首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
Stopped-flow kinetic studies of the formation of ferrioxamine B were performed. Formation of the complex follows the rate law
where Ka is the acid dissociation constant of the iron(III) aquo species in 0.1 M formate buffer. At 25°C k1 = 3.94 × 102M?1 sec?1, k2Ka = 1.18 × 10?1 sec?1, k3 = 3.6 × 10?1 sec?1. Activation parameters for k1 are ΔH = 11.7 kcal mole?1 and ΔS = ?8 cal K?1 mole?1. An associative mechanism is proposed. Attachment of the first chelate ring is the slow step and favorably positions the second chelate ring for attachment. Coordination of two chelate rings favorably positions the third chelate ring for attachment. These results are compared to kinetics of formation of model complexes and to a previous study of the formation of ferrioxamine B in which attachment of the third chelate ring was proposed as the slow step  相似文献   

2.
β-Xylosidase was purified 25 fold from a culture filtrate by ammonium sulfate fractionation, DEAE-Sephadex chromatography, column electrophoresis, gel filtration on Biogel P-100, and isoelectric focusing. The purified β-xylosidase was found to be homogeneous on SDS (sodium dodecyl sulfate) polyacrylamide gel electrophoresis and on disc electrophoresis. A molecular weight of 101,000 was estimated by chromatography on Sephadex G-200, and 102,000 was obtained by SDS polyacrylamide gel electrophoresis. The purified p-xylosidase had an isoelectric point at pH 4.45, and contained 4.5% carbohydrate residue. The optimum activity for the enzyme was found to be at pH 4.5 and 55°C. The enzyme activity was inhibited by Hg2 +, and N-bromosuccinimide at a concentration of 1 x 10?3 m. The purified enzyme hydrolyzed phenyl β-d-xyloside (ko13.0 sec”1), p-nitrophenyl β-d-xyloside (ko=2l.3 sec?1), o-nitrophenyl β-d-xyloside (ko = 22.2 sec?1), o-chlorophenyl β-d-xyloside (ko = 20.0 sec?1), p-methylphenyl β-d-xyloside (ko~9.0 sec?1), o-methylphenyl β-d-xyloside (ko= 10.7 sec?1), p-methoxyphenyl β-d-xyloside (ko=10.3 sec?1), o-methoxyphenyl β-d-xyloside (&;o=10.9 sec?1), xylobiose (ko = 36A sec?1), xylotriose (ko = 34.5 sec?1), xylotetraose (ko~HA sec?1), and xylopentaose (ko= 13.0 sec?1). On enzymic hydrolysis of phenyl β-d-xyloside, the reaction product was found to be β-d-xylose with retention of configuration. The purified p-xylosidase was practically free of α-xylosidase and β-glucosidase activities.  相似文献   

3.
β-Xylosidase was purified 662 fold from a culture filtrate by ammonium sulfate fractionation, gel filtration on Biogel P-100, DEAE-Sephadex chromatography, and gel filtration on Sephadex G-200. With isoelectric focusing, the purified β-xylosidase found to be homogeneous on SDS (sodium dodecyl sulfate) polyacrylamide gel electrophoresis. The molecular weight was estimated by gel filtration to be 240,000, and 116,000 by SDS polyacrylamide gel electrophoresis. The purified β-xylosidase had an isoelectric point at pH 3.25, and contained 4% carbohydrate residue. The optimum pH was found to be in the range of 4.5 ~ 5, and the optimum temperature was 55°C. The enzyme activity was inhibited by Hg2 +, SDS, and N-bromosuccinimide at a concentration of 1 × 10?3 m, and also p-chloromercuribenzoate at a concentration of 1 × 10?4m. The purified enzyme hydrolyzed phenyl β-d-xyloside (ko = 302.6 sec?1),β-nitrophenyl β-d-xyloside (ko = 438.9 sec?1), o-nitrophenyl β-d-xyloside (ko = 431.0 sec?1), p-chlorophenyl β-d-xyloside (ko = 207.9 sec?1), o-chlorophenyl β-d-xyloside (ko = 211.8 sec?1), β-methylphenyl β-d-xyloside ko = 96.5 sec?1), o-methylphenyl β-d-xyloside (ko = 83.1 sec?1), p-methoxyphenyl β-d-xyloside (ko = 99.3 sec?1), o-methoxyphenyl β-d-xyloside (ko= 100.0 sec?1), xylobiose (ko = 992A sec?1), xylotriose (ko = 1321.9 sec?1), xylotetraose (ko = 7S9.1 sec?1) and xylopentaose (ko = 508.0 sec?1). On enzymic hydrolysis of phenyl β-d-xyloside, the reaction product was found to be β-d-xylose with retention of the configuration. The purified β-xylosidase was practically free of a-xylosidase and β-glucosidase activities.  相似文献   

4.
The nonenzymatic reduction of nitrosobenzene by NADPH and NADH in aqueous buffer solution at 25°C is described. Both reactants quantitatively convert nitrosobenzene to phenylhydroxylamine. Rate constants for reduction (kr) were determined spectrophotometrically and found to be identical at pH 5.7 and 7.4 and independent of buffer concentration. The values of kNADH (124–149 M?1 sec?1) and kNADPH (131–170 M?1 sec?1) are essentially identical. The reaction is not subject to general catalysis or specific salt effects. The oxidation of phenylhydroxylamine by NAD(P) to nitrosobenzene is only stimulated by a factor of 1.2 over oxidation in its absence (when the ratio of NADP: phenylhydroxylamine was 8:1).  相似文献   

5.
Reaction rates were measured for the low pH-induced Cu ligand modification of azurins from Pseudomonas aeruginosa and Alcaligenes faecalis. Loss of the intense absorption band at 625 nm obeyed a rate law:
where Az is the concentration of azurin in its native oxidized form possessing the 625 nm band. For Pseudomonas aeruginosa at 25°C, n = 1 and k = 4.0 × 10-2 sec-1 M-1 in citrate buffer but 1.2 × 10-2 sec-1 M-1 in phosphate buffer. For Alcaligenes faecalis, n = 3 and k = 1.5 × 104 sec-1 M-3 in citrate and 2.6 × 103 sec-1 M-3 in phosphate. In equilibrium experiments on Alcaligenes faecalis azurin in citrate buffer, the pH-dependent change to the low pH form exhibited an apparent transition pK of 3.1. The form of the rate law implies a mechanistic scheme that contains fast equilibrium protonation steps prior to a rate limiting ligand rearrangement.  相似文献   

6.
The reactions of copper(II)-ahphatic 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 ammo acid oxidations, as changes in the rate law could be induced by modest variations in reductant concentration (at fixed [Cu(II)]o), pH, and the structure of the redox partners. With excess cysteine present at 25°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 × 103M?1sec?1), Cu(Me5dien)2+ (2.7 × 103M?1 sec?1), Cu(en)22+ (2.1 × 103M?1 sec?1), and Cu(dien)22+ (4.7 × 103 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 mtermediacy of di-μ-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°C) 7.5 × 107 M?2 sec?1, I = 0.2 M). 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 purposed, coupled with attack of a noncoordinated cysteine molecule on a vacated coordination position to stabilize the (Me6(tren)Cu(I) product.  相似文献   

7.
The binding of cis(c)- and trans(t)-Pt(NH3)2Cl2 to DNA at platinum/DNA-nucleotide ratios (Ri) of 0.1 or less has been studied by means of radioactive 195mPt-labeled compounds. Kinetic data are consistent with the following scheme:
At 25°C and pH 5–6 in 5 mM NaClO4, the values for the rate constants in the above scheme for the c-isomer are k2 = 2.2 × 10?5 sec?1, k7 = 0.32 (sec M)?1, and k8 = 143 (sec M)?1; for the t-isomer the values are k2 < 0.5 × 10?5 sec?1 and k7 = 0.95 (sec M)?1. Platinum-DNA adducts do not undergo detectable exchange after 3 days at 37°C, indicating the absence of a dynamic equillibrium. For both isomers the rate of binding is the same for single- and double-stranded DNA. The conclusions derived from Ag+ and H+ titration studies are consistent with binding at guanine N(7) for Ri < 0.1. The reaction rate is competitively inhibited by various salts and buffers and is suppressed by raising the pH (50% inhibition of initial rates at pH 7.3). At 37°C and pH 7 in 0.15 M NaCl, 6–8% of both the c- and t-isomers bind to DNA in 24 h, suggesting that both compounds should bind to DNA under biological conditions.  相似文献   

8.
The kinetics of the reduction by aniline and a series of substituted anilines of a peroxidatically active intermediate, formed by oxidation of deuteroferriheme with hydrogen peroxide, have been studied by stopped-flow spectrophotometry. The reaction with aniline was first order with respect to [intermediate] and showed first-order saturation kinetics with respect to [aniline]. The second-order rate constant was 2.0 ± 0.2 × 105 M?1 sec?1 at 25°C (independent of pH in the range 6.60–9.68) compared with the value of 2.4 × 105 M?1 sec?1 for the reaction of aniline with horseradish peroxidase Compound I. The effect of aniline substituents upon reactivity towards the heme intermediate closely paralled those reported for reaction with the enzymic intermediate. Anilines bearing electron-donating substituents reacted more rapidly and those bearing electron-withdrawing substituents more slowly than the unsubstituted amine. The rate constants for the heme intermediate reactions (kdfh)found to be related to those for the enzymic reactions (khrp) by the equation:log kDFH= 0.65log kHRP+ 1.96 with a correlation coefficient of 0. 98.  相似文献   

9.
Interaction between duodenase (a granase family member) from bovine duodenal mucosa and recombinant antichymotrypsin (rACT) and its P1 variants has been studied. Association rate constants (k a) were 11, 6.8, and 17 mM?1·sec?1 for rACT, ACT L358M, and ACT L358R, respectively. Natural antitrypsin (AT) compared to ACT was a 20 times more effective duodenase inhibitor (in terms of k a). Duodenase interacted with P1 variants of ACT via a suicide mechanism with stoichiometry of the process SI = 1.2. The nature of the P1 residue of the inhibitor did not influence the interaction if other residues did not meet conformational requirements of the duodenase substrate-binding pocket. Also, interaction of duodenase with ACT variants containing residues from AT reaction center loop (rACT P2-P3′, rACT P3-P4′, rACT P4-P3′, and rACT P6-P4′) was studied. The inhibition type ([E]0 = 1·10?7 M, 25°C) was revealed to be reversible-like, and efficacy of inhibition decreased with increase in the substituted part of the reactive center loop. Constants of inhibition (K i) were measured. Efficacy of interaction between the enzyme (duodenase) and inhibitor depends on topochemical correspondence between a substrate-binding pocket of the enzyme and substrate structure.  相似文献   

10.
Horse heart carboxymethylated cytc (CM-cytc) displays myoglobin-like properties. Here, the effect of cardiolipin (CL) liposomes on the nitrite reductase activity of ferrous CM-cytc [CM-cytc-Fe(II)], in the presence of sodium dithionite, is reported between pH 5.5 and 7.6, at 20.0 °C. Cytc-Fe(II) displays a very low value of the apparent second-order rate constant for the NO2 ?-mediated conversion of cytc-Fe(II) to cytc-Fe(II)-NO [k on = (7.3 ± 0.7) × 10?2 M?1 s?1; at pH 7.4], whereas the value of k on for NO2 ? reduction by CM-cytc-Fe(II) is 1.1 ± 0.2 M?1 s?1 (at pH 7.4). CL facilitates the NO2 ?-mediated nitrosylation of CM-cytc-Fe(II) in a dose-dependent manner, the value of k on for the NO2 ?-mediated conversion of CL–CM-cytc-Fe(II) to CL–CM-cytc-Fe(II)-NO (5.6 ± 0.6 M?1 s?1; at pH 7.4) being slightly higher than that for the NO2 ?-mediated conversion of CL–cytc-Fe(II) to CL–cytc-Fe(II)-NO (2.6 ± 0.3 M?1 s?1; at pH 7.4). The apparent affinity of CL for CM-cytc-Fe(II) is essentially pH independent, the average value of B being (1.3 ± 0.3) × 10?6 M. In the absence and presence of CL liposomes, the nitrite reductase activity of CM-cytc-Fe(II) increases linearly on lowering pH and the values of the slope of the linear fittings of Log k on versus pH are ?1.05 ± 0.07 and ?1.03 ± 0.03, respectively, reflecting the involvement of one proton for the formation of the transient ferric form, NO, and OH?. These results indicate that Met80 carboxymethylation and CL binding cooperate in the stabilization of the highly reactive heme-Fe atom of CL–CM-cytc.  相似文献   

11.
The kinetics of the binding of cyanide to ferric chloroperoxidase have been studied at 25°C and ionic strength 0.11 M using a stopped-flow apparatus. The dissociation constant (KCN) of the peroxidase-cyanide complex and both forward (k+) and reverse (k?) rate constants are independent of the H+ concentration over the pH range 2.7 to 7.1. The values obtained are kcn = (9.5 ± 1.0) × 10-5 M, k+. = (5.2 ± 0.5) × 104 M?1 sec?1 and k- = (5.0± 1.4) sec-1. In the presence of 0 06 M potassium nitrate the affinity of cyanide for chloroperoxidase decreases due to the inhibition of the forward reaction. The dissociation rate is not affected. The nitrate anion exerts its influence by binding to a protonated form of the enzyme, whereas the cyanide binds to the unprotonated form. Binding of nitrate results in an apparent shift towards higher pKa values of the ionization of a crucial heme-linked acid group. Hence the influence of this group can be detected in the accessible pH range. Extrapolation to zero nitrate concentration yields a value of 3.1±0.3 for the pKa of the heme-linked acid group.  相似文献   

12.
The acid-catalyzed hydrolysis of heparin from Cu(II) complex was studied as a function of time and temperature. Four independent calculations showed that the hydrolysis, during the 5-hr period examined, obeys the first-order kinetic law. Specific rate constants, calculated at 50°C, 57°C, 65°C, 71°C, and 80°C, were 3.3 × 10?5 sec?1, 6.5 × 10?5 sec?1, 10.4 × 10?5 sec?1, 15.1 × 10?5 sec?1, and 26.6 × 10?5 sec?1, respectively. Arrhenius plots of the data yielded 14.7 kcal as the energy of activation. An independent run of the self-hydrolysis of heparin at 57°C also obeyed first-order kinetics and its specific rate constant of 6.4 × 10?5 sec?1 is in excellent agreement with that of the hydrolysis of Cu(II)-heparin at 57°C. The anticoagulant activity of heparin and of the Cu(II)-heparin are not appreciably different. Further, the inactivation of heparin closely parallels Cu(II) release from the Cu(II) complex which in turn parallels desulfation.  相似文献   

13.
Genes of β-mannosidase 97 kDa, GH family 2 (bMann9), β-mannanase 48 kDa, GH family 5 (bMan2), and α-galactosidase 60 kDa, GH family 27 (aGal1) encoding galactomannan-degrading glycoside hydrolases of Myceliophthora thermophila C1 were successfully cloned, and the recombinant enzymes were purified to homogeneity and characterized. bMann9 displays only exo-mannosidase activity, the K m and k cat values are 0.4 mM and 15 sec?1 for p-nitrophenyl-β-D-mannopyranoside, and the optimal pH and temperature are 5.3 and 40°C, respectively. bMann2 is active towards galac-tomannans (GM) of various structures. The K m and k cat values are 1.3 mg/ml and 67 sec?1 for GM carob, and the optimal pH and temperature are 5.2 and 69°C, respectively. aGal1 is active towards p-nitrophenyl-α-D-galactopyranoside (PNPG) as well as GM of various structures. The K m and k cat values are 0.08 mM and 35 sec?1 for PNPG, and the optimal pH and temperature are 5.0 and 60°C, respectively.  相似文献   

14.
Transport of electrons in spinach photosystem II (PSII) whose oxygen-evolving complex (OEC) contains heterogeneous metal clusters 2Mn2Fe and 3Mn1Fe was studied by measuring the fluorescence induction kinetics (FIK). PSII(2Mn,2Fe) and PSII(3Mn,1Fe) preparations were produced using Cadepleted PSII membranes (PSII(–Ca)). It was found that FIK in PSII(2Mn,2Fe) membranes is similar in form to FIK in PSII(–Ca) samples, but the fluorescence yield is lower in PSII(2Mn,2Fe). The results demonstrate that, just as in PSII(–Ca) preparations, there is electron transfer from the metal cluster in the OEC to the primary plastoquinone electron acceptor QA. They also show that partial substitution of Mn cations with Fe has no effect on the electron transport on the acceptor side of PSII. Thus, these data demonstrate the possibility of water oxidation either by the heterogeneous metal cluster or just by the manganese dimer. We established that FIK in PSII(3Mn,1Fe) preparations are similar in form to FIK in PSII(2Mn,2Fe) membranes but PSII(3Mn,1Fe) is characterized by a slightly higher maximal fluorescence yield, Fmax. The electron transfer rate in PSII(3Mn,1Fe) preparations significantly (by a factor of two) increases in the presence of Ca2+, whereas Ca2+ has hardly any effect on the electron transport in PSII(2Mn,2Fe) membranes. In Mndepleted PSII membranes, FIK reaches its maximum (the so-called peak K), after which the fluorescence yield starts to decrease as the result of two factors: the oxidation of reduced primary plastoquinone Q A ? and the absence of electron influx from the donor side of PSII. The replacement of Mn cations by Fe in PSII(?Mn) preparations leads to fluorescence saturation and disappearance of the K peak. This is possibly due to the deceleration of the charge recombination process that takes place between reduced primary electron acceptor Q A ? and oxidized tyrosine Y Z +. which is an electron carrier between the OEC and the primary electron donor P680.  相似文献   

15.
The reaction of parsley 2Fe-2S ferredoxin in the normal oxidized state with eaq? generated by pulse radiolysis techniques has been studied at ~25°C, pH 7–8, I = 0.10 M (NaClO4). Rate constants ke (eaq? decay) and kp (protein absorbance change) are the same, second-order rate constant 9.7 × 109 M?1 sec?1. The reaction exhibits close to 100% efficiency. With 8Fe-8S ferredoxin from Clostridium pasteurianum under identical conditions it now appears that kp (although sometimes significantly smaller) is equal to ke. Varying efficiencies are also observed with this protein depending on the batch used. The reasons for such variable behavior are not fully understood. With oxidized and reduced forms of Chromatium v. high-potential iron-sulfur protein (HIPIP), ke and kp are essentially the same, but the highest efficiency observed is only ~50%. The prevailing pattern is therefore that rate constants ke and kp are generally in step for proteins having a single (or identical) active site(s). When the active site is buried as with HIPIP the efficiency of the reaction appears to decrease.  相似文献   

16.
Bacterial manganese(II) oxidation has a profound impact on the biogeochemical cycling of Mn and the availability of the trace metals adsorbed to the surfaces of solid Mn(III, IV) oxides. The Mn(II) oxidase enzyme was tentatively identified in Pseudomonas putida GB-1 via transposon mutagenesis: the mutant strain GB-1-007, which fails to oxidize Mn(II), harbors a transposon insertion in the gene cumA. cumA encodes a putative multicopper oxidase (MCO), a class of enzymes implicated in Mn(II) oxidation in other bacterial species. However, we show here that an in-frame deletion of cumA did not affect Mn(II) oxidation. Through complementation analysis of the oxidation defect in GB-1-007 with a cosmid library and subsequent sequencing of candidate genes we show the causative mutation to be a frameshift within the mnxS1 gene that encodes a putative sensor histidine kinase. The frameshift mutation results in a truncated protein lacking the kinase domain. Multicopy expression of mnxS1 restored Mn(II) oxidation to GB-1-007 and in-frame deletion of mnxS1 resulted in a loss of oxidation in the wild-type strain. These results clearly demonstrated that the oxidation defect of GB-1-007 is due to disruption of mnxS1, not cumA::Tn5, and that CumA is not the Mn(II) oxidase. mnxS1 is located upstream of a second sensor histidine kinase gene, mnxS2, and a response regulator gene, mnxR. In-frame deletions of each of these genes also led to the loss of Mn(II) oxidation. Therefore, we conclude that the MnxS1/MnxS2/MnxR two-component regulatory pathway is essential for Mn(II) oxidation in P. putida GB-1.In living cells, manganese (Mn) is an essential trace element, required for enzymes such as superoxide dismutase and in photosystem II (7). In the environment, Mn cycles between a soluble reduced form [Mn(II)] and an insoluble oxidized form [Mn(III, IV)] that can adsorb other trace metals from the environment and serve as potent oxidizing agents. Thus, redox cycling of Mn has a profound effect on the bioavailability and geochemical cycling of many essential or toxic elements (40). Microorganisms, particularly bacteria, are capable of catalyzing the oxidation of Mn(II), thereby increasing the rate of formation of Mn(III, IV) by several orders of magnitude (39). Since Mn(III, IV) oxides are able to bind trace metals, the bacteria that catalyze their formation are good candidates for bioremediation of heavy metal contaminated sites (26, 39).Although bacterial Mn(II) oxidation is widespread, little is known about the physiological function of oxidation (40). The oxidation of Mn(II) to Mn(III) or Mn(IV) is thermodynamically favorable; thus, bacteria may derive energy from this reaction, although this has never been unequivocally proven (40). In addition, Mn(II) oxidation could protect cells from reactive oxygen species (4) or UV irradiation (11). Since oxidation occurs on the cell surface, the bacteria become coated with the solid Mn(IV) oxides, which may also provide protection from toxic heavy metals, predation, or phage infection (40). As a strong oxidant, Mn(IV) oxides could allow the bacteria to degrade refractory organic matter to low-molecular-weight compounds that could then be used to support bacterial growth (38). Conversely, Mn(II) oxidation may be a side reaction or the result of nonspecific interactions with cellular products (15). Identifying signals or conditions that regulate oxidation could provide some insight into the role of Mn(II) oxidation in the cell. Aside from a requirement for oxygen (28) and iron (27, 30), as well as the observation that oxidation occurs in stationary phase (23), very little is known about this regulation.The enzymes responsible for Mn(II) oxidation have been tentatively identified from some species of bacteria and in several cases the enzyme is a putative multicopper oxidase (MCO). MCOs are a family of enzymes that use four Cu ion cofactors to catalyze oxidation of diverse substrates such as metals and organic compounds (33). This family of enzymes is found in plants and fungi (laccase) and humans (ceruloplasmin), as well as in bacteria (35). Some fungi have been shown to use a laccase enzyme to oxidize Mn(II) (20). In both Leptothrix discophora SS-1 and Pedomicrobium sp. strain ACM 3067, the Mn(II)-oxidizing MCO was identified genetically (mofA [10] and moxA [31], respectively). A third MCO—MnxG—was identified both biochemically and genetically as the Mn(II) oxidase in Bacillus sp. strain SG-1 and related strains (14, 43). Recent work with the Mn(II)-oxidizing alphaproteobacterium Aurantimonas manganoxydans SI85-9A1 and Erythrobacter sp. strain SD21 has identified a second class of enzyme involved in Mn(II) oxidation: the heme-binding peroxidase named MopA (3). This class of enzyme had previously been shown to be used by fungi to oxidize Mn(II) (29), in some cases in concert with an MCO (34).Pseudomonas putida GB-1 is a Mn(II)-oxidizing bacterium (9) whose genetic tractability and ease of growth under standard laboratory conditions make it an ideal model system for studying the physiology and mechanism of Mn(II) oxidation. Consequently, several random transposon mutagenesis screens have been undertaken with this organism to identify genes required for Mn(II) oxidation. These screens have identified several categories of genes as important for oxidation or the export of the oxidase to the cell surface: the ccm operon of c-type cytochrome synthesis genes (8, 13), genes encoding components of the trichloroacetic acid (TCA) cycle and the tryptophan biosynthesis pathway (8) and genes encoding a general secretory pathway (12). The Mn(II) oxidation-defective mutant GB-1-007 has a transposon insertion in the gene cumA that encodes a putative MCO (6). Therefore, P. putida GB-1 has been thought to use a similar mechanism as L. discophora SS-1, Pedomicrobium sp. strain ACM 3067, and Bacillus sp. to oxidize Mn(II).Because the available data suggested that CumA was an MCO essential for Mn(II) oxidation, we wanted to study its function in greater detail. We were hampered in this, however, by the fact that the transposon insertion in cumA resulted in a growth defect due to its polar effect on expression of the downstream cumB gene (6). In order to assess the role of CumA in Mn(II) oxidation without the complications arising from polarity, we generated an in-frame deletion of cumA and tested the ability of the resulting ΔcumA strain to form Mn(IV) oxides. Our results showed that cumA is dispensable for Mn(II) oxidation and have instead revealed a complex two-component regulatory pathway essential for Mn(II) oxidation in P. putida GB-1.  相似文献   

17.
Peroxynitrite-mediated oxidation of ferrous nitrosylated myoglobin (Mb(II)-NO) involves the transient ferric nitrosylated species (Mb(III)-NO), followed by NO dissociation and formation of ferric myoglobin (Mb(III)). In contrast, peroxynitrite-mediated oxidation of ferrous oxygenated myoglobin (Mb(II)-O2) involves the transient ferrous deoxygenated and ferryl derivatives (Mb(II) and Mb(IV)O, respectively), followed by Mb(III) formation. Here, kinetics of peroxynitrite-mediated oxidation of ferrous carbonylated horse heart myoglobin (Mb(II)-CO) is reported. Values of the first-order rate constant for peroxynitrite-mediated oxidation of Mb(II)-CO (i.e., for Mb(III) formation) and of the first-order rate constant for CO dissociation from Mb(II)-CO (i.e., for Mb(II) formation) are h = (1.2 ± 0.2) × 10−2 s−1 and koff(CO) = (1.4 ± 0.2) × 10−2 s−1, respectively, at pH 7.2 and 20.0 °C. The coincidence of values of h and koff(CO) indicates that CO dissociation represents the rate limiting step of peroxynitrite-mediated oxidation of Mb(II)-CO.  相似文献   

18.
19.
Nicotinamidase is involved in the maintenance of NAD+ homeostasis and in the NAD+ salvage pathway of most prokaryotes, and it is considered as a possible drug target. The gene (ASAC_0847) encoding a hypothetical nicotinamidase has been found in the genome of the thermophilic archaeon Acidilobus saccharovorans. The product of this gene, NA_As0847, has been expressed in Escherichia coli, isolated, and characterized as a Fe2+-containing nicotinamidase (k cat/K m = 427 mM?1·sec?1)/pyrazinamidase (k cat/K m = 331 mM?1·sec?1). NA_As0847 is a homodimer with molecular mass 46.4 kDa. The enzyme has high thermostability (T1/2 (60°C) = 180 min, T1/2 (80°C) = 35 min) and thermophilicity (Topt = 90°C, Ea = 30.2 ± 1.0 kJ/mol) and broad pH interval of activity, with the optimum at pH 7.5. Special features of NA_As084 are the presence of Fe2+ instead of Zn2+ in the active site of the enzyme and inhibition of the enzyme activity by Zn2+ at micromolar concentrations. Analysis of the amino acid sequence revealed a new motif of the metal-binding site (DXHXXXDXXEXXXWXXH) for homological archaeal nicotinamidases.  相似文献   

20.
The reaction of mercuric acetate with polypeptides in an appropriate buffer system has been found to result in the selective binding of two atoms of mercury to each tyrosine and histidine residue. These heavy atom labels are stable to the high chloride concentrations used to displace the excess mercury (II) from other binding sites on the polypeptide. The kinetics and stoichiometry of these reactions have been studied by the binding of radioactive (203Hg) mercuric acetate to synthetic polymers of these amino acids and by ultraviolet-visible spectroscopy. For a polymer containing tyrosine the mercuration kinetics closely match those for the following mechanism:
At 60°C, and in a buffer containing 0.05M TRIS-acetate, k1 was determined to be 9.47 ± 0.27 M?1 min?1. The best match to the data was for k1/k1 = 5.5 It was discovered that there is an inverse relationship between k1 and the TRIS buffer concentration. The activation energy of k1 was determined to be 18.9 ± 0.1 kcal/mole.Chemical analyses of the products obtained from the reaction of mercuric acetate with tyrosine amide, L(-)-histidine and the methyl ester of L(-)-histidine have established that mercuration results in the formation of a Hg-C bond at the C3 and C5 sites on the phenolic ring in tyrosine and at the C4 site in the imidazole ring in histidine. The site of the second mercury retained by histidine in the presence of high chloride concentrations is uncertain but does involve the amine functions of the imidazole ring.The reaction conditions employed also cause the oxidation of methionine and cysteine to methionine sulfoxide and cysteic acid, respectively. Cystine, however, resists oxidation.  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