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 822 毫秒
1.
Identification of the intermediates and determination of their structures in the reduction of dioxygen to water by cytochrome c oxidase (CcO) are particularly important to understanding both O2 activation and proton pumping by the enzyme. In this work, we report the products of the rapid reaction of O2 with the mixed valence form (CuA2+, heme a3+, heme a32+-CuB1+) of the enzyme. The resonance Raman results show the formation of two ferryl-oxo species with characteristic Fe(IV)=O stretching modes at 790 and 804 cm−1 at the peroxy oxidation level (PM). Density functional theory calculations show that the protein environment of the proximal H-bonded His-411 determines the strength of the distal Fe(IV)=O bond. In contrast to previous proposals, the PM intermediate is also formed in the reaction of Y167F with O2. These results suggest that in the fully reduced enzyme, the proton pumping νFe(IV)=O = 804 cm−1 to νFe(IV)=O = 790 cm−1 transition (P→F, where P is peroxy and F is ferryl) is triggered not only by electron transfer from heme a to heme a3 but also by the formation of the H-bonded form of the His-411-Fe(IV)=O conformer in the proximal site of heme a3. The implications of these results with respect to the role of an O=Fe(IV)-His-411-H-bonded form to the ring A propionate of heme a3-Asp-399-H2O site and, thus, to the exit/output proton channel (H2O) pool during the proton pumping P→F transition are discussed. We propose that the environment proximal to the heme a3 controls the spectroscopic properties of the ferryl intermediates in cytochrome oxidases.  相似文献   

2.
The R2 subunit of Escherichia coli ribonucleotide reductase contains a diiron site that reacts with O2 to produce a tyrosine radical (Y122·). In wild-type R2 (R2-wt), the first observable reaction intermediate is a high-valent [FeIII-FeIV] state called compound X, but in related diiron proteins such as methane monooxygenase, 9-desaturase, and ferritin, peroxodiiron(III) complexes have been characterized. Substitution of iron ligand D84 by E within the active site of R2 allows an intermediate (-1,2-peroxo)diiron species to accumulate. To investigate the possible involvement of a bridging peroxo species within the O2 activation sequence of R2-wt, we have characterized the iron-nitrosyl species that form at the diiron sites in R2-wt, R2-D84E, and R2-W48F/D84E by using vibrational spectroscopy. Previous work has shown that the diiron center in R2-wt binds one NO per iron to form an antiferromagnetically coupled [{FeNO}7]2 center. In the wt and variant proteins, we also observe that both irons bind one NO to form a {FeNO}7 dimer where both Fe–N–O units share a common vibrational signature. In the wt protein, (Fe–NO), (Fe–N–O), and (N–O) bands are observed at 445, 434 and 1742 cm–1, respectively, while in the variant proteins the (Fe–NO) and (Fe–N–O) bands are observed ~10 cm–1 higher and the (N–O) ~10 cm–1 lower at 1735 cm–1. These results demonstrate that all three proteins accommodate fully symmetric [{FeNO}7]2 species with two identical Fe–N–O units. The formation of equivalent NO adducts in the wt and variant proteins strongly favors the formation of a symmetric bridging peroxo intermediate during the O2 activation process in R2-wt.  相似文献   

3.
Thirteen different polypeptide subunits, each in one copy, five phosphatidyl ethanolamines and three phosphatidyl glycerols, two hemes A, three Cu ions, one Mg ion, and one Zn ion are detectable in the crystal structure of bovine heart cytochrome c oxidase in the fully oxidized form at 2.8 Å resolution. A propionate of hems a, a peptide unit (–CO–NH–), and an imidazole bound to CuA are hydrogen-bonded sequentially, giving a facile electron transfer path from CuA to heme a. The O2 binding and reduction site, heme a 3, is 4.7 Å apart from CuB. Two possible proton transfer paths from the matrix side to the cytosolic side are located in subunit I, including hydrogen bonds and internal cavities likely to contain randomly oriented water molecules. Neither path includes the O2 reduction site. The O2 reduction site has a proton transfer path from the matrix side possibly for protons for producing water. The coordination geometry of CuB and the location of Tyr244 in subunit I at the end of the scalar proton path suggests a hydroperoxo species as the two electron reduced intermediate in the O2 reduction process.  相似文献   

4.
IsdGs are heme monooxygenases that break open the tetrapyrrole, releasing the iron, and thereby allowing bacteria expressing this protein to use heme as a nutritional iron source. Little is currently known about the mechanism by which IsdGs degrade heme, although the products differ from those generated by canonical heme oxygenases. A synthesis of time-resolved techniques, including in proteo mass spectrometry and conventional and stopped-flow UV/visible spectroscopy, was used in conjunction with analytical methods to define the reaction steps mediated by IsdG from Staphylococcus aureus and their time scales. An apparent meso-hydroxyheme (forming with k = 0.6 min−1, pH 7.4, 10 mm ascorbate, 10 μm IsdG-heme, 22 °C) was identified as a likely common intermediate with the canonical heme oxygenases. Unlike heme oxygenases, this intermediate does not form with added H2O2 nor does it convert to verdoheme and CO. Rather, the next observable intermediates (k = 0.16 min−1) were a set of formyloxobilin isomers, similar to the mycobilin products of the IsdG homolog from Mycobacterium tuberculosis (MhuD). These converted in separate fast and slow phases to β-/δ-staphylobilin isomers and formaldehyde (CH2O). Controlled release of this unusual C1 product may support IsdG''s dual role as both an oxygenase and a sensor of heme availability in S. aureus.  相似文献   

5.
The mechanism for the reaction between nitric oxide (NO) and O2 bound to the heme iron of myoglobin (Mb), including the following isomerization to nitrate, has been investigated using hybrid density functional theory (B3LYP). Myoglobin working as a NO scavenger could be of importance, since NO reversibly inhibits the terminal enzyme in the respiration chain, cytochrome c oxidase. The concentration of NO in the cell will thus affect the respiration and thereby the synthesis of ATP. The calculations show that the reaction between NO and the heme-bound O2 gives a peroxynitrite intermediate whose O–O bond undergoes a homolytic cleavage, forming a NO2 radical and myoglobin in the oxo-ferryl state. The NO2 radical then recombines with the oxo-ferryl, forming heme-bound nitrate. Nine different models have been used in the present study to examine the effect on the reaction both by the presence and the protonation state of the distal His64, and by the surroundings of the proximal His93. The barriers going from the oxy-Mb and nitric oxide reactant to the peroxynitrite intermediate and further to the oxo-ferryl and NO2 radical are around 10 and 7 kcal/mol, respectively. Forming the product, nitrate bound to the heme iron has a barrier of less than ~7 kcal/mol. The overall reaction going from a free nitric oxide and oxy-Mb to the heme bound nitrate is exergonic by more than 30 kcal/mol.  相似文献   

6.
Inhibition of terminal oxidases by nitric oxide (NO) has been extensively investigated as it plays a role in regulation of cellular respiration and pathophysiology. Cytochrome bd is a tri-heme (b558, b595, d) bacterial oxidase containing no copper that couples electron transfer from quinol to O2 (to produce H2O) with generation of a transmembrane protonmotive force. In this work, we investigated by stopped-flow absorption spectroscopy the reaction of NO with Escherichia coli cytochrome bd in the fully oxidized (O) state. We show that under anaerobic conditions, the O state of the enzyme binds NO at heme d with second-order rate constant kon = 1.5 ± 0.2 × 102 M−1 s−1, yielding a nitrosyl adduct (d3+–NO or d2+–NO+) with characteristic optical features (an absorption increase at 639 nm and a red shift of the Soret band). The reaction mechanism is remarkably different from that of O cytochrome c oxidase in which the heme–copper binuclear center reacts with NO approximately three orders of magnitude faster, forming nitrite. The data allow us to conclude that in the reaction of NO with terminal oxidases in the O state, CuB is indispensable for rapid oxidation of NO into nitrite.  相似文献   

7.
《Luminescence》2003,18(5):259-267
High‐valent oxo‐iron(IV) species are commonly proposed as the key intermediates in the catalytic mechanisms of iron enzymes. Water‐soluble iron(III) tetrakis‐5,10,15,20‐(N‐methyl‐4‐pyridyl)porphyrin (Fe(III)TMPyP) has been used as a model of heme‐enzyme to catalyse the hydrogen peroxide (H2O2) oxidation of various organic compounds. However, the mechanism of the reaction of Fe(III)TMPyP with H2O2 has not been fully established. In this study, we have explored the kinetic simulation of the reaction of Fe(III)TMPyP with H2O2 and of the catalytic reactivity of FeTMPyP in the luminescent peroxidation of luminol. According to the mechanism that has been established in this work, Fe(III)TMPyP is oxidized by H2O2 to produce (TMPyP)·+Fe(IV)=O (k1 = 4.5 × 104/mol/L/s) as a precursor of TMPyPFe(IV)=O. The intermediate, (TMPyP)·+Fe(IV)=O, represented nearly 2% of Fe(III)TMPyP but it does not accumulate in suf?cient concentration to be detected because its decay rate is too fast. Kinetic simulations showed that the proposed scheme is capable of reproducing the observed time courses of FeTMPyP in various oxidation states and the decay pro?les of the luminol chemiluminescence. It also shows that (TMPyP)·+Fe(IV)=O is 100 times more reactive than TMPyPFe(IV)=O in most of the reactions. These two species are responsible for the initial sharp and the sustained luminol emissions, respectively. Copyright © 2003 John Wiley & Sons, Ltd.  相似文献   

8.
Human cystathionine β-synthase (CBS), a novel heme-containing pyridoxal 5′-phosphate enzyme, catalyzes the condensation of homocysteine and serine or cysteine to produce cystathionine and H2O or H2S, respectively. The presence of heme in CBS has limited spectrophotometric characterization of reaction intermediates by masking the absorption of the pyridoxal 5′-phosphate cofactor. In this study, we employed difference stopped-flow spectroscopy to characterize reaction intermediates formed under catalytic turnover conditions. The reactions of l-serine and l-cysteine with CBS resulted in the formation of a common aminoacrylate intermediate (kobs = 0.96 ± 0.02 and 0.38 ± 0.01 mm−1 s−1, respectively, at 24 °C) with concomitant loss of H2O and H2S and without detectable accumulation of the external aldimine or other intermediates. Homocysteine reacted with the aminoacrylate intermediate with kobs = 40.6 ± 3.8 s−1 and re-formed the internal aldimine. In the reverse direction, CBS reacted with cystathionine, forming the aminoacrylate intermediate with kobs = 0.38 ± 0.01 mm−1 s−1. This study provides the first insights into the pre-steady-state kinetic mechanism of human CBS and indicates that the reaction is likely to be limited by a conformational change leading to product release.  相似文献   

9.
The amino terminus of bovine pro-opiomelanocortin (N-POMC1–77) is partially processed in the intermediate lobe of the pituitary to N-POMC1–49 and lys-γ3 -melanotropin. Two pools of N-POMC1–77 were isolated which were differentially glycosylated at threonine45, while N-POMC1–49 isolated from bovine intermediate lobe extracts existed in a non-glycosylated form. This suggested that differential O-linked glycosylation of N-POMC1–77 may regulate cleavage at the Arg49-Lys50 processing site. We tested this hypothesis by incubating N-POMC1–77 glycoforms with purified pro-opiomelanocortin converting enzyme. Only non-O-glycosylated N-POMC1–77 and O-glycosylated N-POMC1–77 with truncated oligosaccharide sidechains were sensitive to cleavage and generated predominantly lys-γ3 -melanotropin, identified by high-performance liquid chromatography. These data provide the first functional evidence to support a role for differential O-linked glycosylation in the regulation of the processing of the N-terminus of bovine POMC.  相似文献   

10.
Magnetite-producing magnetotactic bacteria collected from the oxic–anoxic transition zone of chemically stratified marine environments characterized by O2/H2S inverse double gradients, contained internal S-rich inclusions resembling elemental S globules, suggesting they oxidize reduced S compounds that could support autotrophy. Two strains of marine magnetotactic bacteria, MV-1 and MV-2, isolated from such sites grew in O2-gradient media with H2S or thiosulfate (S2O32–) as electron sources and O2 as electron acceptor or anaerobically with S2O32– and N2O as electron acceptor, with bicarbonate (HCO3)/CO2 as sole C source. Cells grown with H2S contained S-rich inclusions. Cells oxidized S2O32– to sulfate (SO42–). Both strains grew microaerobically with formate. Neither grew microaerobically with tetrathionate (S4O62–), methanol, or Fe2+ as FeS, or siderite (FeCO3). Growth with S2O32– and radiolabeled 14C-HCO3 showed that cell C was derived from HCO3/CO2. Cell-free extracts showed ribulose 1,5-bisphosphate carboxylase/oxygenase (RubisCO) activity. Southern blot analyses indicated the presence of a form II RubisCO (cbbM) but no form I (cbbL) in both strains. cbbM and cbbQ, a putative post-translational activator of RubisCO, were identified in MV-1. MV-1 and MV-2 are thus chemolithoautotrophs that use the Calvin–Benson–Bassham pathway. cbbM was also identified in Magnetospirillum magnetotacticum. Thus, magnetotactic bacteria at the oxic–anoxic transition zone of chemically stratified aquatic environments are important in C cycling and primary productivity.  相似文献   

11.
The potential for C4 photosynthesis was investigated in five C3-C4 intermediate species, one C3 species, and one C4 species in the genus Flaveria, using 14CO2 pulse-12CO2 chase techniques and quantum-yield measurements. All five intermediate species were capable of incorporating 14CO2 into the C4 acids malate and aspartate, following an 8-s pulse. The proportion of 14C label in these C4 products ranged from 50–55% to 20–26% in the C3-C4 intermediates F. floridana Johnston and F. linearis Lag. respectively. All of the intermediate species incorporated as much, or more, 14CO2 into aspartate as into malate. Generally, about 5–15% of the initial label in these species appeared as other organic acids. There was variation in the capacity for C4 photosynthesis among the intermediate species based on the apparent rate of conversion of 14C label from the C4 cycle to the C3 cycle. In intermediate species such as F. pubescens Rydb., F. ramosissima Klatt., and F. floridana we observed a substantial decrease in label of C4-cycle products and an increase in percentage label in C3-cycle products during chase periods with 12CO2, although the rate of change was slower than in the C4 species, F. palmeri. In these C3-C4 intermediates both sucrose and fumarate were predominant products after a 20-min chase period. In the C3-C4 intermediates, F. anomala Robinson and f. linearis we observed no significant decrease in the label of C4-cycle products during a 3-min chase period and a slow turnover during a 20-min chase, indicating a lower level of functional integration between the C4 and C3 cycles in these species, relative to the other intermediates. Although F. cronquistii Powell was previously identified as a C3 species, 7–18% of the initial label was in malate+aspartate. However, only 40–50% of this label was in the C-4 position, indicating C4-acid formation as secondary products of photosynthesis in F. cronquistii. In 21% O2, the absorbed quantum yields for CO2 uptake (in mol CO2·[mol quanta]-1) averaged 0.053 in F. cronquistii (C3), 0.051 in F. trinervia (Spreng.) Mohr (C4), 0.052 in F. ramosissima (C3-C4), 0.051 in F. anomala (C3-C4), 0.050 in F. linearis (C3-C4), 0.046 in F. floridana (C3-C4), and 0.044 in F. pubescens (C3-C4). In 2% O2 an enhancement of the quantum yield was observed in all of the C3-C4 intermediate species, ranging from 21% in F. ramosissima to 43% in F. pubescens. In all intermediates the quantum yields in 2% O2 were intermediate in value to the C3 and C4 species, indicating a co-function of the C3 and C4 cycles in CO2 assimilation. The low quantum-yield values for F. pubescens and F. floridana in 21% O2 presumably reflect an ineffcient transfer of carbon from the C4 to the C3 cycle. The response of the quantum yield to four increasing O2 concentrations (2–35%) showed lower levels of O2 inhibition in the C3-C4 intermediate F. ramosissima, relative to the C3 species. This indicates that the co-function of the C3 and C4 cycles in this intermediate species leads to an increased CO2 concentration at the site of ribulose-1,5-bisphosphate carboxylase/oxygenase and a concomitant decrease in the competitive inhibition by O2.Abbreviations PEP phosphoenolpyruvate - PGA 3-phosphoglycerate - RuBP ribulose-1,5-bisphosphate  相似文献   

12.
Many bioactive peptides, such as hormones and neuropeptides, require amidation at the C terminus for their full biological activity. Peptidylglycine α-hydroxylating monooxygenase (PHM) performs the first step of the amidation reaction—the hydroxylation of peptidylglycine substrates at the Cα position of the terminal glycine. The hydroxylation reaction is copper- and O2-dependent and requires 2 equiv of exogenous reductant. The proposed mechanism suggests that O2 is reduced by two electrons, each provided by one of two nonequivalent copper sites in PHM (CuH and CuM). The characteristics of the reduced oxygen species in the PHM reaction and the identity of the reactive intermediate remain uncertain. To further investigate the nature of the key intermediates in the PHM cycle, we determined the structure of the oxidized form of PHM complexed with hydrogen peroxide. In this 1.98-Å-resolution structure (hydro)peroxide binds solely to CuM in a slightly asymmetric side-on mode. The O–O interatomic distance of the copper-bound ligand is 1.5 Å, characteristic of peroxide/hydroperoxide species, and the Cu–O distances are 2.0 and 2.1 Å. Density functional theory calculations using the first coordination sphere of the CuM active site as a model system show that the computed energies of the side-on L3CuM(II)–O2 2? species and its isomeric, end-on structure L3CuM(I)–O2 ·? are similar, suggesting that both these intermediates are significantly populated within the protein environment. This observation has important mechanistic implications. The geometry of the observed side-on coordinated peroxide ligand in L3CuM(II)O2 2? is in good agreement with the results of a hybrid quantum mechanical–molecular mechanical optimization of this species.  相似文献   

13.
Solid dipyridine hemes which are unreactive toward oxygen lose both pyridine ligands upon heating under vacuum to give a solid which takes up O2, reversibly, one O2 per heme. Replacement of 16O2 by 18O2 reduces only infrared bands near 1660 and 1590 cm?1, frequencies near the vibrational band for gaseous O2. No FeO bands are detected. EPR spectra reveal a free radical and ferric iron; Mössbauer, NMR and infrared spectra support an iron(III) oxidation state. Limited molecular weight data indicate a dimer. Possibly two dioxygen molecules are held sandwich fashion between two porphyrins via donor-acceptor interactions, which are facilitated by electron transfer from iron(II) into the porphyrin forming a π-anion. Such O2 bonding is not found in oxy Hb and Mb or in oxyhemerythrin but may occur with cytochrome c oxidase and other oxygen utilizing (or producing) heme and other proteins.  相似文献   

14.
1H NMR spectroscopy was applied to study the reactions of cis-[Pd(L)(H2O)2]2+ complexes (L is en, pic and dpa) with the N-acetylated tripeptides L-methionylglycylglycine, MeCOMet–Gly–Gly, and glycyl–L-methionyl–glycine, MeCOGly–Met–Gly. All reactions were performed in the pH range 2.0–2.5 with equimolar amounts of the cis-[Pd(L)(H2O)2]2+ complex and the tripeptide at 60 °C. The hydrolytic reactions of the cis-[Pd(en)(H2O)2]2+, cis-[Pd(pic)(H2O)2]2+ and cis-[Pd(dpa)(H2O)2]2+ complexes with MeCOMet–Gly–Gly were regioselective and only the amide bond involving the carboxylic group of methionine was cleaved. However, in the reactions of these three Pd(II) complexes with MeCOGly–Met–Gly, two amide bonds, Met–Gly and MeCO–Gly, were cleaved. From UV–Vis spectrophotometry studies, it was found that the rate-determining step of these hydrolytic reactions is the monodentate coordination of the corresponding Pd(II) complex to the sulfur atom of the methionine side chain. The rate of the cleavage of these amide bonds is dependent on the nature of the bidentate coordinated diamine ligand L (en > pic > dpa). The hydrolytic reaction of cis-[Pd(L)(H2O)2]2+-type complexes with MeCOMet–Gly–Gly, containing the methionine side chain in the terminal position of the peptide, is regioselective while in the reaction of these Pd(II) complexes with MeCOGly–Met–Gly, none selective cleavage of the peptide occurs. This study contributes to a better understanding of the selective cleavage of methionine-containing peptides employing palladium(II) complexes as catalysts.  相似文献   

15.
Activation of the pro-drug isoniazid (INH) as an anti-tubercular drug in Mycobacterium tuberculosis involves its conversion to isonicotinyl-NAD, a reaction that requires the catalase-peroxidase KatG. This report shows that the reaction proceeds in the absence of KatG at a slow rate in a mixture of INH, NAD+, Mn2+, and O2, and that the inclusion of KatG increases the rate by >7 times. Superoxide, generated by either Mn2+- or KatG-catalyzed reduction of O2, is an essential intermediate in the reaction. Elimination of the peroxidatic process by mutation slows the rate of reaction by 60% revealing that the peroxidatic process enhances, but is not essential for isonicotinyl-NAD formation. The isonicotinyl-NAD•+ radical is identified as a reaction intermediate, and its reduction by superoxide is proposed. Binding sites for INH and its co-substrate, NAD+, are identified for the first time in crystal complexes of Burkholderia pseudomallei catalase-peroxidase with INH and NAD+ grown by co-crystallization. The best defined INH binding sites were identified, one in each subunit, on the opposite side of the protein from the entrance to the heme cavity in a funnel-shaped channel. The NAD+ binding site is ∼20 Å from the entrance to the heme cavity and involves interactions primarily with the AMP portion of the molecule in agreement with the NMR saturation transfer difference results.  相似文献   

16.
An electron-rich iron(III) porphyrin complex (meso-tetramesitylporphinato)iron(III) chloride [Fe(TMP)Cl], was found to catalyze the epoxidation of olefins by aqueous 30% H2O2 when the reaction was carried out in the presence of 5-chloro-1-methylimidazole (5-Cl-1-MeIm) in aprotic solvent. Epoxides were the predominant products with trace amounts of allylic oxidation products, indicating that Fenton-type oxidation reactions were not involved in the olefin epoxidation reactions. cis-Stilbene was stereospecifically oxidized to cis-stilbene oxide without giving isomerized trans-stilbene oxide product, demonstrating that neither hydroperoxy radical (HOO·) nor oxoiron(IV) porphyrin [(TMP)FeIV=O] was responsible for the olefin epoxidations. We also found that the reactivities of other iron(III) porphyrin complexes such as (meso-tetrakis(2,6-dichlorophenyl)porphinato)iron(III) chloride [Fe(TDCPP)Cl], (meso-tetrakis(2,6-difluorophenyl)porphinato)iron(III) chloride [Fe(TDFPP)Cl], and (meso-tetrakis(pentafluorophenyl)porphinato)iron(III) chloride [Fe(TPFPP)Cl] were significantly affected by the presence of the imidazole in the epoxidation of olefins by H2O2. These iron porphyrin complexes did not yield cyclohexene oxide in the epoxidation of cyclohexene by H2O2 in the absence of 5-Cl-1-MeIm in aprotic solvent; however, addition of 5-Cl-1-MeIm to the reaction solutions gave high yields of cyclohexene oxide with the formation of trace amounts of allylic oxidation products. We proposed, on the basis of the results of mechanistic studies, that the role of the imidazole is to decelerate the O–O bond cleavage of an iron(III) hydroperoxide porphyrin (or H2O2–iron(III) porphyrin adduct) and that the intermediate transfers its oxygen to olefins prior to the O–O bond cleavage.  相似文献   

17.
-Amino acid deaminase ( -AAD) from Proteus myxofaciens was cloned and over-expressed in Escherichia coli K12. This enzyme has a broad substrate specificity, working on both natural and unnatural -amino acids. Of the 20 naturally occurring -amino acids, -AAD prefers amino acid substrates that have aliphatic, aromatic or sulfur-containing side chains; those with charged side chains (–CO2 or –NH3+) are poor or non-substrates. Enzyme activity was monitored using a microtiter-plate-based assay, which measures the formation of phenylpyruvic acid from -phenylalanine. The reaction has an absolute requirement for O2, releases NH3 and does not produce H2O2. Substrate comparisons were carried out by using an O2 electrode to measure the O2 utilization rates. Studies on partially purified enzyme show a pH optimum of 7.5 with a subunit molecular weight of approximately 51 kDa. Additional purification and characterization strategies will be presented. The use of whole cells containing -AAD will be discussed to prepare chiral pharmaceutical intermediates.  相似文献   

18.
A novel effect of the inhibition of the decomposition of amino acids to carbonates on addition of imidazole (HIm) to a reacting system containing equimolar amounts of copper and zinc metal powders, an amino acid [glycine (Hgly), aspartic acid (H2Asp) or glycylglycine (H2gg)] (1:1:2) and excess hydrogen peroxide (H2O2) resulting in formation of a mixed metal mixed ligand peroxo complex compound was observed, because in the absence of imidazole the corresponding reaction system yields only a mixed metal peroxo carbonate. For the resulting complex compounds, the homogeneity, i.e. [Cu(Zn)(O2 2–)(Gly)2(HIm)(H2O)], [Cu(Zn)(O2 2–)(Asp)(HIm)(H2O)2] or [Cu(Zn)2(O2 2–)2(gg)(HIm)(H2O)4], molecular formula, presence of peroxo group and coordination environment were established by combined physicochemical evidence from elemental and thermogravimetric analysis in air and argon atmospheres, electron spin resonance and electronic and IR spectral data. It is noteworthy to mention that the corresponding carboxylic acids of the above-mentioned amino acids, i.e. acetic and succinic acids, either do not decompose to carbonates in the absence of imidazole or form novel homogeneous peroxo mixed metal mixed ligand complex compounds as described above in the presence of imidazole. This suggests an important and significant mutual influence (in vitro) of biologically active chromophores like peroxo ions, imidazole and amino groups in the above-mentioned chemical reactions containing bioactive metals such as copper and zinc.  相似文献   

19.
Cell-free extracts of Lactobacillus plantarum contain non-proteinaceous compounds which mimic superoxide dismutase activity. Using the test system in which O 2 is generated by xanthine oxidase, superoxide dismutase activity is found in cell-free extracts, where proteins are removed by precipitation. This activity is strongly decreased after dialysis of cell-free extracts. Superoxide dismutase activity was also investigated by means of pulse radiolysis. Cell-free extracts of Escherichia coli were also investigated as a comparison, which were known to contain superoxide dismutase. With cell-free extracts of both L. plantarum and E. coli the decay of O 2 was markedly increased. However, the type of reaction of the O 2 decay was of first order in the presence of E. coli extracts due to superoxide dismutase(s), and of second order in the presence of L. plantarum extracts, indicating that O 2 elimination is not an enzymic reaction. Mn2+ phosphate(s) might be responsible for the observed elimination of O 2 . The production of O 2 is not detectable during NADH-, lactate- or pyruvate oxidase reactions in L. plantarum extracts.  相似文献   

20.
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