首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
The processes that are photoinduced by [Ru(bpz)(3)](2+) (bpz = 2,2'-bipyrazyl) in the presence of Cu/Zn superoxide dismutase (Cu/Zn SOD) are investigated by laser flash photolysis and electron paramagnetic resonance (EPR) spectroscopy; they are compared to those of the system [Ru(bpy)(3)(2+)-Cu/Zn SOD]. Although the mechanism is complicated, primary and secondary reactions can be evidenced. First, the excited [Ru(bpz)(3)](2+) complex is quenched reductively by Cu/Zn SOD with the production of a reduced complex and an oxidized enzyme. The oxidation site of Cu/Zn SOD is proposed to correspond to amino acids located on the surface of the protein. Afterward and only when this reductive electron transfer to the excited complex has produced enough oxidized protein, another electron-transfer process can be evidenced. In this case, however, the charge-transfer process takes place in the other direction, i.e., from the excited complex to the Cu(II) center of the SOD with the formation of Ru(III) and Cu(I) species. This proposed mechanism is supported by the fact that [Ru(bpy)(3)](2+), which is less photo-oxidizing than [Ru(bpz)(3)](2+), exhibits no photoreaction with Cu/Zn SOD. Because Ru(III) species are generated as intermediates with [Ru(bpz)(3)](2+), they are proposed to be responsible for the enhancement of [poly(dG-dC)](2) and [poly(dA-dT)](2) oxidation observed when Cu/Zn SOD is added to the [Ru(bpz)(3)](2+)-DNA system.  相似文献   

2.
Cu,Zn-superoxide dismutase (SOD1) is a copper- and zinc-dependent enzyme. The main function of SOD1 is believed to be the scavenging and detoxification of superoxide radicals. Nevertheless, the last 30 years have seen a rapid accumulation of evidence indicating that SOD1 may also act as a peroxidase, an alternative function that was implicated in the onset and progression of familial amyotrophic lateral sclerosis. Although SOD1 peroxidase activity and its dependence on carbon dioxide have been well described, the molecular basis of the SOD1 peroxidase cycle remains obscure, because none of the proposed catalytic intermediates have so far been identified. In view of recent observations, we hypothesized that the SOD1 peroxidase cycle relies on two steps: 1) reduction of SOD-Cu(II) by hydrogen peroxide followed by 2) oxidation of SOD-Cu(I) by peroxymonocarbonate, the product of the spontaneous reaction of bicarbonate with hydrogen peroxide, to produce SOD-Cu(II) and carbonate radical anion. This hypothesis has been investigated through electron paramagnetic resonance and nuclear magnetic resonance to provide direct evidence for a peroxycarbonate-driven, SOD1-catalyzed carbonate radical production. The results gathered herein indicate that peroxymonocarbonate () is a key intermediate in the SOD1 peroxidase cycle and identify this species as the precursor of carbonate radical anions.Cytosolic Cu,Zn-superoxide dismutase (SOD1)2 is a metal-dependent enzyme capable of accelerating the rate of spontaneous superoxide dismutation into O2 and H2O2 through the redox cycling of its copper ion (1, 2). SOD1 is widely distributed in mammalian cells and tissues and has been demonstrated to be located in the cytosol and in the intermembrane space of the mitochondria (see Ref. 3 and references therein). Because of that, SOD1 is believed to be a major player in the first line defense against reactive oxygen species, in particular superoxide anion.In addition to its dismutase activity, SOD1 possesses a well described but incompletely understood peroxidase activity which is dependent on hydrogen peroxide and markedly stimulated by small oxidizable anions such as nitrite and the ubiquitous carbon dioxide (3-12). The peroxidase activity of SOD1 has been proposed to impact the onset and progression of familial amyotrophic lateral sclerosis, a severely debilitating fatal disease characterized by the selective death of motor neurons (13-18). Although several reports exist in the literature indicating the formation of SOD1 aggregates and accumulation as a potential cause in the pathology progression, conflicting hypotheses are still under debate concerning the mechanisms that lead to the formation of SOD1-protein aggregates (19-21). Although some support the suggestion that free radical-induced covalent cross-links among SOD1 amino acids play a fundamental role in aggregate formation (22, 23), others support the view that metal loss from the enzyme structure leads to an unstable apo-form of SOD1 with increased capacity to form aggregates (24, 25). A detailed understanding of the SOD1 peroxidase cycle is essential to unraveling the mechanisms through which SOD1 aggregates are produced.The SOD1 peroxidase cycle is initiated when SOD1-Cu(II) is reduced by H2O2 or its deprotonated form (12), the peroxide anion (HOO-), to SOD1-Cu(I). This latter species is subsequently oxidized to a hypervalent intermediate (proposed to be either SOD1-Cu(III), SOD1-Cu(II)-·OH, or SOD1-Cu=O) (8, 9) that remains to be characterized. The reduction of this hypervalent intermediate by small anions is supposed to close the cycle, leading to the native enzyme and diffusible highly reactive radicals derived from the anionic substrates (6, 10).During its peroxidase cycle, a considerable fraction of SOD1 is inactivated due to the oxidation of the copper-binding histidines to oxohistidine, presumably by the hypervalent intermediate, in a process that can be prevented by the presence of reducing substrates and, in their absence, unavoidably leads to copper loss (26). Although this process is well described for the heme-dependent peroxidase cycle, current literature data (9, 27-30) and the fact that the proposed SOD1-bound hypervalent copper states (Cu(III), Cu(II)=O, and Cu(II)/.OH) have never been characterized suggested to us that an alternative mechanism may take place, leading to production from and H2O2 by the enzyme, a process that does not involve copper oxidation beyond the thermodynamically stable Cu(II) form. In the presence of , a significant fraction of H2O2 is promptly converted to through the perhydration of CO2 (27, 28, 31, 32). The peroxo bond in peroxymonocarbonate can be cleaved by reduced metals to produce and H2O (30 33) (Reactions 1-5),where Reactions 1 and 2 represent SOD1 reduction, Reactions 3 and 4 represent SOD1 oxidation, and Reaction 5 represents peroxycarbonate formation.Interestingly, studies employing molecular modeling of the SOD1 active site indicate that and H2O2 gain access to the SOD1 active site, where they react to produce in close proximity to the copper ion (29). This interaction of with Cu(I) may result in production.On the basis of these new data, we hypothesized that SOD1-Cu(I), which is the predominant form of SOD1 exposed to excess hydrogen peroxide (8, 9), is oxidized back to the native form of the enzyme by more efficiently than by H2O2 itself or HOO-. The latter two oxidations would slowly produce .OH radicals (or the equivalent SOD1-Cu(III), SOD1-Cu(II)=O, or Cu(II)/·OH) in the enzyme active site, leading to the observed inactivation of SOD1 (see Scheme 1). Here we present data that strongly support this hypothesis; they indicate that is a key substrate for reduced SOD1, which mediates SOD1-Cu(I) reoxidation back to the resting SOD1-Cu(II) severalfold faster than H2O2 itself and, in doing so, serves as the carbonate radical anion precursor.Open in a separate windowSCHEME 1.Schematic representation of the peroxidase catalytic cycle of Cu,Zn-SOD in the presence of /CO2. Native Cu,Zn-SOD is reduced by the peroxide anion, which gains access to the copper through the enzyme''s anion channel. Reduced Cu,Zn-SOD is reoxidized by the peroxycarbonate anion (), which is in equilibrium with H2O2//CO2, leading to carbonate radical anion production. Superoxide anion (), the product of HOO--induced SOD1 reduction, can oxidize SOD1-Cu(I) back to its resting SOD-Cu(II) state at diffusion-limited rates; however, it can alternatively reduce another molecule of SOD1-Cu(II) to SOD1-Cu(I), considerably accelerating the rate of SOD1-Cu(II) reduction by H2O2. Whether will act as a reductant of SOD1-Cu(II) or an oxidant of SOD1-Cu(I) will depend on the ratio of SOD1-Cu(II)/SOD1-Cu(I) at a given time, because the rate constants for the reaction of with both SOD1 states are close to the diffusion limits.  相似文献   

3.
A new quinolone-metal complex was prepared by a hydrothermal reaction in the presence of L-histidine that served as a reducing agent for a metal. The title compound [Cu(II)(cfH)(2)(Cu(I)Cl(2))(2)] (1) is a mixed-valence Cu(II)-Cu(I) complex, which contains two ciprofloxacin (cfH) molecules bonded to the central copper(II) atom and two almost planar [Cu(I)Cl(2)](-) moieties. Both metal centers are connected through two bridging atoms (chloride and quinolone oxygen). The electrochemical methods (differential-pulse polarography and cyclovoltammetric measurements) confirmed the presence of various copper-ciprofloxacin complex species in aqueous solution at low concentrations used in biological activity tests and also indicated that the equilibria in this system are very complex. The biological properties of the title compound and some previously isolated copper-ciprofloxacin complexes ([Cu(cfH)(2)Cl(2)].6H(2)O (2) and [CuCl(cfH)(phen)]Cl.2H(2)O (3)) (phen=1, 10-phenantroline) were determined and compared. The DNA gyrase inhibition tests and antibacterial activity tests have shown that the effect of copper complexes is comparable to that of free quinolone. Additionally, an interesting DNA cleavage activity of the title compound was also discovered.  相似文献   

4.
Tyrosinase (Ty) is a type 3 copper protein involved in the rate-limiting step of melanin synthesis. It is shown that the endogenous Trp fluorescence of tyrosinase from Streptomyces antibioticus is remarkably sensitive to the redox state. The fluorescence emission intensity of the [(Cu(I) Cu(I)] reduced species is more than twice that of the oxygen-bound [Cu(II)-O(2)(2-)-Cu(II)] form. The emission intensity of the oxidized [Cu(II)-OH(-)-Cu(II)] protein (Ty(met)) appears to be dependent on an acid-base equilibrium with a pK(a) value of 4.5 +/- 0.1. The binding of fluoride was studied under pseudo first-order conditions using stopped-flow fluorescence spectroscopy. The kinetic parameters k(on), K(d), and the fraction of fluorescence emission quenched upon fluoride binding show a similar pH dependence as above with an average pK(a) value of 4.62 +/- 0.05. Both observations are related to the dissociation of Cu(2)-bridging hydroxide at low pH. It is further shown that Ty is rapidly inactivated at low pH and that halide protects the enzyme from this inactivation. All results support the hypothesis that halide displaces hydroxide as the Cu(2)-bridging ligand in Ty(met). The relevance of the experimental findings for the catalytic cycle is discussed. The data are consistent with the data obtained from other techniques, validating the use of fluorescence quenching as a sensitive and effective tool in studying ligand binding and substrate conversion.  相似文献   

5.
Zhao X  Yeung N  Wang Z  Guo Z  Lu Y 《Biochemistry》2005,44(4):1210-1214
The electrochemical properties of an engineered heme-copper center in myoglobin have been investigated by UV-visible spectroelectrochemistry. In the cyanide-bridged, spin-coupled heme-copper center in an engineered myoglobin, the presence of Zn(II) in the Cu(B) center raises the heme reduction potential from -85 to 49 mV vs NHE. However, in the cyanide-free, spin-decoupled derivative of the same protein, the presence of Zn(II) in the Cu(B) center exerts little influence on the heme reduction potentials (77 and 80 mV vs NHE, respectively, in the absence and in the presence of Zn(II)). Similar trends have also been observed when copper ion is present in the Cu(B) center, although on a smaller scale, due to reduction of Cu(II) to Cu(I) prior to heme reduction. These results show that the presence of a metal ion in the designed Cu(B) center has a significant effect on the redox potential of heme Fe only when the two metal centers are coupled through a bridging ligand between the two metal centers, indicating that spin coupling plays an important role in redox potential regulation. In addition, the presence of a single positively charged Cu(I) center in the Cu(B) center resulted in a much lower increase (16 mV) in heme reduction potential than that of two positively charged Zn(II) (118 mV). Therefore, the heme reduction potential must be lowered after the first electron transfer to reduce heme Fe(3+)-Cu(B)(2+) to Fe(3+)-Cu(B)(+). To raise the heme reduction potential to make the second electron transfer (i.e., reduction of Fe(3+)-Cu(B)(+) to Fe(2+)-Cu(B)(+)) to be favorable, most likely a proton or decoupling of the heme-copper center is needed in the heme-copper site. These findings provide a strong argument for a thermodynamic driving force basis for redox-regulated proton transfer in heme-copper oxidases.  相似文献   

6.
Aberrant interactions of copper and zinc ions with the amyloid-beta peptide (Abeta) potentiate Alzheimer's disease (AD) by participating in the aggregation process of Abeta and in the generation of reactive oxygen species (ROS). The ROS production and the neurotoxicity of Abeta are associated with copper binding. Metallothionein-3 (Zn(7)MT-3), an intra- and extracellularly occurring metalloprotein, is highly expressed in the brain and downregulated in AD. This protein protects, by an unknown mechanism, cultured neurons from the toxicity of Abeta. Here, we show that a metal swap between Zn(7)MT-3 and soluble and aggregated Abeta(1-40)-Cu(II) abolishes the ROS production and the related cellular toxicity. In this process, copper is reduced by the protein thiolates forming Cu(I)(4)Zn(4)MT-3, in which an air-stable Cu(I)(4)-thiolate cluster and two disulfide bonds are present. The discovered protective effect of Zn(7)MT-3 from the copper-mediated Abeta(1-40) toxicity may lead to new therapeutic strategies for treating AD.  相似文献   

7.
The reactions of copper(II)-aliphatic polyamine complexes with cysteine, cysteine methyl ester, penicillamine, and glutathione have been investigated, with the goal of understanding the relationship between RS- -Cu(II) adduct structure and preferred redox decay pathway. Considerable mechanistic flexibility exists within this class of mercapto amino acid oxidations, as changes in the rate law could be induced by modest variations in reductant concentration (at fixed [Cu(II)]0), pH, and the structure of the redox partners. With excess cysteine present at 25 degrees C, pH 5.0, I = 0.2 M (NaOAc), decay of 1:1 cys-S- -Cu(II) transient adducts was found to be first order in both cys-SH and transient. Second-order rate constants characteristic of Cu(dien)2+(6.1 X 10(3) M-1 sec-1), Cu(Me5dien)2+ (2.7 X 10(3) M-1 sec-1), Cu(en)22+ (2.1 X 10(3) M-1 sec-1), and Cu(dien)22+ (4.7 X 10(3) M-1 sec-1) are remarkably similar, considering substantial differences in the composition and geometry of the oxidant first coordination sphere. A mechanism involving attack of cysteine on the coordinated sulfur atom of the transient, giving a disulfide anion radical intermediate, is proposed to account for these results. Moderate reactivity decreases in the cysteine-Cu(dien)2+, Cu(Me5dien)2+ reactions with increasing [H+] (pH 4-6) reflect partial protonation of the polyamine ligands. A very different rate law, second order in the RS- -Cu(II) transient and approximately zeroth order in mercaptan, applies in the pH 5.0 oxidations of cysteine methyl ester, penicillamine, and glutathione by Cu(dien)2+ and Cu(Me5dien)2+. This behavior suggests the intermediacy of di-mu-mercapto-bridged binuclear Cu(II) species, in which a concerted two-electron change yields the disulfide and Cu(I) products. Similar hydroxo-bridged intermediates are proposed to account for the transition from first- to second-order transient dependence in cysteine oxidations by Cu(dien)2+ and Cu(Me5dien)2+ as the pH is increased from 5 to 7. Yet another rate law, second order in transient and first order in cysteine, applies in the pH 5.0 oxidation of cysteine by Cu(Me6tren)2+ (k(25 degrees C) 7.5 X 10(7) M-2 sec-1, I = 0.2M). Steric rigidity of this trigonal bipyramidal oxidant evidently protects the coordinated sulfur atom from attack in a RSSR- -forming pathway. Formation of a coordinated disulfide in the rate-determining step is proposed, coupled with attack of a noncoordinated cysteine molecule on a vacated coordination position to stabilize the (Me6tren)Cu(I) product.  相似文献   

8.
Generation of superoxide during the enzymatic action of tyrosinase.   总被引:3,自引:0,他引:3  
Evidence for the generation of superoxide anion in an enzymatic action of tyrosinase is reported. In the dopatyrosinase reaction, 1 mol of O2 is required for the production of 2 mol of dopaquinone, 1 mol of dopachrome, and 1/4 mol of O2-. Superoxide dismutase and 2-methyl-6-phenyl-3,7-dihydroimidazo[1,2-a]pyrazin-3-one (a chemiluminescence probe and O2 trap) do not inhibit the rate of dopachrome formation from dopa in the presence of tyrosinase, indicating that free O2- is not utilized for metabolizing dopa. ESR studies for the accumulation of semiquinone radicals generated from tyrosine and N-acetyltyrosine in the presence of tyrosinase imply that O2- is not generated by the semiquinone + O2 reaction. Since the addition of H2O2 and dopa to tyrosinase promotes the release of O2- and formation of dopachrome, the Cu(II)O2-Cu(I) complex could be formed as a intermediate (an active form of tyrosinase); [Cu(II)]2 + H2O2 in equilibrium Cu(I)O2-Cu(II) + 2H+.  相似文献   

9.
In contrast with BLM(or TALM)-Cu(II) complex system, Cu(I)-O2 system of BLM(or TALM) as well as the corresponding Fe(II) system evidently produces reactive oxygen radicals as detected by ESR spin trapping. The sulfhydryl compound strongly prevented the generation of hydroxyl radical in BLM(or TALM)-Cu(I)-O2 system. TALM forms metal complexes similar to BLM. The action mechanism of BLM and TALM has been proposed to be substantially same.  相似文献   

10.
In this work, we investigated the oxidative modification of histidine residues induced by peroxidase and thiol oxidase activities of bovine copper-zinc superoxide dismutase (Cu-ZnSOD) using NMR and pulse EPR spectroscopy. 1D NMR and 2D-NOESY were used to determine the oxidative damage at the Zn(II) and Cu(II) active sites as well as at distant histidines. Results indicate that during treatment of SOD with hydrogen peroxide (H(2)O(2)) or cysteine in the absence of bicarbonate anion (HCO(3)(-)), both exchangeable and nonexchangeable protons were affected. Both His-44 and His-46 in the Cu(II) active site were oxidized based on the disappearance of NOESY cross-peaks between CH and NH resonances of the imidazole rings. In the Zn(II) site, only His-69, which is closer to His-44, was oxidatively modified. However, addition of HCO(3)(-) protected the active site His residues. Instead, resonances assigned to the His-41 residue, 11 ? away from the Cu(II) site, were completely abolished during both HCO(3)(-)-stimulated peroxidase activity and thiol oxidase activity in the presence of HCO(3)(-) . Additionally, ESEEM/HYSCORE and ENDOR studies of SOD treated with peroxide/Cys in the absence of HCO(3)(-) revealed that hyperfine couplings to the distal and directly coordinated nitrogens of the His-44 and His-46 ligands at the Cu(II) active site were modified. In the presence of HCO(3)(-), these modifications were absent. HCO(3)(-)-mediated, selective oxidative modification of histidines in SOD may be relevant to understanding the molecular mechanism of SOD peroxidase and thiol oxidase activities.  相似文献   

11.
Copper(II) ternary complexes based on the novel benzothiazole- N-sulfonamides, HL1 ( N-2-(4-methylbenzothiazole)benzenesulfonamide) and HL2 ( N-2-(6-nitrobenzothiazole)naphthalenesulfonamide) ligands, and pyridine have been synthesized and characterized. Complexes [Cu(L1)(2)(py)(2)] (1). and [Cu(L2)(2)(py)(2)] (2). were chemically characterized and their structures determined by means of single crystal X-ray analysis. In both compounds the Cu(II) ion is coordinated to four N atoms in a nearly square planar arrangement. The Cu-N bond distances are similar to those of Cu(2)Zn(2)SOD. The SOD mimetic activity of the complexes was determined both in vitro and in vivo. For determining the SOD-like activity of the complexes in vivo, we have developed a new method based on the complexes' protective effect on a delta sod1mutant of Saccharomyces cerevisiae against free radicals generated by hydrogen peroxide and menadione as well as free radicals produced in the cellular respiration process. The results have shown that complex 1 presents a protective action against oxidative stress induced by menadione or H(2)O(2) and that both complexes 1 and 2 protect against free radicals generated in cellular respiration.  相似文献   

12.
The complexation between copper(II) and the antihypertensive drug oxprenolol (HOxp) was studied both in methanol and slightly alkaline aqueous media at Cu:HOxp molar ratio from 1:1 to 1:10. Copper(lI) forms two types of complexes-a mononuclear violet one, CuOxp2, with bidentately bound ligands and a green dimeric one, Cu2Oxp2Cl2, in which the two Cu(II) centres are linked by the ligand through oxygen bridges. The crystal structure of the Cu2Oxp2Cl2 complex consists of two crystallographically non-equivalent centrosymmetric copper dimers. Each copper atom is four-coordinated in a distorted square-planar environment. The Cu2O2 structural core is characterized by a Cu1-O1-Cu1' angle of 104.15(13)degrees (Cu2-O2-Cu2' 104.30(13) degrees) and a relatively short Cu1-Cu1' separation of 3.026(1) A (Cu2-Cu2'-3.023(1) A). Magnetic susceptibility and EPR measurements indicate an antiferromagnetic coupling of the copper(II) centers.  相似文献   

13.
Using X-band electron paramagnetic resonance (EPR) and electron nuclear double resonance (ENDOR) spectroscopy at liquid helium temperatures, the Cu(II) coordination geometry at the active site of bovine and human copper,zinc-superoxide dismutases (bSOD1 and hSOD1) treated with H(2)O(2) and bicarbonate (HCO(3)(-)) was examined. The time course EPR of wild type human SOD1 (WT hSOD1), W32F hSOD1 mutant (tryptophan 32 substituted with phenylalanine), and bSOD1 treated with H(2)O(2) and HCO(3)(-) shows an initial reduction of active site Cu(II) to Cu(I) followed by its oxidation back to Cu(II) in the presence of H(2)O(2). However, HCO(3)(-) induced a Trp-32-derived radical from WT hSOD1 but not from bSOD1. The mutation of Trp-32 by phenylalanine totally eliminated the Trp-32 radical signal generated from W32F hSOD1 treated with HCO(3)(-) and H(2)O(2). Further characterization of the free radical was performed by UV irradiation of WT hSOD1 and bSOD1 that generated tryptophanyl and tyrosyl radicals. Both proton ((1)H) and nitrogen ((14)N) ENDOR studies of bSOD1 and hSOD1 in the presence of H(2)O(2) revealed a change in the geometry of His-46 (or His-44) and His-48 (or His-46) coordinated to Cu(II) at the active site of WT hSOD1 and bSOD1, respectively. However, in the presence of HCO(3)(-) and H(2)O(2), both (1)H and (14)N ENDOR spectra were almost identical to those derived from native bSOD1. We conclude that HCO(3)(-)-derived oxidant does not alter significantly the Cu(II) active site geometry and histidine coordination to Cu(II) in SOD1 as does H(2)O(2) alone; however, the oxidant derived from HCO(3)(-) (i.e. carbonate anion radical) reacts with surface-associated Trp-32 in hSOD1 to form the corresponding radical.  相似文献   

14.
The effect of Zn(2+) or Cu(2+) ions on Mn-depleted photosystem II (PS II) has been investigated using EPR spectroscopy. In Zn(2+)-treated and Cu(2+)-treated PS II, chemical reduction with sodium dithionite gives rise to a signal attributed to the plastosemiquinone, Q(A)(*)(-), the usual interaction with the non-heme iron being lost. The signal was identified by Q-band EPR spectroscopy which partially resolves the typical g-anisotropy of the semiquinone anion radical. Illumination at 200 K of the unreduced samples gives rise to a single organic free radical in Cu(2+)-treated PS II, and this is assigned to a monomeric chlorophyll cation radical, Chl a(*)(+), based on its (1)H-ENDOR spectrum. The Zn(2+)-treated PS II under the same conditions gives rise to two radical signals present in equal amounts and attributed to the Chl a(*)(+) and the Q(A)(*)(-) formed by light-induced charge separation. When the Cu(2+)-treated PS II is reduced by sodium ascorbate, at >/=77 K electron donation eliminates the donor-side radical leaving the Q(A)(*)(-) EPR signal. The data are explained as follows: (1) Cu(2+) and Zn(2+) have similar effects on PS II (although higher concentrations of Zn(2+) are required) causing the displacement of the non-heme Fe(2+). (2) In both cases chlorophyll is the electron donor at 200 K. It is proposed that the lack of a light-induced Q(A)(*)(-) signal in the unreduced Cu(2+)-treated sample is due to Cu(2+) acting as an electron acceptor from Q(A)(*)(-) at low temperature, forming the Cu(+) state and leaving the electron donor radical Chl a(*)(+) detectable by EPR. (3) The Cu(2+) in PS II is chemically reducible by ascorbate prior to illumination, and the metal can therefore no longer act as an electron acceptor; thus Q(A)(*)(-) is generated by illumination in such samples. (4) With dithionite, both the Cu(2+) and the quinone are reduced resulting in the presence of Q(A)(*)(-) in the dark. The suggested high redox potential of Cu(2+) when in the Fe(2+) site in PS II is in contrast to the situation in the bacterial reaction center where it has been shown in earlier work that the Cu(2+) is unreduced by dithionite. It cannot be ruled out however that Q(A)-Cu(2+) is formed and a magnetic interaction is responsible for the lack of the Q(A)(-) signal when no exogenous reductant is present. With this alternative possibility, the effects of reductants would be explained as the loss of Cu(2+) (due to formation of Cu(+)) leading to loss of the Cu(2+) from the Fe(2+) site due to the binding equilibrium. The quite different binding and redox behavior of the metal in the iron site in PS II compared to that of the bacterial reaction center is presumably a further reflection of the differences in the coordination of the iron in the two systems.  相似文献   

15.
The known action of Cu, Zn superoxide dismutase (Cu(2)Zn(2)SOD) that converts O(2)(-) to O(2) and H(2)O(2) plays a crucial role in protecting cells from toxicity of oxidative stress. However, the overproduction of Cu(2)Zn(2)SOD does not result in increased protection but rather creates a variety of unfavorable effects, suggesting that too much Cu(2)Zn(2)SOD may be injurious to the cells. The present study examined the DNA cleavage activity mediated by a Cu(n)SOD that contains 1-4 copper ions, in order to obtain an insight into the aberrant copper-mediated oxidative chemistry in the enzyme. A high SOD activity was observed upon metallation of the apo-form of Cu(2)Zn(2)SOD with Cu(II), indicating that nearly all of the Cu(II) in the Cu(n)SOD is as active as the Cu(II) in the copper site of fully active Cu(2)Zn(2)SOD. Using a supercoiled DNA as substrate, significant DNA cleavage was observed with the Cu(n)SOD in the presence of hydrogen peroxide or mercaptoethanol, whereas DNA cleavage with free Cu(II) ions can occur only <5% under the same conditions. Comparison with other proteins shows that the DNA cleavage activity is specific to some proteins including the Cu(n)SOD. The steady state study suggests that a cooperative action between the SOD protein and the Cu(II)may appear in the DNA cleavage activity, which is independent of the number of Cu(II) in the Cu(n)SOD. The kinetic study shows that a two-stage reaction was involved in DNA cleavage. The effects of various factors including EDTA, radical scavengers, bicarbonate anion, and carbon dioxide gas molecules on the Cu(n)SOD-mediated DNA cleavage activity were also investigated. It is proposed that DNA cleavage occurs via both hydroxyl radical oxidation and hydroxide ion hydrolysis pathways. This work implies that any form of the copper-containing SOD enzymes (including Cu(2)Zn(2)SOD and its mutants) might have the DNA cleavage activity.  相似文献   

16.
Cu,Zn SOD, but not Mn SOD, catalyzes the oxidation of 3-hydroxyanthranilic acid (3-HA) under aerobic conditions. In the absence of O2, the Cu(II) of the enzyme is reduced by 3-HA. One plausible mechanism involves the reduction of the active site Cu(II) to Cu(I), which is then reoxidized by the O2- generated by autoxidation of the anthranilyl or other radicals on the pathway to cinnabarinate. We may call this the superoxide reductase, or SOR, mechanism. Another possibility invokes direct reoxidation of the active site Cu(I) by the anthranilyl, or other organic radicals, or by the peroxyl radicals generated by addition of O2 to these organic radicals. Such oxidations catalyzed by Cu,Zn SOD could account for the deleterious effects of the mutant Cu,Zn SODs associated with familial amyotrophic lateral sclerosis and of the overproduction or overadministration of wild-type Cu,Zn SOD.  相似文献   

17.
Two Cu(II) complexes with cyanoguanidine (cnge) and o-phenanthroline, [Cu(o-phen)(2)(cnge)](NO(3))(2).2H(2)O (1) and [Cu(o-phen)(cnge)(H(2)O)(NO(3))(2)] (2), have been synthesized using different experimental techniques and characterized by elemental analyses, FTIR, diffuse and UV-vis spectra and EPR and magnetic moment measurements techniques. The crystal structures of both complexes were solved by X-ray diffraction methods. Complex (1) crystallizes in the monoclinic space group C2/c with a=12.621(5), b=31.968(3), c=15.39(1)A, beta=111.68(4) degrees, and Z=8 and complex (2) in the monoclinic space group P2(1)/n with a=10.245(1), b=13.923(2), c=12.391(2)A, beta=98.07(1) degrees, and Z=4. The environments of the copper(II) center are trigonal bipyramidal (TBP) for [Cu(o-phen)(2)(cnge)](2+) and an elongated octahedron for [Cu(o-phen)(cnge)(H(2)O)(NO(3))(2)]. Solution studies have been performed to determine the species distribution. The superoxide dismutase (SOD) activities of both complexes have also been tested in order to determine if these compounds mimic the enzymatic action of the enzyme SOD that protects cells against peroxide radicals.  相似文献   

18.
Dicopper complexes of the following benzimidazole-containing ligands have been studied as possible models for the active site of hemocyanin: EDTB (N,N,N',N'-tetrakis-(2-benzimidazolylmethyl)-1,2-ethanediamine), EGTB (1,1,10,10-tetrakis-(2-benzimidazolylmethyl)-1,10-diaza-4,7- dioxadecane), and MEGTB (1,1,10,10-tetrakis-(1-methylbenzimidazol-2-y lmethyl)-1,10-diaza-4,7-dioxadecane). The initial oxygenation product of Cu2(EDTB)(ClO4)2 in Me2SO gives optical absorption maxima at 315 nm (epsilon = 3750 M-1 cm-1) and 690 nm (epsilon = 100 M-1 cm-1). The fluorescence emission intensities of Cu2(EDTB)(ClO4)2 at 400 and 700 nm (excitation at 350 nm) decreases rapidly on exposure to air. This suggests oxidation of Cu2(I) to Cu2(II). The x-ray absorption edge spectra suggest that both coppers in the oxygenation product, analyzed as Cu2(EDTB)(ClO4)2(O).3H2O, are Cu(II). From spectrophotometric titration of Cu2(MEGTB)Cl4 with azide, formation constant of the Cu2(MEGTB)N3Cl3 complex has been obtained. Data from cyclic voltammetry experiments suggest that in the presence of azide, Cu(II)(N3)Cu(II) species is present.  相似文献   

19.
The reactivity of nitrite towards the copper(II) and copper(I) centers of a series of complexes with tridentate nitrogen donor ligands has been investigated. The ligands are bis[(1-methylbenzimidazol-2-yl)methyl]amine (1-bb), bis[2-(1-methylbenzimidazol-2-yl)ethyl]amine (2-bb), and bis[2-(3,5-dimethyl-1-pyrazolyl)ethyl]amine (ddah) and carry two terminal benzimidazole (1-bb, 2-bb) or pyrazole (ddah) rings and a central amine donor residue. While 2-bb and ddah form two adjacent six-membered chelate rings on metal coordination, 1-bb forms two smaller rings of five members. The binding affinity of nitrite and azide to the Cu(II) complexes (ClO4 as counterion) has been determined in solution. The association constants for the two ligands are similar, but nitrite is a slightly stronger ligand than azide when it binds as a bidentate donor. The X-ray crystal structure of the nitrite complex [Cu(ddah)(NO2)]ClO4 (final R=0.056) has been determined: triclinic P1ˉspace group, a=8.200(2) ?, b=9.582(3) ?, c=15.541(4) ?. It may be described as a perchlorate salt of a “supramolecular” species resulting from the assembly of two complex cations and one sodium perchlorate unit. The copper stereochemistry in the complex is intermediate between SPY and TBP, and nitrite binds to Cu(II) asymmetrically, with Cu-O distances of 2.037(2) and 2.390(3) ? and a nearly planar CuO2N cycle. On standing, solutions of [Cu(ddah)(NO2)]ClO4 in methanol produce the dinuclear complex [Cu(ddah)(OMe)]2(ClO4)2, containing dibridging methoxy groups. In fact the crystal structure analysis (final R=0.083) showed that the crystals are built up by dinuclear cations, arranged on a crystallographic symmetry center, and perchlorate anions. Electrochemical analysis shows that binding of nitrite to the Cu(II) complexes of 2-bb and ddah shifts the reduction potential of the Cu(II)/Cu(I) couple towards negative values by about 0.3 V. The thermodynamic parameters of the Cu(II)/Cu(I) electron transfer have also been analyzed. The mechanism of reductive activation of nitrite to nitric oxide by the Cu(I) complexes of 1-bb, 2-bb, and ddah has been studied. The reaction requires two protons per molecule of nitrite and Cu(I). Kinetic experiments show that the reaction is first order in [Cu(I)] and [H+] and exhibits saturation behavior with respect to nitrite concentration. The kinetic data show that [Cu(2-bb)]+ is more efficient than [Cu(1-bb)]+ and [Cu(ddah)]+ in reducing nitrite. Received: 19 November 1999 / Accepted: 20 January 2000  相似文献   

20.
A site-directed mutant of spinach plastocyanin, Pc(Tyr83-His), has been modified by covalent attachment of a photoactive [Ru(bpy)2(im)]2+ complex to the His83 residue. The residue is surface exposed and located about 10–12?Å from the copper ion at the entrance of a proposed natural electron transfer pathway from cytochrome f. Electron transfer within the Ru-Pc complex has been studied with time-resolved optical spectroscopy using two different approaches. In the first, the fully reduced [Cu(I), Ru(II)] protein was photoexcited and subsequently oxidized by an external quencher, forming the [Cu(I), Ru(III)] protein. This was followed by an electron transfer from reduced Cu(I) to Ru(III). In the second method, the initially oxidized Cu(II) ion acted as an internal quencher for excited Ru(II) and the photoinduced reduction of the Cu(II) ion was followed by a thermal recombination with the Ru(III) ion. The reoxidation of the Cu ion, which has an estimated driving force of 0.56?eV, occured with a rate constant k et?=?(9.5±1.0)×106?s–1, observed with both methods. The results suggest a strong electronic coupling (H DA>0.3?cm–1) along the Ru-His(83)-Cys(84)-Cu pathway.  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