Your browser doesn't support javascript.
loading
Mostrar: 20 | 50 | 100
Resultados 1 - 20 de 27
Filtrar
Mais filtros










Base de dados
Intervalo de ano de publicação
1.
Biochemistry ; 51(23): 4693-703, 2012 Jun 12.
Artigo em Inglês | MEDLINE | ID: mdl-22616883

RESUMO

Mammalian (Clade 3) catalases utilize NADPH as a protective cofactor to prevent one-electron reduction of the central reactive intermediate Compound I (Cpd I) to the catalytically inactive Compound II (Cpd II) species by re-reduction of Cpd I to the enzyme's resting state (ferricatalase). It has long been known that ascorbate/ascorbic acid is capable of reducing Cpd I of NADPH-binding catalases to Cpd II, but the mode of this one-electron reduction had hitherto not been explored. We here demonstrate that ascorbate-mediated reduction of Cpd I, generated by addition of peroxoacetic acid to NADPH-free bovine liver catalase (BLC), requires specific binding of the ascorbate anion to the NADPH binding pocket. Ascorbate-mediated Cpd II formation was found to be suppressed by added NADPH in a concentration-dependent manner, for the achievement of complete suppression at a stoichiometric 1:1 NADPH:heme concentration ratio. Cpd I → Cpd II reduction by ascorbate was similarly inhibited by addition of NADH, NADP(+), thio-NADP(+), or NAD(+), though with 0.5-, 0.1-, 0.1-, and 0.01-fold reduced efficiencies, respectively, in agreement with the relative binding affinities of these dinucleotides. Unexpected was the observation that although Cpd II formation is not observed in the presence of NADP(+), the decay of Cpd I is slightly accelerated by ascorbate rather than retarded, leading to direct regeneration of ferricatalase. The experimental findings are supported by molecular mechanics docking computations, which show a similar binding of NADPH, NADP(+), and NADH, but not NAD(+), as found in the X-ray structure of NADPH-loaded human erythrocyte catalase. The computations suggest that two ascorbate molecules may occupy the empty NADPH pocket, preferably binding to the adenine binding site. The biological relevance of these findings is discussed.


Assuntos
Ácido Ascórbico/metabolismo , Catalase/metabolismo , NADP/metabolismo , Animais , Ácido Ascórbico/química , Aspergillus niger/enzimologia , Sítios de Ligação , Catalase/química , Bovinos , Humanos , Modelos Moleculares , NADP/análogos & derivados , Ligação Proteica , Conformação Proteica
2.
Plant Physiol Biochem ; 59: 80-9, 2012 Oct.
Artigo em Inglês | MEDLINE | ID: mdl-22277729

RESUMO

The nitric oxide-specific fluorescent probe Fluorescent Nitric Oxide Cheletropic Trap (FNOCT) 8a was applied to intact tobacco (Nicotiana tabacum cv. Samsun) roots to detect sites of nitric oxide formation and NO distribution. Three week old tobacco seedlings were gently removed from the sand culture pots with intact roots and transferred to small Petri dishes, whose base was replaced by a thin coverslip. Intact roots were subjected to FNOCT 8a to localize NO-dependent fluorescence in these roots; controls with an exogenous NO donor confirmed the presence and distribution of the probe in the roots. To confirm the NO-dependent fluorescence, roots were incubated with the three different NO scavengers cPTIO {2-(4-Carboxyphenyl)-4,4,5,5-tetramethylimidazoline-L-oxyl-3-oxide}, methylene blue and sodium diethyl dithiocarbamate (DCC) followed by incubation with FNOCT 8a. Methylene blue and DCC were able to completely quench NO-dependent fluorescence, cPTIO quenched partially. The roots were incubated in the presence of NaNO2 and NaNO3, which are substrates for nitrite:nitric oxide reductase (NI-NOR) and plasma membrane-bound nitrate reductase (PM-NR), respectively. The NO-dependent fluorescence was more or less same at the root tips upon treatment with NaNO2, while the overall fluorescence was reduced in the presence of NaNO. Fluorescence from the living roots was visualized by inverted confocal laser scanning microscope (CLSM) using UV laser (excitation 360 nm and emission 408 nm). A specialized apparatus has been devised by the authors for analysis of intact roots as described in the methods section of this paper. Intact roots were chosen for microscopic observation rather than incised roots to avoid production of NO due to stress or physical injury.


Assuntos
Corantes Fluorescentes , Nicotiana/metabolismo , Nitratos/farmacologia , Óxido Nítrico/metabolismo , Nitrito de Sódio/farmacologia , Membrana Celular/enzimologia , Citosol/metabolismo , Proteínas de Membrana/metabolismo , Microscopia Confocal , Nitrato Redutase/metabolismo , Oxirredutases/metabolismo , Proteínas de Plantas/metabolismo , Raízes de Plantas/citologia , Raízes de Plantas/efeitos dos fármacos , Raízes de Plantas/metabolismo , Plântula/citologia , Plântula/efeitos dos fármacos , Plântula/metabolismo , Nicotiana/citologia , Nicotiana/efeitos dos fármacos
3.
Chemistry ; 16(36): 11121-32, 2010 Sep 24.
Artigo em Inglês | MEDLINE | ID: mdl-20680937

RESUMO

The synthesis and the structural and spectroscopic characterization of nonfluorescent, pyrene-based cyclic o-quinodimethanes are reported. These compounds react efficiently with nitric oxide (NO) in a formal cheletropic manner, by which the fluorescent aromatic pyrene system is regenerated. The NO trapping capabilities and kinetics of the fluorescent nitric oxide cheletropic traps (FNOCTs) are assessed in THF and buffered aqueous solution by ESR, UV/Vis, and fluorescence spectroscopy, by employing NO solutions and NO released from N-diazeniumdiolates (NONOates). Prototypal biological applications include the quantitation of NO production from cultured rat alveolar macrophages and the endothelium of porcine aorta, which demonstrate a sensitivity for NO detection in the nanomolar range.


Assuntos
Aorta/química , Endotélio/química , Macrófagos Alveolares/química , Óxido Nítrico/análise , Óxido Nítrico/química , Pirenos/química , Animais , Aorta/metabolismo , Técnicas de Cultura de Células , Cristalografia por Raios X , Endotélio/metabolismo , Macrófagos Alveolares/metabolismo , Estrutura Molecular , Ratos , Ratos Wistar , Sensibilidade e Especificidade , Espectrometria de Fluorescência/métodos
4.
Chem Biol Drug Des ; 73(5): 494-501, 2009 May.
Artigo em Inglês | MEDLINE | ID: mdl-19366358

RESUMO

In previous studies we reported on the catalase-like activity and antioxidative properties of a non-heme Fe(III)-tetraaza[14]annulene complex, 5,4-didehydro-5,9,14,18-tetraaza-di(2,2-dimethyl-[5,6]benzo[1,3]dioxolo)[a,h]cyclotetradecene--Fe(III) chloride (TAA-1/Fe). We proposed that intracellular application of the parent, iron-free tetraaza[14]annulene ligand, TAA-1, as precursor would allow antioxidative defense along two lines, i.e. by chelation of potentially toxic cellular iron ions and, subsequently, by catalase-mimic activity. We here set out to establish whether the active catalase mimic is indeed formed intracellularly when cells are loaded with the ligand. When isolated rat hepatocytes were preloaded with TAA-1, they were protected against iron-induced cell injury and oxidative stress elicited by exposure to the membrane-permeable iron complex Fe(III)/8-hydroxyquinoline. After lysis of the cells, followed by ultrafiltration to remove endogenous catalase, the lysate exhibited catalase-like activity, while lysates of control cells not treated with TAA-1 showed no catalase-like activity. By comparison with authentic TAA-1/Fe, an intracellular formation of 2.0 +/- 0.3 microm of the active catalase mimic in native hepatocytes exposed to TAA-1 and of 6.5 +/- 1.0 microm in hepatocytes exposed to both TAA-1 and iron ions was estimated. The intracellular formation of the active catalase mimic thus renders TAA-1 an attractive compound for protection against iron- and/or hydrogen peroxide-dependent cell injuries.


Assuntos
Compostos Aza/metabolismo , Benzodioxóis/metabolismo , Compostos Férricos/metabolismo , Hepatócitos/metabolismo , Ferro/toxicidade , Animais , Compostos Aza/síntese química , Compostos Aza/química , Benzodioxóis/síntese química , Benzodioxóis/química , Catalase/metabolismo , Sobrevivência Celular , Células Cultivadas , Compostos Férricos/síntese química , Compostos Férricos/química , Peróxido de Hidrogênio/metabolismo , Ferro/farmacologia , Masculino , Estresse Oxidativo , Oxiquinolina/farmacologia , Ratos , Ratos Wistar
5.
Chemphyschem ; 9(16): 2397-405, 2008 Nov 10.
Artigo em Inglês | MEDLINE | ID: mdl-18956403

RESUMO

Zwitterionic diazeniumdiolates of the form RN[N(O)NO(-)](CH(2))(2)NH(2) (+)R, where R=CH(3) (1), (CH(2))(3)CH(3) (2), (CH(2))(5)CH(3) (3), and (CH(2))(7)CH(3) (4) were synthesized by reaction of the corresponding diamines with nitric oxide. Spectrophotometrically determined pK(a)(O) values, attributed to protonation at the terminal oxygen of the diazeniumdiolate group, show shifts to higher values in dependence of the chain lengths of R. The pH dependence of the decomposition of NO donors 1-3 was studied in buffered solution between pH 5 and 8 at 22 degrees C, from which pK(a)(N) values for protonation at the amino nitrogen, leading to release of NO, were estimated. It is shown that the decomposition of these diazeniumdiolates is markedly catalyzed by anionic SDS micelles. First-order rate constants for the decay of 1-4 were determined in phosphate buffer pH 7.4 at 22 degrees C as a function of SDS concentration. Micellar binding constants, K(SM), for the association of diazeniumdiolates 1-3 with the SDS micelles were also determined, again showing a significant increase with increasing length of the alkyl side chains. The decomposition of 1-3 in micellar solution is quantitatively described by using the pseudo-phase ion-exchange (PIE) model, in which the degree of micellar catalysis is taken into account through the ratio of the second-order rate constants (k(2m)/k(2w)) for decay in the micelles and in the bulk aqueous phase. The decay kinetics of 1-3 were further studied in the presence of cosolvents and nonionic surfactants, but no effect on the rate of NO release was observed. The kinetic data are discussed in terms of association to the micelle-aqueous phase interface of the negatively charged micelles. The apparent interfacial pH value of SDS micelles was evaluated from comparison of the pH dependence of the first-order decay rate constants of 2 and 3 in neat buffer and the rate data obtained for the surfactant-mediated decay. For a bulk phase of pH 7.4, an interfacial pH of 5.7-5.8 was determined, consistent with the distribution of H(+) in the vicinity of the negatively charged micelles. The data demonstrate the utility of 2 and 3 as probes for the determination of the apparent pH value in the Stern region of anionic micelles.


Assuntos
Compostos Azo/química , Micelas , Dodecilsulfato de Sódio/química , Tensoativos/química , Catálise , Concentração de Íons de Hidrogênio , Interações Hidrofóbicas e Hidrofílicas , Cinética , Doadores de Óxido Nítrico/química , Água
6.
J Am Chem Soc ; 130(23): 7345-56, 2008 Jun 11.
Artigo em Inglês | MEDLINE | ID: mdl-18479132

RESUMO

X-ray structures of the 13 different monofunctional heme catalases published to date were scrutinized in order to gain insight in the mechanism by which NADPH in Clade 3 catalases may protect the reactive ferryloxo intermediate Compound I (Cpd I; por (*+)Fe (IV)O) against deactivation to the catalytically inactive intermediate Compound II (Cpd II; porFe (IV)O). Striking similarities in the molecular network of the protein subunits encompassing the heme center and the surface-bound NADPH were found for all of the Clade 3 catalases. Unique features in this region are the presence of a water molecule (W1) adjacent to the 4-vinyl group of heme and a serine residue or a second water molecule hydrogen-bonded to both W1 and the carbonyl group of a threonine-proline linkage, with the proline in van der Waals contact with the dihydronicotinamide group of NADPH. A mechanism is proposed in which a hydroxyl anion released from W1 undergoes reversible nucleophilic addition to the terminal carbon of the 4-vinyl group of Cpd I, thereby producing a neutral porphyrin pi-radical ferryloxo (HO-por (*)Fe (IV)O) species of reduced reactivity. This structure is suggested to be the elusive Cpd II' intermediate proposed in previous studies. An accompanying proton-shifting process along the hydrogen-bonded network is believed to facilitate the NADPH-mediated reduction of Cpd I to ferricatalase and to serve as a funnel for electron transfer from NADPH to the heme center to restore the catalase Fe (III) resting state. The proposed reaction paths were fully supported as chemically reasonable and energetically feasible by means of density functional theory calculations at the (U)B3LYP/6-31G* level. A particularly attractive feature of the present mechanism is that the previously discussed formation of protein-derived radicals is avoided.


Assuntos
Catalase/química , Catalase/metabolismo , Compostos Férricos/química , NADP/química , NADP/metabolismo , Cristalografia por Raios X , Heme/química , Heme/metabolismo , Modelos Moleculares , Teoria Quântica , Relação Estrutura-Atividade , Termodinâmica
7.
Inorg Chem ; 46(26): 11416-30, 2007 Dec 24.
Artigo em Inglês | MEDLINE | ID: mdl-18001111

RESUMO

The development of enzyme mimics of catalase which decompose hydrogen peroxide to water and molecular oxygen according to the 2:1 stoichiometry of native catalase and in aqueous solution at pH 7 and at micromolar concentrations of the enzyme model and hydrogen peroxide is reported. For this purpose, iron(III) complexes of 1,4,8,11-tetraaza[14]annulenes are prepared by various procedures. Efficacious preparations utilize reaction of the [N4] macrocyles with FeII salts in the presence of triphenylamine, followed by gentle oxidation of the FeII complexes by molecular oxygen or by tris(4-bromophenyl)aminium hexachloroantimonate. The complexes are characterized by SQUID magnetometry and by Mössbauer, EPR, and UV/vis spectrometry. In the solid state, the iron(III) center of the catalytically active complexes exists in the intermediate (quartet, S = 3/2) spin state. Several of these complexes decompose hydrogen peroxide in aqueous buffer solution at pH 7.2 at room temperature with turnover numbers between 40 and 80. The apparent second-order rate constant for hydrogen peroxide decomposition is in the range of 1400-2400 M(-1) s(-1), about 3 orders of magnitude lower than the value for native catalase. Besides oxygen production, a non-oxygen releasing pathway of hydrogen peroxide decomposition is unveiled.


Assuntos
Compostos Aza/química , Compostos Aza/metabolismo , Materiais Biomiméticos/química , Materiais Biomiméticos/metabolismo , Catalase/metabolismo , Compostos Férricos/química , Compostos Policíclicos/química , Compostos Aza/síntese química , Materiais Biomiméticos/síntese química , Elétrons , Peróxido de Hidrogênio/química , Ligantes , Magnetismo , Estrutura Molecular , Oxigênio/química , Oxigênio/metabolismo , Compostos Policíclicos/síntese química , Compostos Policíclicos/metabolismo , Espectrofotometria , Temperatura
8.
J Phys Chem B ; 111(45): 12985-92, 2007 Nov 15.
Artigo em Inglês | MEDLINE | ID: mdl-17958349

RESUMO

A new pyrene-based fluorescent probe for the determination of critical micelle concentrations (CMC) is described. The title compound 1 is obtained in five steps, starting from pyrene. Fluorescence spectroscopic properties of 1 are studied in homogeneous organic solvents and aqueous micellar solutions. In a wide range of organic solvents, probe 1 exhibits a characteristic monomer emission of the pyrene fluorophore, with three distinct peak maxima at 382, 404, and 425 nm. The spectra change dramatically in aqueous solution, where no monomer emission of the pyrene fluorophore is detected. Instead, only strong excimer fluorescence with a broad, red-shifted emission band at lambda(max) = 465 nm is observed. In micellar aqueous solution, a superposition of the monomer and excimer emission is found. The appearance of the monomer emission in micellar solution can be explained on the basis of solubilization of 1 by the surfactant micelles. The ratio of the monomer to excimer fluorescence intensities of 1 is highly sensitive to changes in surfactant concentration. This renders 1 a versatile and sensitive probe molecule for studying the micellization of ionic and nonionic surfactants. For a representative selection of common surfactants, the critical micelle concentrations in aqueous solution are determined, showing excellent agreement with established literature data.


Assuntos
Corantes Fluorescentes/química , Micelas , Pirenos/química , Tensoativos/química , Cetrimônio , Compostos de Cetrimônio/química , Cristalografia por Raios X , Corantes Fluorescentes/síntese química , Interações Hidrofóbicas e Hidrofílicas , Estrutura Molecular , Pirenos/síntese química , Dodecilsulfato de Sódio/química , Espectrometria de Fluorescência , Eletricidade Estática , Água/química
9.
Chemistry ; 13(15): 4230-45, 2007.
Artigo em Inglês | MEDLINE | ID: mdl-17323385

RESUMO

Non-heme iron(III) complexes of 14-membered tetraaza macrocycles have previously been found to catalytically decompose hydrogen peroxide to water and molecular oxygen, like the native enzyme catalase. Here the mechanism of this reaction is theoretically investigated by DFT calculations at the (U)B3LYP/6-31G* level, with focus on the reactivity of the possible spin states of the FeIII complexes. The computations suggest that H2O2 decomposition follows a homolytic route with intermediate formation of an iron(IV) oxo radical cation species (L.+FeIV==O) that resembles Compound I of natural iron porphyrin systems. Along the whole catalytic cycle, no significant energetic differences were found for the reaction proceeding on the doublet (S=1/2) or on the quartet (S=3/2) hypersurface, with the single exception of the rate-determining O--O bond cleavage of the first associated hydrogen peroxide molecule, for which reaction via the doublet state is preferred. The sextet (S=5/2) state of the FeIII complexes appears to be unreactive in catalase-like reactions.


Assuntos
Catalase , Compostos Férricos/química , Peróxido de Hidrogênio/química , Porfirinas/química , Catálise
10.
Chembiochem ; 8(3): 341-52, 2007 Feb 12.
Artigo em Inglês | MEDLINE | ID: mdl-17219451

RESUMO

Chelatable cellular iron, and chelatable mitochondrial iron in particular, has yet to be well characterized, so the overall strength with which these "loosely bound" iron ions (presumably mainly Fe(II)) are intracellularly/intramitochondrially bound is unclear. We have previously reported the first selective mitochondrial iron indicator: rhodamine B 4-[(1,10-phenanthrolin-5-yl)aminocarbonyl]benzyl ester (RPA). With this compound as a model, we have now developed two additional mitochondrial iron indicators with very different iron-binding affinities and have applied these to the study of the chelatable iron pool in the mitochondria of isolated rat liver cells. With the new indicator rhodamine B 4-[(2,2'-bipyridin-4-yl)aminocarbonyl]benzyl ester (RDA), with 2,2'-bipyridine as chelating unit (log beta(3)=17.5), essentially the same iron concentration (16.0+/-1.9 microM) was determined as with RPA (log beta(3)=21.1), despite the four orders of magnitude difference in Fe(II)-binding affinity. This not only demonstrates the reliability of the procedure, but also confirms that iron complexation by these indicators does not induce any significant release of iron from the iron-storage proteins on the timescale of the experiment. In contrast, the indicator rhodamine B 4-[bis(pyridin-2-ylmethyl)aminomethyl]benzyl ester (PIRO), with an N,N-bis(pyridin-2-ylmethyl)amine group as chelating component (log beta(2)=12.2), could not compete against the array of endogenous ligands. The intramitochondrial concentrations of the three indicators were determined to be in the range of 100 microM: that is, about three orders of magnitude lower than the total concentration of endogenous compounds that might chelate iron ions. It is therefore estimated that chelatable mitochondrial iron ions are bound by endogenous ligands with apparent stability constants (log K(app)) of between 9 and 14.


Assuntos
Corantes Fluorescentes/síntese química , Quelantes de Ferro/química , Ferro/análise , Mitocôndrias/metabolismo , Piridinas/síntese química , Rodaminas/síntese química , Animais , Células Cultivadas , Corantes Fluorescentes/química , Hepatócitos/metabolismo , Ferro/metabolismo , Quelantes de Ferro/análise , Mitocôndrias/efeitos dos fármacos , Estrutura Molecular , Piridinas/química , Ratos , Rodaminas/química
11.
Chemistry ; 12(34): 8786-97, 2006 Nov 24.
Artigo em Inglês | MEDLINE | ID: mdl-16952125

RESUMO

During the past years, there has been increasing interest in endogenous nitric oxide storage compounds. Recently, we briefly reported on the ascorbate-dependent release of nitric oxide ((.)NO) from N-nitrosotryptophan derivatives. In the present study, the underlying mechanism of (.)NO release is studied in more detail, primarily utilizing N-acetyl-N-nitrosotryptophan (NANT) as a model compound. The initial rate of the ascorbate-induced release of nitric oxide has been found to correspond to the rate of NANT decay. In this process, N-acetyltryptophan (NAT) is produced almost quantitatively. The final yield of nitrite amounted to around 90 % with respect to the applied amount of NANT. However, the total release of nitric oxide was only 60 %, as determined by using an FNOCT-4(fluorescent nitric oxide cheletropic trap number 4) assay. Besides nitric oxide, a second volatile product, dinitrogen oxide (N(2)O), has been identified by using (15)N NMR spectrometry, strongly indicating the intermediacy of nitroxyl (HNO). The formation of intermediate ascorbyl radical anions during the NANT-ascorbate reaction has been monitored by using ESR spectrometry. Unexpectedly, it was found that the primary oxidized product of vitamin C, dehydroascorbic acid (DHA), efficiently consumes nitric oxide. Since ESR spectrometry further revealed that ascorbyl radical anions are also generated during the spontaneous decay of DHA, the DHA-nitric oxide reaction is related to recombination of (.)NO with the thus formed ascorbyl radical anions. A conclusively established mechanism of the NANT-ascorbate reaction is presented, with O-nitrosoascorbate as a key intermediate, as additionally supported by CBS-QB3 calculations. The present study suggests that vitamin C and its oxidation products can chemically counterbalance endogenous nitric oxide levels.


Assuntos
Ácido Ascórbico/química , Óxido Nítrico/química , Compostos Nitrosos/química , Triptofano/análogos & derivados , Ácido Desidroascórbico/química , Espectroscopia de Ressonância de Spin Eletrônica , Radicais Livres/química , Espectroscopia de Ressonância Magnética , Estrutura Molecular , Nitritos/química , Óxidos de Nitrogênio/química , Nitrosação , Oxirredução
12.
Free Radic Res ; 40(1): 67-74, 2006 Jan.
Artigo em Inglês | MEDLINE | ID: mdl-16298761

RESUMO

Heme catalases are considered to degrade two molecules of H(2)O(2) to two molecules of H(2)O and one molecule of O(2) employing the catalatic cycle. We here studied the catalytic behaviour of bovine liver catalase at low fluxes of H(2)O(2) (relative to catalase concentration), adjusted by H(2)O(2)-generating systems. At a ratio of a H(2)O(2) flux (given in microM/min(- 1)) to catalase concentration (given in microM) of 10 min(- 1) and above, H(2)O(2) degradation occurred via the catalatic cycle. At lower ratios, however, H(2)O(2) degradation proceeded with increasingly diminished production of O(2). At a ratio of 1 min(- 1), O(2) formation could no longer be observed, although the enzyme still degraded H(2)O(2). These results strongly suggest that at low physiological H(2)O(2) fluxes H(2)O(2) is preferentially metabolised reductively to H(2)O, without release of O(2). The pathways involved in the reductive metabolism of H(2)O(2) are presumably those previously reported as inactivation and reactivation pathways. They start from compound I and are operative at low and high H(2)O(2) fluxes but kinetically outcompete the reaction of compound I with H(2)O(2) at low H(2)O(2) production rates. In the absence of NADPH, the reducing equivalents for the reductive metabolism of H(2)O(2) are most likely provided by the protein moiety of the enzyme. In the presence of NADPH, they are at least in part provided by the coenzyme.


Assuntos
Catalase/metabolismo , Peróxido de Hidrogênio/química , Fígado/enzimologia , Animais , Catalase/antagonistas & inibidores , Catálise , Bovinos , Peróxido de Hidrogênio/metabolismo , Cinética , NADP/química , NADP/metabolismo , Espectrofotometria Ultravioleta
13.
Free Radic Biol Med ; 37(9): 1369-83, 2004 Nov 01.
Artigo em Inglês | MEDLINE | ID: mdl-15454276

RESUMO

Hydrogen peroxide is involved in many types of cell injury and exerts most of its injurious effects in conjunction with chelatable iron. We previously described a synthetic nonporphyrin iron-containing catalase mimic, TAA-1/Fe. Its ligand TAA-1 was designed for application in biological systems in which it is supposed to fulfill a dual task: it should chelate cellular labile iron and thus form the active catalase mimic, thereby decreasing levels of redox-active iron and enhancing the degradation of hydrogen peroxide. Here, we tested these novel compounds in cellular systems, i.e., in cultured hepatocytes and liver endothelial cells. Both the iron complex, i.e., the complete mimic, and the ligand, i.e., the putative precursor of this mimic, provided protection against endothelial cell injury induced by exogenous hydrogen peroxide. Furthermore, the ligand--but not (or less so) the complex--strongly protected both cell types against iron-dependent hypothermic injury and hepatocytes against iron-induced cell injury and against iron-dependent, histidine-induced injury. Together, these results demonstrate that the putative catalase mimic precursor TAA-1 is able to protect cells against iron- and/or hydrogen peroxide-dependent cell injuries and that--in line with our initial concept--it is likely to exert its protection by both iron chelation and hydrogen peroxide degradation.


Assuntos
Catalase/metabolismo , Sobrevivência Celular/efeitos dos fármacos , Hepatócitos/patologia , Peróxido de Hidrogênio/toxicidade , Ferro/toxicidade , Animais , Hepatócitos/efeitos dos fármacos , Hipotermia Induzida , Cinética , Masculino , Ratos , Ratos Wistar
14.
Chemistry ; 10(11): 2707-21, 2004 Jun 07.
Artigo em Inglês | MEDLINE | ID: mdl-15195302

RESUMO

The effect of pressure on the oxidation of hydroarenes 3-9 with 2,3-dichloro-5,6-dicyano-1,4-quinone (DDQ; 1 a) or o-chloranil (10), leading to the corresponding arenes, has been investigated. The activation volumes were determined from the pressure dependence of the rate constants of these reactions monitored by on-line UV/Vis spectroscopic measurements in an optical high-pressure cell (up to 3500 bar). The finding that they are highly negative and only moderately dependent on the solvent polarity (DeltaV( not equal ) = -13 to -25 in MTBE and -15 to -29 cm(3) mol(-1) in MeCN/AcOEt, 1:1) rules out the formation of ionic species in the rate-determining step and is good evidence for a hydrogen atom transfer mechanism leading to a pair of radicals in the rate-determining step, as was also suggested by kinetic measurements, studies of kinetic isotope effects, and spin-trapping experiments. The strong pressure dependence of the kinetic deuterium isotope effect for the reaction of 9,10-dihydroanthracene 5/5-9,9,10,10-D(4) with DDQ (1 a) can be attributed to a tunneling component in the hydrogen transfer. In the case of formal 1,3-dienes and enes possessing two vicinal C--H bonds, which have to be cleaved during the dehydrogenation, a pericyclic hydrogen transfer has to considered as one mechanistic alternative. The comparison of the kinetic deuterium isotope effects determined for the oxidation of tetralin 9/9-1,1,4,4-D(4)/9-2,2,3,3-D(4)/9-D(12) either with DDQ (1 a) or with thymoquinone 1 c indicates that the reaction with DDQ (1 a) proceeds in a stepwise manner through hydrogen atom transfer, analogously to the oxidations of 1,4-dihydroarenes, whereas the reaction with thymoquinone 1 c is concerted, following the course of a pericyclic hydrogen transfer. The difference in the mechanistic courses of these two reactions may be explained by the effect of the CN and Cl substituents in 1 a, which stabilize a radical intermediate better than the alkyl groups in 1 c. The mechanistic conclusions are substantiated by DFT calculations.


Assuntos
Cloranila/análogos & derivados , Hidrogênio/química , Quinonas/síntese química , Benzoquinonas/síntese química , Benzoquinonas/química , Cloranila/síntese química , Cloranila/química , Espectroscopia de Ressonância de Spin Eletrônica , Modelos Moleculares , Estrutura Molecular , Oxirredução , Pressão , Quinonas/química
15.
J Hepatol ; 40(4): 607-15, 2004 Apr.
Artigo em Inglês | MEDLINE | ID: mdl-15030976

RESUMO

BACKGROUND/AIMS: We previously described that the cold-induced apoptosis of cultured hepatocytes and liver endothelial cells is mediated by an increase in the cellular chelatable iron pool-in the absence of any increase in O(2)(.-)/H(2)O(2) formation. As this is an unusual mechanism, we here set out to assess whether an increase in cellular chelatable iron per se is sufficient to trigger cell injury/apoptosis. METHODS: Cultured rat hepatocytes were acutely loaded with iron using the membrane-permeable complex Fe(III)/8-hydroxyquinoline and incubated under otherwise 'physiological' conditions. RESULTS: Incubation with Fe(III)/8-hydroxyquinoline (15 microM/30 microM) increased the cellular chelatable iron and induced strong hepatocellular injury with morphological features of apoptosis, but also of necrosis. The iron-induced cell injury was oxygen-dependent, and although it was not inhibitable by extracellular catalase, it was strongly inhibited by the novel membrane-permeable catalase mimic TAA-1/Fe. The experimentally induced increase in cellular chelatable iron triggered a mitochondrial permeability transition (MPT) as assessed using double-staining with calcein and tetramethylrhodamine methyl ester. The MPT inhibitor cyclosporine A partially and the well-known inhibitor combination trifluoperazine+fructose completely inhibited the iron-induced cell injury/apoptosis. CONCLUSIONS: These results show that iron per se can induce cell injury/apoptosis and that this injury is mediated via an MPT.


Assuntos
Hepatócitos/metabolismo , Ferro/metabolismo , Animais , Apoptose/efeitos dos fármacos , Células Cultivadas , Frutose/farmacologia , Hepatócitos/citologia , Hepatócitos/efeitos dos fármacos , Ferro/farmacologia , Mitocôndrias Hepáticas/efeitos dos fármacos , Mitocôndrias Hepáticas/metabolismo , Oxiquinolina/farmacologia , Permeabilidade/efeitos dos fármacos , Ratos , Espécies Reativas de Oxigênio/metabolismo
16.
Chem Biodivers ; 1(1): 65-97, 2004 Jan.
Artigo em Inglês | MEDLINE | ID: mdl-17191776

RESUMO

In view of the important role arginine plays in living organisms as the free amino acid and, especially, as a residue in peptides and proteins, the homologous beta-homoarginines are central in our investigations of beta-peptides (Fig. 1). The preparation of beta2-homoarginine derivatives suitably protected for solution- or solid-phase peptide syntheses is described with full experimental detail (9 and 12 in Scheme 1). The readily available Fmoc-beta3 hArg(Boc)2-OH is used for manual solid-phase synthesis of beta3-oligoarginines (on Rink amide or Rink amide AM resin) either by single amino acid coupling (Scheme 3) or, much better, by dimer-fragment coupling (Scheme 4). In this way, beta3-oligoarginine amides composed of 4, 6, 7, 8, and 10 residues, both with and without fluorescein labelling, were synthesized (Schemes 2-4), purified by preparative HPLC and identified by high-resolution mass spectrometry. The free amino acids (R)- and (S)-H-beta2 hArg-OH and (S)-H-beta3 hArg-OH were tested for their ability to function as substrates for NO synthase (iNOS); the beta3-oligoarginine amides (5, 6, and 7 residues) were tested for antibacterial (against six pathogens) and hemolytic (against rat and human erythrocytes) activities. All test results were negative: none of the free beta-homoarginines induced NO formation (Fig. 3), and there was no lysis of erythrocytes (concentrations up to 100 microM; Table 1), and no significant antibiotic activity (MIC > or = 64 microg/ml; Table 2). Cell-penetration studies with the fluorescence-labelled, peptidase-resistant beta3-oligoarginine amides were carried out with HeLa cells and human foreskin keratinocytes (HFKs). The results obtained with fluorescence microscopy are: i) the longer-chain beta-oligoarginine amides (8 and 10 residues; Figs. 4-6) enter the cells and end up in the nuclei, especially in the nucleoli, irrespective of temperature (37 degrees and 4 degrees with HFKs) or pretreatment with NaN3 (with HFKs), indicating a non-endocytotic and non-energy-dependent uptake mechanism; ii) the beta-tetraarginine derivative occupies the cell surface but does not enter the cells (with HeLa); iii) the cell-growth rate of the HFKs is not affected by a 1-microM concentration of the fluorescence-labelled beta-octaarginine amide (Fig. 7), i.e., there is no antiproliferative effect. In vivo experiments with mouse skin and the beta-octaarginine derivative show migration of the beta-peptide throughout the epidermis (Fig. 8). As a contribution to understanding the mechanism, we have also studied the behavior of fluorescence-labelled beta-octa- and beta-decaarginine amides (TFA salts) towards giant unilamellar vesicles (GUVs) built of neutral (POPC) or anionic (POPC/POPG mixtures) phospholipids: the beta-oligoarginine amides bind tightly to the surface of anionic GUVs but do not penetrate the lipid bilayer (Fig. 9) as they do with living cells. In contrast, a beta-heptapeptide FL-22, which had been used as a negative control sample for the cell-penetration experiments, entered the GUVs of negative surface charge. Thus, the mechanisms of cell and GUV-model penetration appear to be different. Finally, the possible applications and implications of the 'protein transduction' by beta-oligoarginines are discussed.


Assuntos
Homoarginina/química , Homoarginina/farmacologia , Animais , Arginina/química , Arginina/farmacologia , Células Cultivadas , Células HeLa , Humanos , Queratinócitos/citologia , Queratinócitos/efeitos dos fármacos , Queratinócitos/fisiologia , Masculino , Camundongos , Ratos
17.
Free Radic Biol Med ; 35(12): 1664-78, 2003 Dec 15.
Artigo em Inglês | MEDLINE | ID: mdl-14680689

RESUMO

We previously described that the cold-induced apoptosis of cultured hepatocytes is mediated by an increase in the cellular chelatable iron pool. We here set out to assess whether a mitochondrial permeability transition (MPT) is involved in cold-induced apoptosis. When cultured hepatocytes were rewarmed after 18 h of cold (4 degrees C) incubation in cell culture medium or University of Wisconsin solution, the vast majority of cells rapidly lost mitochondrial membrane potential. This loss was due to MPT as assessed by confocal laser scanning microscopy and as evidenced by the inhibitory effect of the MPT inhibitors trifluoperazine plus fructose. The occurrence of the MPT was iron-dependent: it was strongly inhibited by the iron chelators 2,2'-dipyridyl and deferoxamine. Addition of trifluoperazine plus fructose also strongly inhibited cold-induced apoptosis, suggesting that the MPT constitutes a decisive intermediate event in the pathway leading to cold-induced apoptosis. Further experiments employing the non-site-specific iron indicator Phen Green SK and specifically mitochondrial iron indicators and chelators (rhodamine B-[(1,10-phenanthrolin-5-yl)aminocarbonyl]benzyl ester, RPA, and rhodamine B-[(2,2'-bipyridin-4-yl)aminocarbonyl]benzyl ester, RDA) suggest that it is the cold-induced increase in cytosolic chelatable iron that triggers the MPT and that mitochondrial chelatable iron is not involved in this process.


Assuntos
Apoptose , Temperatura Baixa , Hepatócitos/citologia , Quelantes de Ferro/química , Ferro/metabolismo , Mitocôndrias/metabolismo , Animais , Células Cultivadas , Hepatócitos/metabolismo , Ferro/química , Masculino , Permeabilidade , Ratos , Ratos Wistar
18.
J Am Chem Soc ; 125(47): 14425-34, 2003 Nov 26.
Artigo em Inglês | MEDLINE | ID: mdl-14624590

RESUMO

The cycloaddition of thiobenzophenone S-methylide to thiobenzophenone, an experimentally well-known reaction, was studied, using (U)HF/3-21G* for finding stationary points and (U)B3LYP/6-31G*//(U)HF/3-21G* single-point calculations for energies. Some optimizations were performed by (U)B3LYP/ 6-31G* to check the reliability of the calculations. The comparison of the concerted pathways and stepwise reactions via C,C-biradicals and C,S-zwitterions showed that the formation of a tetraphenyl-substituted C,C-biradical and its ring closure to 4,4,5,5-tetraphenyl-1,3-dithiolane constitutes the energetically most probable pathway of product formation, despite the fact that the regioisomeric 2,2,4,4-tetraphenyl-substituted product is more favorable by 17 kcal mol(-1). Model calculations on bond dissociation energies showed that (U)B3LYP with various basis sets overestimates radical stabilization, whereas CBS-QB3 closely reproduced experimental values. Results with the BLYP functional are similar to those with B3LYP. The consequences of the overestimation of radical stability for the cycloaddition mechanism involving biradicals are discussed. Thiobenzophenone S-methylide, if not captured by a dipolarophile, dimerizes to 2,2,3,3-tetraphenyl-1,4-dithiane. Calculation disclosed likewise a tetraphenyl-substituted C,C-biradical as intermediate.

19.
Arch Biochem Biophys ; 418(2): 133-50, 2003 Oct 15.
Artigo em Inglês | MEDLINE | ID: mdl-14522585

RESUMO

The yields of nitrate and nitrite from decomposition of peroxynitrite in phosphate buffer at 37 degrees C were determined in the pH range 1-14. The NO(2)(-)/NO(3)(-) yields showed a stepwise variation with pH, with inflection points at approximately pH 3.1, 5.8, 6.8, 8.0, and 11.9. Nitrite formation increased strongly above pH 7 at the expense of nitrate, but above pH 12 nitrate again became the major product (80% at pH 14). At this pH, the Arrhenius parameters were E(a)=24.1+/-0.2kcal mol(-1) and A=(4.9+/-1.3)x10(12)s(-1). The yields of NO(2)(-), NO(3)(-), and O(2) measured at pH 5.8, 7.4, and 8.5 as a function of the initial peroxynitrite concentration (50-1000 microM) were linear only at pH 5.8. In the presence of carbon dioxide, oxygen production at pH 7.5 and pH 10 was found to be linear on the CO(2) concentration. The experimental observations were satisfactorily reproduced by kinetic simulations including principal component analyses. These data strongly suggest that the chemistry of peroxynitrite is exclusively mediated by z.rad;NO(2) and HO(z.rad;) radicals in the absence, and by z.rad;NO(2) and CO(3)(z.rad;-) radicals in the presence of CO(2).


Assuntos
Dióxido de Carbono/química , Modelos Químicos , Nitratos/química , Nitritos/química , Oxigênio/química , Ácido Peroxinitroso/química , Simulação por Computador , Radicais Livres/química , Concentração de Íons de Hidrogênio , Cinética , Reprodutibilidade dos Testes , Sensibilidade e Especificidade , Superóxidos/química , Temperatura
20.
J Biol Chem ; 278(47): 46403-13, 2003 Nov 21.
Artigo em Inglês | MEDLINE | ID: mdl-12963736

RESUMO

Enzymatic reduction of physiological Fe(III) complexes of the "labile iron pool" has not been studied so far. By use of spectrophotometric assays based on the oxidation of NAD(P)H and formation of [Fe(II) (1,10-phenanthroline)3]2+ as well as by utilizing electron paramagnetic resonance spectrometry, it was demonstrated that the NAD(P)H-dependent flavoenzyme lipoyl dehydrogenase (diaphorase, EC 1.8.1.4) effectively catalyzes the one-electron reduction of Fe(III) complexes of citrate, ATP, and ADP at the expense of the co-enzymes NAD(P)H. Deactivated or inhibited lipoyl dehydrogenase did not reduce the Fe(III) complexes. Likewise, in the absence of NAD(P)H or in the presence of NAD(P)+, Fe(III) reduction could not be detected. The fact that reduction also occurred in the absence of molecular oxygen as well as in the presence of superoxide dismutase proved that the Fe(III) reduction was directly linked to the enzymatic activity of lipoyl dehydrogenase and not mediated by O2. Kinetic studies revealed different affinities of lipoyl dehydrogenase for the reduction of the low molecular weight Fe(III) complexes in the relative order Fe(III)-citrate > Fe(III)-ATP > Fe(III)-ADP (half-maximal velocities at 346-485 microm). These Fe(III) complexes were enzymatically reduced also by other flavoenzymes, namely glutathione reductase (EC 1.6.4.2), cytochrome c reductase (EC 1.6.99.3), and cytochrome P450 reductase (EC 1.6.2.4) with somewhat lower efficacy. The present data suggest a (patho)physiological role for lipoyl dehydrogenase and other flavoenzymes in intracellular iron metabolism.


Assuntos
Difosfato de Adenosina/análogos & derivados , Di-Hidrolipoamida Desidrogenase/metabolismo , Flavoproteínas Transferidoras de Elétrons/metabolismo , Compostos Férricos/metabolismo , Difosfato de Adenosina/metabolismo , Trifosfato de Adenosina/metabolismo , Animais , Clostridium/enzimologia , Di-Hidrolipoamida Desidrogenase/química , Espectroscopia de Ressonância de Spin Eletrônica , Flavoproteínas Transferidoras de Elétrons/química , Cinética , Ligantes , Miocárdio/enzimologia , Análise Espectral , Suínos
SELEÇÃO DE REFERÊNCIAS
DETALHE DA PESQUISA
...