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 oxidation of ibuprofen with H2O2 catalysed by Horseradish peroxidase (HRP), Cl8TPPS4Fe(III)(OH2)2 and Cl8TPPS4Mn(III)(OH2)2 in AOT reverse micelles gives 2-(4'-isobutyl-phenyl)ethanol (5) and p-isobutyl acetophenone (6) in moderate yields. The reaction of ibuprofen (2) with H2O2 catalysed by HRP form carbon radicals by the oxidative decarboxylation, which on reaction with molecular oxygen to form hydroperoxy intermediate, responsible for the formation of the products 5 and 6. The yields of different oxidation products depend on the pH, the water to surfactant ratio (Wo), concentration of Cl8TPPS4Fe(III)(OH2)2 and Cl8TPPS4Mn(III)(OH2)2 and amount of molecular oxygen present in AOT reverse micelles. The formation of 2-(4'-isobutyl phenyl)ethanol (5) may be explained by the hydrogen abstraction from ibuprofen by high valent oxo-manganese(IV) radical cation, followed by decarboxylation and subsequent recombination of either free hydroxy radical or hydroxy iron(III)/manganese(III) porphyrins. The over-oxidation of 5 with high valent oxo-manganese, Mn(IV)radical cation intermediate form 6 in AOT reverse micelles by abstraction and recombination mechanism.  相似文献   

2.
The efficiencies of several metalloporphyrin complexes at catalyzing the disproportionation of superoxide have been determined at pH 10 in both carbonate and borate buffer systems. Catalytic rate constants were obtained for the iron(III) and cobalt(III) derivatives of tetrakis(4-N-methylpyridyl) porphine, for tetraphenylporphinesulfonatoferrate(III) and for hemin. In addition, the effects of added bovine serum albumin and imidazole were studied. The order of catalytic efficiency is FeTMpyP greater than FeTMpyP(Im)2 greater than FeTPPS(Im)2 approximately FeTPPS approximately FeTPPS.BSA approximately Fe(EDTA) greater than or approximately CoTMpyP greater than hemin(Im)2 greater than or approximately hemin.  相似文献   

3.
Zhao L  Ma R  Li J  Li Y  An Y  Shi L 《Biomacromolecules》2008,9(10):2601-2608
Micellization of poly(ethylene glycol)-block-poly(4-vinylpyridine) (PEG114-b-P4VP61) induced by 5,10,15,20-tetrakis-(4-sulfonatophenyl)-porphyrin (TPPS) in acidic solutions were studied by dynamic and static light scattering, atomic force microscope, and UV-vis spectroscopy. The resultant complex micelles had a core-shell structure with the electrostatically complex TPPS/P4VP as the core and the soluble PEG as the shell. The anionic TPPS in the micellar core formed J-aggregates at pH 1.5-2.5 and H-aggregates at pH 3.0-4.0, respectively. Interconversion between the J-aggregates and the H-aggregates was carried out by adjusting the pH value of the micelle solutions. It is worth noting that the micelles showed strong split Cotton effect in the circular dichroism spectra although TPPS and the copolymer were all achiral. The resulting chirality sign could be selected by the hydrodynamic forces of a stirring vortex. Positive or negative chiral signals appeared when stirring clockwise or anticlockwise.  相似文献   

4.
The equilibrium behavior of cationic iron(III) meso-tetrakis(4-N-methyl-pyridiniumyl) porphyrin, Fe(III)TMPyP, in aqueous solution was studied as a function of pH by optical absorption, EPR and (1)H NMR spectroscopies. The presence of several Fe(III)TMPyP species in solution was unequivocally demonstrated: monomeric porphyrin species (a monoaqueous five-coordinated complex, a diaaqueous six-coordinated complex and a monoaqueous-hydroxo six-coordinated complex), a micro-oxo dimer and a bis-hydroxo complex. The addition of salt to the porphyrin solution leads to a simplification of the equilibrium as a function of pH. In this case, only three species were observed in solution: a monomeric porphyrin species, a micro-oxo dimer and a bis-hydroxo complex. Optical absorption, EPR and (1)H NMR spectra contributed to the characterization of these species. Four critical pH values (pK) for Fe(III)TMPyP were obtained in pure buffer and only three pK values were observed in the presence of NaCl. The addition of salt favors the presence of the dimeric species in solution and simplifies the equilibrium in the acidic pH range.  相似文献   

5.
The aggregate morphology of meso-tetrakis(4-sulfonatophenyl) porphyrin (TPPS(4)) in aqueous solution is investigated by using small angle x-ray scattering (SAXS) technique. Measurements were performed at pH 4.0 and 9.0 to monitor the pH influence on the structural parameters of the aggregates. Radii of gyration were obtained from distance distribution functions p(r) analysis. The experimental data of TPPS(4) at pH 4.0 showed well-defined oscillations characteristic of large aggregates in contrast to the SAXS curve of 5 mM TPPS(4) at pH 9.0, where both a significant decrease in the intensity and the disappearance of the oscillation peaks suggest the dissociation of the aggregate. A 340-A long "hollow" cylinder with shell thickness of 20 A, compatible to the porphyrin molecule dimension, represents well the scattering curve of the aggregates at pH 4.0. According to the fitting parameters, 26 porphyrin molecules self-associate into a ringlike configuration in the plane of the cylinder cross-section. The total number of porphyrin molecules in the whole aggregate was also estimated as approximately 3000. The model compatible to SAXS data of a hollow cylinder with J-aggregation in the cross-section and H-aggregation (columnar stacking) between the cylinder layers is consistent with optical absorption spectroscopic data both in the literature and obtained in this work.  相似文献   

6.
The aggregation of meso-tetra(4-sulfonatophenyl)porphyrin (H(2)TPPS(4-)) in phosphate solutions was investigated as a function of pH, concentration, time, ionic strength, and solution preparation (either from dilution of a freshly prepared 2 mM stock or by direct preparation of μM solution concentrations) using a combination of complementary analytical techniques. UV-vis and fluorescence spectroscopy indicated the formation of staggered, side-by-side (J-type) assemblies. Their size and self-associative behavior were determined using analytical ultracentrifugation and small-angle X-ray scattering. Our results indicate that in neutral and basic solutions of H(2)TPPS(4-), porphyrin dimers and trimers are formed at micromolar concentrations and in the absence of NaCl to screen any ionic interactions. At these low concentrations and pH 4, the protonated H(4)TPPS(2-) species self-assembles, leading to the formation of particularly stable aggregates bearing 25 ± 3 macrocycles. At higher concentrations, these structures further organize or reorganize into tubular, rod-like shapes of various lengths, which were imaged by cryogenic and freeze-fracture transmission electron microscopy. Micron-scale fibrillar aggregates were obtained even at micromolar concentrations at pH 4 when prepared from dilution of a 2 mM stock solution, upon addition of NaCl, or both.  相似文献   

7.
The nitrosylation of two water-soluble iron-porphyrins, the anionic Fe(III)-meso-tetrakis(p-sulfonatophenyl)porphyrin (FeTPPS(4)) and the cationic Fe(III)-meso-tetrakis(4-N-methylpyridiniumyl)porphyrin (FeTMPyP), by the nitric oxide donor S-nitroso-N-acetylpenicillamine (SNAP) was studied using optical absorption spectroscopy. The influence of ionic and non-ionic micelles on rates of nitric oxide transfer was investigated. Initially, the effect of the micelles on the pH-dependent equilibrium between monomeric and micro-oxo-dimeric species of the iron-porphyrins was examined. It is not affected in micelle-porphyrin systems with electric charges identical in sign. The non-ionic micelles of polidocanol induce a small negative pK shift. In contrast, the dimerization equilibrium of porphyrins in oppositely charged micellar phases is displaced to lower pH by approximately 2 units. Nitric oxide binding to monomers and micro-oxo-dimers was examined at pH 5.0 and 8.0, respectively. Contrary to nitrosylation by NO gas, SNAP induces reductive nitrosylation. There was no observed NO-Fe(III)porphyrin intermediate. Nitrosylation rates were obtained and compared in aqueous buffer and in micellar systems. Monomers nitrosylate much faster than micro-oxo-dimers. Oppositely charged micelles prevent nitrosylation of the iron-porphyrins or considerably enhance nitrosylation times. Nitrosylation rates are comparable to transnitrosylation rates between several S-nitrosothiols and thiol-containing proteins, suggesting biological relevance for the process.  相似文献   

8.
Radical production during the photolysis of deaerated aqueous alkaline solutions (pH 11) of some water-soluble porphyrins was investigated. Metal-free and metallo complexes of tetrakis (4-N-methylpyridyl)porphyrin (TMPyP) and tetra (4-sulphonatophenyl)porphyrin (TPPS4) were studied. Evidence for the formation of OH radicals during photolysis at 615, 545, 435, 408 and 335 nm of Fe(III) TPPS4 is presented. Fe(III) TMPyP, Mn(III) TPPS4 and Mn(III) TMPyP also gave OH radicals but only during photolysis at 335 nm. The method of spin trapping with 5,5-dimethyl-1-pyrroline-1-oxide (DMPO) and 4-pyridyl-1-oxide-N-tert-butylnitrone (POBN) combined with e.s.r. was used for the detection of OH, H and hydrated electrons. With the spin trap DMPO, photolysis generated DMPO-OH adducts under certain conditions but no DMPO-H adducts could be observed. With POBN, no POBN-H adducts were found. The formation of OH was confirmed by studying competition reactions for OH between the spin traps and OH scavengers (formate, isopropanol) and the concomitant formation of the CO-2 adduct and the (CH3)2COH adduct with both DMPO and POBN. The photochemical generation of OH radicals was pH dependent; at pH 7.5 no OH radicals could be detected. Photolysis (615-335 nm) of dicyanocomplexes of the Fe(III) porphyrins did not produce OH radicals. When corresponding Cu(II), Ni(II), Zn(II) and metal-free porphyrins were photolysed at 615 and 335 nm, no OH radicals could be spin trapped. These results tend to associate the well-known phenomenon of photoreduction of Fe(III) and Mn(III) porphyrins with the formation of OH radicals. This process is described mainly as the photoreduction of the metal ion by the ligand-bound hydroxyl ion via an intramolecular process.  相似文献   

9.
In this paper, we present a study about the influence of the porphyrin metal center and meso ligands on the biological effects of meso-tetrakis porphyrins. Different from the cationic meso-tetrakis 4-N-methyl pyridinium (Mn(III)TMPyP), the anionic Mn(III) meso-tetrakis (para-sulfonatophenyl) porphyrin (Mn(III)TPPS4) exhibited no protector effect against Fe(citrate)-induced lipid oxidation. Mn(III)TPPS4 did not protect mitochondria against endogenous hydrogen peroxide and only delayed the swelling caused by tert-BuOOH and Ca2+. Fe(III)TPPS4 exacerbated the effect of the tert-BuOOH, and both porphyrins did not significantly affect Fe(II)citrate-induced swelling. Consistently, Fe(III)TPPS4 predominantly promotes the homolytic cleavage of peroxides and exhibits catalytic efficiency ten-fold higher than Mn(III)TPPS4. For Mn(III)TPPS4, the microenvironment of rat liver mitochondria favors the heterolytic cleavage of peroxides and increases the catalytic efficiency of the manganese porphyrin due to the availability of axial ligands for the metal center and reducing agents such as glutathione (GSH) and proteins necessary for Compound II (oxomanganese IV) recycling to the initial Mn(III) form. The use of thiol reducing agents for the recycling of Mn(III)TPPS4 leads to GSH depletion and protein oxidation and consequent damages in the organelle.  相似文献   

10.
M Merkx  M W Pinkse  B A Averill 《Biochemistry》1999,38(31):9914-9925
The pH dependence of the catalytic parameters k(cat) and K(M) has been determined for the Fe(III)Fe(II)- and Fe(III)Zn(II)-forms of bovine spleen purple acid phosphatase (BSPAP). The parameter k(cat) was found to be maximal at pH 6.3, and a pK(a) of 5.4-5.5 was obtained for the acidic limb of the k(cat) vs pH profile. Two different EPR spectra were detected for the phosphate complex of the mixed-valent diiron enzyme; their relative amounts depended on the pH, with an apparent pK(a) of 6. The EPR spectra of Fe(III)Fe(II)-BSPAP.PO(4) and Fe(III)Zn(II)-BSPAP.PO(4) at pH 5.0 are similar to those previously reported for Fe(III)Fe(II)-Uf.PO(4) and Fe(III)Zn(II)-Uf.PO(4) complexes at pH 5.0. At higher pH, a new Fe(III)Fe(II)-BSPAP.PO(4) species is formed, with apparent g-values of 1.94, 1.71, and 1.50. The EPR spectrum of Fe(III)Zn(II)-BSPAP does not show significant changes upon addition of phosphate up to 30 mM at pH 6.5, suggesting that phosphate binds only to the spectroscopically silent Zn(II). To determine whether the phosphate complexes were good structural models for the enzyme substrate complexes, these complexes were studied using rapid-freeze EPR and stopped-flow optical spectroscopy. The stopped-flow studies showed the absence of burst kinetics at pH 7.0, which indicates that substrate hydrolysis is rate limiting, rather than phosphate release. The EPR spectrum of Fe(III)Fe(II)-BSPAP.p-NPP is similar, but not identical, to that of the corresponding phosphate complex, both at pH 5 and pH 6.5. We propose that both phosphate and p-NPP bridge the two metal ions at low pH. At higher pH where the enzyme is optimally active, we propose that hydroxide competes with phosphate and p-NPP for coordination to Fe(III) and that both phosphate and p-NPP coordinate only to the divalent metal ion.  相似文献   

11.
To achieve cellular iron deprivation by chelation, it is important to develop chelators with selective metal-binding properties. Selectivity for iron has long been the province of certain oxygen-donor chelators such as desferrioxamine, which target Fe(III) and exploit the strength of a relatively ionic Fe(III)-O interaction. We have been studying novel chelators that possess mechanisms to selectively chelate +2 biometals, particularly tachpyr [N,N',N"-tris(2-pyridylmethyl)-1,3,5-cis,cis-triaminocyclohexane] and derivatives from N,N',N"-trialkylation and pyridine ring alkylation. Metal-exchange and metal-binding competition reactions have been conducted at pH 7.4, 37 degrees C and time periods until no further change was observed (generally 24-48 h). Under anaerobic conditions, tachpyr is strongly selective for iron, binding 95+/-5% Fe(II) versus 5+/-5% Zn(II) in the forms [Fe(tachpyr)](2+) and [Zn(tachpyr)](2+) respectively. Under aerobic conditions, tachpyr complexes Fe(II) more effectively than Fe(III), forming iminopyridyl complexes [Fe(tachpyr-ox-n)](2+) (n=2, 4) by O(2)-induced and iron-mediated oxidative dehydrogenation. Complexes [Fe(tachpyr-ox-n)](2+) are also strongly bound forms of iron that are unaffected by an excess of Zn(II) (75 mol zinc:1 mol iron complex). The preference of tachpyr for iron over zinc under aerobic conditions appears to be hindered by oxidation of Fe(II) to Fe(III), such that the proportions bound are 44+/-10% Fe(II) versus 56+/-10% Zn(II), in the respective forms [Fe(tachpyr-ox-n)](2+) and [Zn(tachpyr)](2+). However, upon addition of the reducing agent Na(2)S(2)O(4) that converts Fe(III) to Fe(II), the binding proportions shift to 76+/-10% Fe(II) versus 24+/-10% Zn(II), demonstrating a clear preference of tachpyr for Fe(II) over Zn(II). Iron(II) is in the low-spin state in [Fe(tachpyr)](2+) and [Fe(tachpyr-ox-n)](2+) (n=2, 4), which is a likely cause of the observed selectivity. N-methylation of tachpyr [giving (N-methyl)(3)tachpyr] results in the loss of selectivity for Fe(II), which is attributed to the steric effect of the methyl groups and a resulting high-spin state of Fe(II) in [Fe(N-methyl)(3)tachpyr)](2+). The relationship of chelator selectivity to cytotoxicity in the tach family will be discussed.  相似文献   

12.
The use of Porphyrin derivatives as photosensitizers in Photodynamic Therapy (PDT) was investigated by means of a molecular docking study. These molecules can bind to intracellular targets such as P-type CaCa2+ ATPase of sarcoplasmic reticulum (SERCA1a). CAChe software was successfully employed for conducting the docking of Tetraphenylporphinesulfonate(TPPS), 5,10,15,20- Tetrakis (4-sulfonatophenyl) porphyrinato Iron(III) Chloride (FeTPPS) and 5,10,15,20-Tetrakis (4-sulfonatophenyl) porphyrinato Iron(III) nitrosyl Chloride (FeNOTPPS) with CaCa2+ ATPase from sarcoplasmic reticulum of rabbit. The results show that FeNOTPPS forms the most stable complex with CaCa2+ ATPase.  相似文献   

13.
The binding of the vasodilator drug papaverine (PAV) to micelles of zwitterionic N-hexadecyl-N,N-dimethyl-3-ammonio-1-propanesulfonate (HPS), cationic cetyltrimethylammonium chloride (CTAC) and anionic sodium dodecylsulfate (SDS) in aqueous solution was studied by 1H NMR and electronic absorption spectroscopy. In the presence of HPS or CTAC, the apparent pK(a) of PAV decreased by about 2 units, while it increased by about 2 units upon binding to SDS. However, the chemical shift patterns of both protonated (PAVH+) and deprotonated (PAV0) forms of PAV are not sensitive to the type of surfactant. The association constants were estimated as 5 +/- 2 M(-1) for PAVH+-CTAC, 8 +/- 3 M(-1) for PAVH+-HPS, (7 +/- 2) x 10(5) M(-1) for PAVH+-SDS, and 1.5 x 10(3) to 3.0 x 10(3) M(-1) for the complexes of PAV0 with all three types of micelles. Using these data, an electrostatic potential difference on the micelle-water interface was calculated as 150 +/- 10 mV for CTAC, 140 +/- 10 mV for HPS and - 140 +/- 10 mV for SDS. The results suggest that PAV aromatic rings are located in the hydrophobic part of the micelle. The electrostatic attraction or repulsion of the protonated quinoline nitrogen and surfactant headgroups changes the affinity of PAV to micelles and, thus, shifts the ionization equilibrium of PAV. The electrostatic potential of HPS micellar surface is determined by the cationic dimethylammonium headgroup fragment, whereas the anionic sulfate fragment attenuates the effective charge of HPS headgroup.  相似文献   

14.
The action of irradiated cationic Fe(III)TMPyP and anionic Fe(III)TPPS4 forms of mesoporphyrins on mitochondrial functions was investigated using experimental conditions that caused minimal effects on mitochondria in the dark. Treatment of mitochondria with 1 microM Fe(III)TMPyP for 2 min decreased the respiratory control by 3% in the dark and 28% after irradiation. Fe(III)TPPS4 (1 microM) had no significant effect on respiratory control under any of the above conditions. Both porphyrins increased the mitochondrial production of reactive oxygen species in the presence of Ca2+; however, the effect of Fe(III)TMPyP was significantly stronger. In both cases, this overproduction was associated with membrane lipid peroxidation. It was also observed that the association constant of Fe(III)TMPyP with mitochondria was 11 times higher than that of Fe(III)TPPS4. In conclusion, the damage to isolated mitochondria induced by Fe(III)TMPyP under illumination was larger than by Fe(III)TPPS4, probably because its cationic charge favors association with the mitochondrial membrane. This is supported by the decrease in the association constant of Fe(III)TMPyP with mitochondria in higher salt medium.  相似文献   

15.
In a previous study (Minotti, G., and Ikeda-Saito, M. (1991) J. Biol. Chem. 266, 20011-20017) we demonstrated the existence of a M(r) 66,000 microsomal iron protein (MIP) which stimulates NADPH oxidation by shunting electrons from NADPH-cytochrome P-450 reducase to its bound Fe(III). In the present study, purified MIP was depleted of iron and the apoMIP was examined for its ability to incorporate Fe(III) upon an incubation with Fe(II). It was found that apoMIP had an oxygen-dependent ferroxidase activity coupled with the incorporation of Fe(III). The reconstituted MIP exhibited a Fe(III) content and an NADPH oxidation activity similar to those of native MIP. However, the reconstitution of MIP from apoMIP and Fe(II) had to be performed in the presence of detergents to prevent the formation of protein aggregates and the oxidative incorporation of an iron which could not react with NADPH-cytochrome P-450 reductase. This redox inactive iron was probably bound nonspecifically to artifactual sites formed by the protein aggregates.  相似文献   

16.
The three-dimensional conformation of ferricytochrome c results from specific folding of the polypeptide chain around the covalently bound heme so that His-18 and Met-80 are axially coordinated to the Fe(III). The Fe(III)-free, porphyrin protein has an intrinsic viscosity, sedimentation coefficient, and circular dichroism indicative of a compact, globular protein conformation comparable to the holoprotein. Both the porphyrin protein and ferricytochrome c are reversibly denatured by guanidinium chloride. Refolding of the porphyrin protein occurs in essentially a single, exceptionally rapid kinetic phase (tau = 14 ms, 0.75 M guanidinium chloride, pH 6.5, 25 degrees C); whereas refolding of ferricytochrome c occurs in two slower kinetic phases (TAU 1 = 0.10 S, TAU 2 = 20 S) UNDER COMPARABLE CONDITIONS. The presence of Fe(III) in the metalloporphyrin of ferricytochrome c thus has a major effect on the protein folding kinetics. The slow kinetic phase is evidently due to this effect of Fe(III) and not to the slow cis-trans isomerism of the peptide bond of proline residues as has been suggested.  相似文献   

17.
Mn(III) and Fe(III) complexes of meso-tetrakis(N-methylpyridinium-4-yl)porphyrin (M-TMePyP) and related hybrid molecules ("metalloporphyrin-ellipticine") were activated by potassium monopersulfate in the presence of variable calf thymus (CT) DNA and NaCl concentrations. Monitored by visible spectroscopy (Soret band), fast degradation of the free metalloprophyrin was observed while the DNA-bound form appeared protected. This direct quantitation of free versus bound metalloporphyrin ratios allowed determination of binding constants: Mn- and Fe-TMePyP respectively bind to CT DNA (5 mM phosphate buffer, 0.1 M NaCl, pH 7) with K = 3 X 10(4) and 1.2 X 10(4) M-1. Mn-TMePyP showed a greater affinity for poly[d(A-T)] (K = 1.2 X 10(5) M-1) than for poly[d(G-C)] (K = 0.2 X 10(4) M-1). This method allowed us access to the intrinsic DNA affinity of the metalloporphyrin moiety of the hybrid molecules "metalloporphyrin-ellipticine".  相似文献   

18.
The binding of several different categories of small molecules to bovine (BSA) and human (HSA) serum albumins has been studied for many years through different spectroscopic techniques to elucidate details of the protein structure and binding mechanism. In this work we present the results of the study of the interactions of BSA and HSA with the anionic sodium dodecyl sulfate (SDS), cationic cethyltrimethylammonium chloride (CTAC) and zwitterionic N-hexadecyl-N,N-dimethyl-3-ammonium-1-propanesulfonate (HPS) monitored by fluorescence spectroscopy of the intrinsic tryptophans at pH 5.0. Similarly to pH 7.0 and 9.0, at low concentrations, the interaction of BSA with these surfactants shows a quenching of fluorescence with Stern-Volmer quenching constants of (1.1+/-0.1)x10(4) M(-1), (3.2+/-0.1)x10(3) M(-1) and (2.1+/-0.1)x10(3) M(-1) for SDS, HPS and CTAC, respectively, which are associated to the 'effective' association constants to the protein. On the interaction of these surfactants with HSA, an opposite effect was observed as compared to BSA, i.e., an enhancement of fluorescence takes place. For both proteins, at low surfactant concentrations, a positive cooperativity was observed and the Hill plot model was used to estimate the number of surfactant binding sites, as well as the association constants of the surfactants to the proteins. It is worthy of notice that the binding constants for the surfactants at pH 5.0 are lower as compared to pH 7.0 and 9.0. This is probably due to fact that the protein at this acid pH is quite compact reducing the accessibility of the surfactants to the hydrophobic cavities in the binding sites. The interaction of myristic acid with both proteins shows a similar fluorescence behaviour, suggesting that the mechanism of the interaction is the same. Recently published crystallographic studies of HSA-myristate complex were used to perform a modelling study with the aim to explain the fluorescence results. The crystallographic structure reveals that a total of five myristic acid molecules are asymmetrically bound in the macromolecule. Three of these sites correspond to higher affinity ones and correlate with high association constants described in the literature. Our models for BSA and HSA with bound SDS suggest that the surfactant could be bound at the same sites as those reported in the crystal structure for the fatty acid. The differences in tryptophan vicinity upon surfactant binding are explored in the models in order to explain the observed spectroscopic changes. For BSA the quenching is due to a direct contact of a surfactant molecule with the indole of W131 residue. It is clear that the binding site in BSA which is very close, in contact with tryptophan W131, corresponds to a lower affinity site, explaining the lower binding constants obtained from fluorescence studies. In the case of HSA the enhancement of fluorescence is due to the removal of static quenching of W214 residue in the intact protein caused by nearby residues in the vicinity of this tryptophan.  相似文献   

19.
Although metal ions can promote amyloid formation from many proteins, their effects on the formation of amyloid from transthyretin have not been previously studied. We therefore screened the effects of Cu(II), Zn(II), Al(III), and Fe(III) on amyloid formation from wild-type (WT) transthyretin as well as its V30M, L55P, and T119M mutants. Cu(II) and Zn(II) promoted amyloid formation from the L55P mutant of transthyretin at pH 6.5 but had little effect on amyloid formation from the other forms of the protein. Zn(II) promoted L55P amyloid formation at pH 7.4 but Cu(II) inhibited it. Cu(II) gave dose-dependent quenching of the tryptophan fluorescence of transthyretin and the fluorescence of 1-anilino-8-naphthalene sulfonate bound to it. Zn(II) gave dose-dependent quenching of the tryptophan but not the 1-anilino-8-naphthalene sulfonate fluorescence. Apparent dissociation constants for Cu(II) and Zn(II) binding at pH 7.4 of approximately 10 nM and approximately 1 microM (approximately 0.4 microM and approximately 5 microM at pH 6.5), respectively, were obtained from the quenching data. Zn(II) enhanced urea-mediated the dissociation of the L55P but not the WT transthyretin tetramer. Cu(II), depending on its concentration, either had no effect or stabilized the WT tetramer but could enhance urea-mediated dissociation of L55P.  相似文献   

20.
Iron(III) meso-tetrakis (4-N-methylpyridiniumyl) porphyrin (FeTMPyP) undergoes a complex equilibrium in aqueous solution as a function of pH. Use of phosphate buffers, a common practice in biomedical applications of porphyrins, suggests the complexation of phosphate anion at the sixth coordination position to the iron, which contributes to the complexity of the equilibrium in the pH range from 1 to 4. In the absence of phosphate the equilibrium is simplified in a similar way as in the presence of high salt concentrations. Combined use of optical absorption, (1)H NMR and infrared spectroscopies, together with the literature data, suggest the formation of hexacoordinated monoaqueous-phosphate FeTMPyP complex in a limited acidic pH range. Discussion of the behavior of cationic FeTMPyP as compared to anionic iron(III) meso-tetrakis (4-sulfonatophenyl) porphyrin (FeTPPS(4)) is presented in regard to equilibrium of different species to explain the observed complex equilibrium.  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