binding specificity of the recombinant P. gingivalis HmuR protein and native gingipains for hemoglobin, hemin, various porphyrins, and metalloporphyrins as assessed by spectrophotometric assays, by affinity chromatography, and by enzyme-linked immunosorbent assay. Protoporphyrin, mesoporphyrin, deuteroporphyrin, hematoporphyrin, and some of their iron, copper, and zinc derivatives were examined to evaluate the role of both the central metal ion and the peripheral substituents on binding to recombinant HmuR and soluble gingipains. Scatchard analysis of hemin binding to Escherichia coli cells expressing recombinant membrane-associated six-His-tagged HmuR yielded a linear plot with a binding affinity of 2.4 x 10(-5) M. Recombinant E. coli cells bound the iron, copper, and zinc derivatives of protoporphyrin IX (PPIX) with similar affinities, and approximately four times more tightly than PPIX itself, which suggests that the active site of HmuR contains a histidine that binds the metal ion in the porphyrin ring. Furthermore, we found that recombinant HmuR prefers the ethyl and vinyl side chains of the PPIX molecule to either the larger hydroxyethyl or smaller hydrogen side chains. Kgp and HRgpA were demonstrated to bind various porphyrins and metalloporphyrins with affinities similar to those for hemin, indicating that the binding of Kgp and HRgpA to these porphyrins does not require a metal within the porphyrin ring. We did not detect the binding of RgpB, the arginine-specific cysteine protease that lacks a C-terminal hemagglutinin domain, to hemoglobin, porphyrins, or metalloporphyrins. Kgp and HRgpA, but not RgpB, were demonstrated to bind directly to soluble recombinant six-His-tagged HmuR. Several possible mechanisms for the cooperation between outer membrane receptor HmuR and proteases Kgp and HRgpA in hemin and hemoglobin binding and utilization are discussed.  相似文献   

16.
The dimerization equilibrium of deuteroporphyrin IX and of mesoporphyrin IX in aqueous solutions were studied by fluorimetric techniques over the 0.01-1 microM concentration range, where dimerization is the dominant aggregation process. Deuteroporphyrin IX was studied at several temperatures over the range 22-37 degrees C, and mesoporphyrin at 25 and 37 degrees C. The magnitudes determined for the dimerization equilibrium constants (25 degrees C, neutral pH, phosphate-buffered saline) are 2.3 X 10(6)M-1 and 5.4 X 10(6)M-1 for the deutero and meso derivatives respectively. The meso, deutero and haemato species tested show a similar temperature effect, namely dimerization decreasing with increasing temperature, indicating the involvement of a negative enthalpy change. Van''t Hoff isochore of the dimerization constants determined for deuteroporphyrin IX was linear within the temperature range of 22-37 degrees C, allowing the calculation of the thermodynamic parameters. For deuteroporphyrin dimerization, those were found to be delta G0 = -36. 4kJ X mol-1; delta H0 = -46. 0kJ X mol-1 and delta S0 = -32.2J X K-1 X mol-1 (at neutral pH, 25 degrees C, phosphate-buffered saline), showing the process to be enthalpy-driven. Similar trends have been found for porphyrin species other than those studied here. Our data fit with a hypothesis giving a major role to the solvent in driving porphyrins to aggregate in aqueous solution. The magnitudes and directions of the energetic changes fit better with the expectation of the '' solvophobic force'' theory predicting enthalpy-driven association, than with the classic hydrophobic bonding, predicting the association to be entropy-driven.  相似文献   

17.
Specificity of the heme requirement for growth of Bacteroides ruminicola   总被引:15,自引:6,他引:9  
Caldwell, D. R. (U.S. Department of Agriculture, Beltsville, Md.), D. C. White, M. P. Bryant, and R. N. Doetsch. Specificity of the heme requirement for growth of Bacteroides ruminicola. J. Bacteriol. 90:1645-1654. 1965.-Previous studies suggested that most strains of Bacteroides ruminicola subsp. ruminicola require heme for growth. Present studies with heme-requiring strain 23 showed that protoheme was replaced by various porphyrins, uroporphyrinogen, coproporphyrinogen, certain iron-free metalloporphyrins, hemes, and certain heme-proteins containing readily removable hemes. Strain 23 utilized a wider range of tetrapyrroles than hemin-requiring bacteria previously studied. Inactive compounds included porphyrin biosynthesis intermediates preceding the tetrapyrrole stage and related compounds; uroporphyrin, chlorophyll, pheophytin, phycoerythrin, bilirubin, pyrrole, FeSO(4) with or without chelating agents; and representative ferrichrome compounds. Strain 23, two other strains representing predominant biotypes of B. ruminicola subsp. ruminicola, and one closely related strain grew in media containing heme-free protoporphyrin, mesoporphyrin, hematoporphyrin, or deuteroporphyrin, apparently inserting iron into several nonvinyl porphyrins. Porphobilinogen and porphyrin synthesis, apparently via the commonly known heme synthesis pathway, occurred during growth of heme-independent B. ruminicola subsp. brevis strain GA33 in a tetrapyrrole-free medium containing delta-aminolevulinic acid, but delta-aminolevulinic acid metabolism to porphobilinogen or porphyrins could not be detected in cells of heme-requiring strain 23 grown in the same medium with hemin added. Growth of strain 23 with uroporphyrinogen, coproporphyrinogen, or protoporphyrin IX replacing hemin suggests that part of the commonly known heme-biosynthesis pathway is present in this strain, but nutritional and metabolic evidence indicates that some or all of the enzymes synthesizing the tetrapyrrole nucleus from linear molecules are lacking or inactive.  相似文献   

18.
Crystallographic studies of L-chain horse spleen apoferritin (HSF) co-crystallized with Pt-hematoporphyrin IX and Sn-protoporphyrin IX have brought significant new insights into structure-function relationships in ferritins. Interactions of HSF with porphyrins are discussed. Structural results show that the nestling properties into HSF are dependent on the porphyrin moiety. (Only protoporphyrin IX significantly interacts with the protein, whereas hematoporphyrin IX does not.) These studies additionally point out the L-chain HSF ability to demetalate metalloporphyrins, a result which is of importance in looking at the iron storage properties of ferritins. In both compound investigated (whether the porphyrin reaches the binding site or not), the complexation appears to be concomitant with the extraction of the metal from the porphyrin. To analyze further the previous results, a three-dimensional alignment of ferritin sequences based on available crystallographic coordinates, including the present structures, is given. It confirms a high degree of homology between these members of the ferritin family and thus allows us to emphasize observed structural differences: 1) unlike L-chain HSF, H-chain human ferritin presents no preformed binding site; and 2) despite the absence of axial ligands, and due to the demetalation, L-chain HSF is able to host protoporphyrin at a similar location to that naturally found in bacterioferritin.  相似文献   

19.
1. Iron protoporphyrin IX was required for the growth of H. influenzae. It could be replaced by protoporphyrin IX. When grown on protoporphyrin evidence was obtained for the presence of Fe porphyrin in the organism. It was concluded that the organism could insert iron into the protoporphyrin ring. 2. In the smooth strains, other porphyrins containing no iron such as deutero-, hemato-, meso-, and coproporphyrins could not replace protoporphyrin for growth. Since protoporphyrin has two vinyl groups which other porphyrins lack, it was concluded that the two vinyl groups were essential for growth. 3. When porphyrins lacking vinyl groups were converted chemically into iron porphyrins and then supplied to the organisms it was found that these iron porphyrins supported growth. It was concluded that the "smooth" organisms were able to insert iron only into the porphyrin containing the vinyl groups; i.e., protoporphyrin. One function of the vinyl groups then was to permit iron to be inserted biologically into the porphyrin ring. 4. An anomalous behavior in the rough Turner strain was observed and discussed. This organism was able to insert iron into mesoporphyrin at low concentrations but was inhibited by this compound at higher concentrations. In all other reactions with the porphyrins this rough strain behaved in the same was as did the smooth strains. 5. All strains which were grown on iron porphyrins lacking vinyl groups could not reduce nitrate to nitrite. When grown on protoporphyrin or Fe protoporphyrin reduction of nitrate occurred. It was concluded that the nitrate-reducing mechanism required the presence of the vinyl groups either for its formation or function. 6. The porphyrins lacking iron and lacking vinyl groups inhibited the growth of H. influenzae on Fe protoporphyrin. The inhibition between a porphyrin and Fe protoporphyrin was a competitive one. It was suggested that the porphyrin inhibited the growth-promoting properties of Fe protoporphyrin by attaching on to a particular apoprotein, thus preventing the formation of a heme catalyst. Likewise, competition between two growth-promoting Fe porphyrins for apoenzymes could be shown to occur. 7. Protoporphyrin and Fe protoporphyrin supported growth. When their propionic acid side chains were esterified they no longer supported growth. It was suggested that the esterified carboxyl groups could not attach to the specific apoproteins to form the heme enzymes and so could not act to support growth. For the same reason the inhibitory action of porphyrins lacking vinyl groups could be prevented by esterifying their propionic acid groups.  相似文献   

20.
The heme biosynthetic pathway culminates with the ferrochelatase-catalyzed ferrous iron chelation into protoporphyrin IX to form protoheme. The catalytic mechanism of ferrochelatase has been proposed to involve the stabilization of a nonplanar porphyrin to present the pyrrole nitrogens to the metal ion substrate. Previously, we hypothesized that the ferrochelatase-induced nonplanar distortions of the porphyrin substrate impose selectivity for the divalent metal ion incorporated into the porphyrin ring and facilitate the release of the metalated porphyrin through its reduced affinity for the enzyme. Using resonance Raman spectroscopy, the structural properties of porphyrins bound to the active site of directly evolved Ni(2+)-chelatase variants are now examined with regard to the mode and extent of porphyrin deformation and related to the catalytic properties of the enzymes. The Ni(2+)-chelatase variants (S143T, F323L, and S143T/F323L), which were directly evolved to exhibit an enhanced Ni(2+)-chelatase activity over that of the parent wild-type ferrochelatase, induced a weaker saddling deformation of the porphyrin substrate. Steady-state kinetic parameters of the evolved variants for Ni(2+)- and Fe(2+)-chelatase activities increased compared to those of wild-type ferrochelatase. In particular, the reduced porphyrin saddling deformation correlated with increased catalytic efficiency toward the metal ion substrate (Ni(2+) or Fe(2+)). The results lead us to propose that the decrease in the induced protoporphyrin IX saddling mode is associated with a less stringent metal ion preference by ferrochelatase and a slower porphyrin chelation step.  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