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.
An enzymatic assay was developed for the spectrophotometric determination of glycolate in urine and plasma. Glycolate was first converted to glyoxylate with glycolate oxidase, and the glyoxylate formed was condensed with phenylhydrazine. The glyoxylate phenylhydrazone formed was then oxidized with K(3)Fe(CN)(6) in the presence of excess phenylhydrazine, and A(515) of the resulting 1, 5-diphenylformazan was measured. Since glycolate oxidase also acts on glyoxylate and L-lactate, the incubation of samples with glycolate oxidase was carried out in 120-170 mM Tris-HCl (pH 8.3) to obtain glyoxylate as its adduct with Tris. The pyruvate formed from lactate was removed by subsequent brief incubation with alanine aminotransferase in the presence of L-glutamate, and alpha-ketoglutarate formed was converted back to L-glutamate by glutamate dehydrogenase and an NADPH generating system. Thus the specificity of the assay relies principally on the substrate specificity of glycolate oxidase, and high sensitivity is provided by the high absorbance of 1,5-diphenylformazan at 515-520 nm. Plasma was deproteinized with perchloric acid, and then neutralized with KOH. Plasma and urine samples were then incubated with approximately 5 mM phenylhydrazine, and then treated with stearate-deactivated activated charcoal to remove endogenous keto and aldehyde acids as their phenylhydrazones. The normal plasma glycolate and urinary glycolate/creatinine ratio for adults determined by this method are approximately 8 microM and approximately 0.036, respectively.  相似文献   

2.
Indoleamine 2,3-dioxygenase is an important mammalian target that catalyses the oxidative cleavage of l-tryptophan to N-formylkynurenine. In this work, the redox properties of recombinant human indoleamine 2,3-dioxygenase (rhIDO) and its H303A variant have been examined for the first time and the spectroscopic and substrate-binding properties of rhIDO and H303A in the presence and absence of substrate are reported. The Fe(3+)/Fe(2+) reduction potential of H303A was found to be -30 +/- 4 mV; in the presence of l-Trp, this value increases to +16 +/- 3 mV. A variety of spectroscopies indicate that ferric rhIDO at pH 6.6 exists as a mixture of six-coordinate, high-spin, water-bound heme and a low-spin species that contains a second nitrogenous ligand; parallel experiments on H303A are consistent either with His303 as the sixth ligand or with His303 linked to a conformational change that affects this transition. There is an increase in the low-spin component at alkaline pH for rhIDO, but this is not due to hydroxide-bound heme. Substrate binding induces a conformational rearrangement and formation of low-spin, hydroxide-bound heme; analysis of the H303A variant indicates that His303 is not required for this conversion and is not essential for substrate binding. The Fe(3+)/Fe(2+) reduction potential of H303A variant is approximately 70 mV lower than that of rhIDO, leading to a destabilization of the ferrous-oxy complex, which is an obligate intermediate in the catalytic process. In comparison with the properties of other heme enzymes, the data can be used to build a more detailed picture of substrate binding and catalysis in indoleamine 2,3-dioxygenase. The wider implications of these results are discussed in the context of our current understanding of the catalytic mechanism of the enzyme.  相似文献   

3.
An optical flow cell provided a means to conveniently measure the rate of successive Fe(2+) oxidation reactions catalyzed by horse spleen ferritin (HoSF) to determine if both ferroxidase and mineral core Fe(2+) oxidation reactions occur. The oxygen concentration and pH were held constant and multiple additions of Fe(2+)/HoSF ratios of 1, 10, 100, 150, 250 and 400 were conducted, creating core sizes ranging from 12 to 2800. During these oxidations, the absence of nonspecific Fe(OH)(3) formation and the presence (>95%) of Fe(OH)(3) deposited within the core of HoSF demonstrated the validity of monitoring iron deposition into HoSF by this procedure. Initial rates for oxidation of 5-50 Fe(2+)/HoSF established that the reaction is overall first order in Fe(2+) concentration. However, when full progress curves were analyzed at a variety of Fe(2+)/HoSF ratios, two first-order reactions (k(1) approximately 0.035 s(-1) and k(2) approximately 0.007 s(-1)) were found to contribute to the overall Fe(2+) oxidation reaction. The proportion of the fast reaction increased with increasing Fe(2+)/HoSF ratio until at approximately 400, it was the dominant reaction. For the Fe(2+)/HoSF ratios examined, the overall rate of iron deposition is independent of the size of the mineral core, a result suggesting that an increasing mineral core size does not enhance the rate of Fe(2+) oxidation. Comparison of successive additions of 1.0 Fe(2+)/HoSF showed that oxidation of the first 8-10 Fe(2+) produced a Fe(III) species with a lower molar absorptivity per Fe(III) than that of the bulk core. Measurement of the H(+)/Fe(2+) ratio confirmed this difference in behavior by giving an H(+)/Fe(2+) ratio of approximately 1.0 below and 2.0 for ratios >30 Fe(2+)/HoSF. The faster reaction was attributed to ferroxidase catalysis and the slow reaction to nonspecific ferroxidase activity of the HoSF protein shell.  相似文献   

4.
The kinetics of ferrous iron oxidation by Leptospirillum ferriphilum (L. ferriphilum) dominated culture was studied in the concentration range of 0.1-20 g Fe(2+)/L and the effect of ferric iron (0-60 g Fe(3+)/L) on Fe(2+) oxidation was investigated at pH below one. Denaturing gradient gel electrophoresis of PCR amplified 16S rRNA genes followed by partial sequencing confirmed that the bacterial community was dominated by L. ferriphilum. In batch assays, Fe(2+) oxidation started without lag phase and the oxidation was completed within 1 to 60 h depending on the initial Fe(2+) concentration. The specific Fe(2+) oxidation rates increased up to around 4 g/L and started to decrease at above 4 g/L. This implies substrate inhibition of Fe(2+) oxidation at higher concentrations. Haldane equation fitted the experimental data reasonably well (R(2) = 0.90). The maximum specific oxidation rate (q(m)) was 2.4 mg/mg VS . h, and the values of the half saturation (K(s)) and self inhibition constants (K(i)) were 413 and 8,650 mg/L, respectively. Fe(2+) oxidation was competitively inhibited by Fe(3+) and the competitive inhibition constant (K(ii)) was 830 mg/L. The time required to reach threshold Fe(2+) concentration was around 1 day and 2.3 days with initial Fe(3+) concentration of 5 and 60 g/L, respectively. The threshold Fe(2+) concentration, below which no further Fe(2+) oxidation occurred, linearly increased with increasing initial Fe(2+) and Fe(3+) concentrations. Fe(2+) oxidation proceeds by L. ferriphilum dominated culture at pH below 1 even in the presence of 60 g Fe(3+)/L. This indicates potential of using and biologically regenerating concentrated Fe(3+) sulfate solutions required, for example, in indirect tank leaching of ore concentrates.  相似文献   

5.
Iron deposition within the iron storage protein ferritin involves a complex series of events consisting of Fe(2+) binding, transport, and oxidation at ferroxidase sites and mineralization of a hydrous ferric oxide core, the storage form of iron. In the present study, we have examined the thermodynamic properties of Fe(2+) binding to recombinant human H-chain apoferritin (HuHF) by isothermal titration calorimetry (ITC) in order to determine the location of the primary ferrous ion binding sites on the protein and the principal pathways by which the Fe(2+) travels to the dinuclear ferroxidase center prior to its oxidation to Fe(3+). Calorimetric titrations show that the ferroxidase center is the principal locus for Fe(2+) binding with weaker binding sites elsewhere on the protein and that one site of the ferroxidase center, likely the His65 containing A-site, preferentially binds Fe(2+). That only one site of the ferroxidase center is occupied by Fe(2+) implies that Fe(2+) oxidation to form diFe(III) species might occur in a stepwise fashion. In dilute anaerobic protein solution (3-5 microM), only 12 Fe(2+)/protein bind at pH 6.51 increasing to 24 Fe(2+)/protein at pH 7.04 and 7.5. Mutation of ferroxidase center residues (E62K+H65G) eliminates the binding of Fe(2+) to the center, a result confirming the importance of one or both Glu62 and His65 residues in Fe(2+) binding. The total Fe(2+) binding capacity of the protein is reduced in the 3-fold hydrophilic channel variant S14 (D131I+E134F), indicating that the primary avenue by which Fe(2+) gains access to the interior of ferritin is through these eight channels. The binding stoichiometry of the channel variant is one-third that of the recombinant wild-type H-chain ferritin whereas the enthalpy and association constant for Fe(2+) binding are similar for the two with an average values (DeltaH degrees = 7.82 kJ/mol, binding constant K = 1.48 x 10(5) M(-)(1) at pH 7.04). Since channel mutations do not completely prevent Fe(2+) binding to the ferroxidase center, iron gains access to the center in approximately one-third of the channel variant molecules by other pathways.  相似文献   

6.
Kinetics and inhibition of Na(+)/K(+)-ATPase and Mg(2+)-ATPase activity from rat synaptic plasma membrane (SPM), by separate and simultaneous exposure to transition (Cu(2+), Zn(2+), Fe(2+) and Co(2+)) and heavy metals (Hg(2+) and Pb(2+)) ions were studied. All investigated metals produced a larger maximum inhibition of Na(+)/K(+)-ATPase than Mg(2+)-ATPase activity. The free concentrations of the key species (inhibitor, MgATP(2-), MeATP(2-)) in the medium assay were calculated and discussed. Simultaneous exposure to the combinations Cu(2+)/Fe(2+) or Hg(2+)/Pb(2+) caused additive inhibition, while Cu(2+)/Zn(2+) or Fe(2+)/Zn(2+) inhibited Na(+)/K(+)-ATPase activity synergistically (i.e., greater than the sum metal-induced inhibition assayed separately). Simultaneous exposure to Cu(2+)/Fe(2+) or Cu(2+)/Zn(2+) inhibited Mg(2+)-ATPase activity synergistically, while Hg(2+)/Pb(2+) or Fe(2+)/Zn(2+) induced antagonistic inhibition of this enzyme. Kinetic analysis showed that all investigated metals inhibited Na(+)/K(+)-ATPase activity by reducing the maximum velocities (V(max)) rather than the apparent affinity (Km) for substrate MgATP(2-), implying the noncompetitive nature of the inhibition. The incomplete inhibition of Mg(2+)-ATPase activity by Zn(2+), Fe(2+) and Co(2+) as well as kinetic analysis indicated two distinct Mg(2+)-ATPase subtypes activated in the presence of low and high MgATP(2-) concentration. EDTA, L-cysteine and gluthathione (GSH) prevented metal ion-induced inhibition of Na(+)/K(+)-ATPase with various potencies. Furthermore, these ligands also reversed Na(+)/K(+)-ATPase activity inhibited by transition metals in a concentration-dependent manner, but a recovery effect by any ligand on Hg(2+)-induced inhibition was not obtained.  相似文献   

7.
The alternative low-spin states of Fe(3+) and Fe(2+) cytochrome c induced by SDS or AOT/hexane reverse micelles exhibited the heme group in a less rhombic symmetry and were characterized by electron paramagnetic resonance, UV-visible, CD, magnetic CD, fluorescence, and Raman resonance. Consistent with the replacement of Met(80) by another strong field ligand at the sixth heme iron coordination position, Fe(3+) ALSScytc exhibited 1-nm Soret band blue shift and epsilon enhancement accompanied by disappearance of the 695-nm charge transfer band. The Raman resonance, CD, and magnetic CD spectra of Fe(3+) and Fe(2+) ALSScytc exhibited significant changes suggestive of alterations in the heme iron microenvironment and conformation and should not be assigned to unfold because the Trp(59) fluorescence remained quenched by the neighboring heme group. ALSScytc was obtained with His(33) and His(26) carboxyethoxylated horse cytochrome c and with tuna cytochrome c (His(33) replaced by Asn) pointing out Lys(79) as the probable heme iron ligand. Fe(3+) ALSScytc retained the capacity to cleave tert-butylhydroperoxide and to be reduced by dithiothreitol and diphenylacetaldehyde but not by ascorbate. Compatible with a more open heme crevice, ALSScytc exhibited a redox potential approximately 200 mV lower than the wild-type protein (+220 mV) and was more susceptible to the attack of free radicals.  相似文献   

8.
Synthesis of oxalic Acid by enzymes from lettuce leaves   总被引:3,自引:0,他引:3       下载免费PDF全文
A rapid purification of lactate dehydrogenase and glycolate oxidase from lettuce (Lactuca sativa) leaves is described. The kinetics of both enzymes are reported in relation to their possible roles in the production of oxalate. Lettuce lactate dehydrogenase behaves like mammalian dehydrogenase, catalyzing the dismutation of glyoxylate to glycolate and oxalate. A model is proposed in which glycolate oxidase in the peroxisomes and lactate dehydrogenase in the cytosol are involved in the production of oxalate. The effect of pH on the balance between oxalate and glycolate produced from glyoxylate suggests that in leaves lactate dehydrogenase may function as part of an oxalate-based biochemical, pH-stat.  相似文献   

9.
Kálmán L  LoBrutto R  Allen JP  Williams JC 《Biochemistry》2003,42(37):11016-11022
The transfer of an electron from exogenous manganese (II) ions to the bacteriochlorophyll dimer, P, of bacterial reaction centers was characterized for a series of mutants that have P/P(+) midpoint potentials ranging from 585 to 765 mV compared to 505 mV for wild type. Light-induced changes in optical and EPR spectra of the mutants were measured to monitor the disappearance of the oxidized dimer upon electron donation by manganese in the presence of bicarbonate. The extent of electron transfer was strongly dependent upon the P/P(+) midpoint potential. The midpoint potential of the Mn(2+)/Mn(3+) couple was calculated to decrease linearly from 751 to 623 mV as the pH was raised from 8 to 10, indicating the involvement of a proton. The electron donation had a second order rate constant of approximately 9 x 10(4) M(-1) s(-1), determined from the linear increase in rate for Mn(2+) concentrations up to 200 microM. Weak dissociation constants of 100-200 microM were found. Quantitative EPR analysis of the six-line free Mn(2+) signal revealed that up to seven manganese ions were associated with the reaction centers at a 1 mM concentration of manganese. The association and the electron transfer between manganese and the reaction centers could be inhibited by Ca(2+) and Na(+) ions. The ability of reaction centers with high potentials to oxidize manganese suggests that manganese oxidation could have preceded water oxidation in the evolutionary development of photosystem II.  相似文献   

10.
Key features differentiating the coordination environment of the two irons in the mixed-valent (Fe(2+),Fe(3+)) diiron site of Desulfovibrio vulgaris rubrerythrin (Rbr(mv)) were determined by continuous wave (CW) and pulsed ENDOR spectroscopy at 35GHz. (14)N ENDOR evidence indicates that a nitrogen is bound only to the Fe(2+) ion of the mixed-valent site. Assuming that this nitrogen is from His131Ndelta, the same one that furnishes an iron ligand in the crystal structure of the diferric site, the ENDOR data allow us to specify the Fe(2+) and Fe(3+) positions within the molecular reference frame. In addition, the (1,2)H ENDOR on Rbr(mv) indicates the presence of a solvent-derived aqua/hydroxo ligand bound either terminally or in a bridging mode to Fe(3+) in the mixed-valent site. The relatively large g anisotropy of Rbr(mv) and weak antiferromagnetic coupling, J approximately -8 cm(-)(1) (in the 2JS(1)*S(2) formalism), between the irons is more consistent with a bridging than terminal hydroxo ligand. gamma-Irradiation was used to cryoreduce Rbr at 77 K, thereby producing a mixed-valent diiron site [(Rbr(ox))(mv)] that retains the structure of the diferric site. The EPR spectrum of (Rbr(ox))(mv) was nearly identical to that of the as-isolated or chemically reduced samples. This near identity implies that the structure of the mixed-valent Rbr diiron site is essentially identical to that of the diferric site, except for protonation of the oxo bridge, which apparently occurred via a proton jump from hydrogen-bonded solvent at 77 K. The EPR spectrum of (Rbr(ox))(mv) thus supports the (14)N ENDOR-assigned His131 ligation to Fe(2+) and assignment of the solvent-derived ligand observed in the (1,2)H ENDOR to a hydroxo bridge between the irons of the mixed-valent diiron site.  相似文献   

11.
Soluble guanylate cyclase is an NO-sensing hemoprotein that serves as a NO receptor in NO-mediated signaling pathways. It has been believed that this enzyme displays no measurable affinity for O(2), thereby enabling the selective NO sensing in aerobic environments. Despite the physiological significance, the reactivity of the enzyme-heme for O(2) has not been examined in detail. In this paper we demonstrated that the high spin heme of the ferrous enzyme converted to a low spin oxyheme (Fe(2+)-O(2)) when frozen at 77 K in the presence of O(2). The ligation of O(2) was confirmed by EPR analyses using cobalt-substituted enzyme. The oxy form was produced also under solution conditions at -7 °C, with the extremely low affinity for O(2). The low O(2) affinity was not caused by a distal steric protein effect and by rupture of the Fe(2+)-proximal His bond as revealed by extended x-ray absorption fine structure. The midpoint potential of the enzyme-heme was +187 mV, which is the most positive among high spin protoheme-hemoproteins. This observation implies that the electron density of the ferrous heme iron is relatively low by comparison to those of other hemoproteins, presumably due to the weak Fe(2+)-proximal His bond. Based on our results, we propose that the weak Fe(2+)-proximal His bond is a key determinant for the low O(2) affinity of the heme moiety of soluble guanylate cyclase.  相似文献   

12.
The regulation of heme oxygenase (HO) activity and its dependence on iron was studied in bovine aortic endothelial cells (BAEC) subjected to hypoxia-reoxygenation (H/R). HO activity was induced by hypoxia (10 h) and continued to increase during the reoxygenation phase. HO-1 protein levels were strongly induced by hypoxia from undetectable levels and remained elevated at least 8 h postreoxygenation. Addition of the Fe(3+) chelator desferrioxamine mesylate (DFO) or the Fe(2+) chelator o-phenanthroline during hypoxia alone or during the entire H/R period inhibited the induction of HO activity and HO-1 protein levels. However, DFO had no effect and o-phenanthroline had a partial inhibitory effect on HO activity and protein levels when added only during reoxygenation. Loading of BAEC with Fe(3+) enhanced the activation of the HO-1 gene by H/R, whereas loading with L-aminolevulinic acid, which stimulates heme synthesis, had little effect. These results suggest that chelatable iron participates in regulating HO expression during hypoxia.  相似文献   

13.
An analysis of the H(2)O(2)-induced breakdown and transformation of different keto-monosaccharides at physiological concentrations reveals that glycolate and other short-chained carbohydrates and organic acids are produced. Depletion of monosaccharides and glycolate synthesis occurs at increased rates as the length of the carbohydrate chain is decreased, and is significantly increased in the presence of trace amounts of Fe(2+) ions (10 microM). Rates of monosaccharide depletion (initial concentration of 3 mM) observed were up to 1.55 mmol h(-1) in the case of fructose, and 2.59 mmol h(-1) in the case of dihydroxyacetone, depending upon pH, H(2)O(2) concentration, temperature and the presence or absence of catalytic amounts of Fe(2+). Glycolate was produced by dihydroxyacetone cleavage at rates up to 0.45 mmol h(-1) in the absence, and up to 1.88 mmol h(-1) in the presence of Fe(2+) ions (pH 8). Besides glycolate, other sugars (ribose, glyceraldehyde, glucose), glucitol (sorbitol) and organic acids (formic and 2-oxogluconic acid) were produced in such H(2)O(2)-induced reactions with fructose or dihydroxyacetone. EPR measurements demonstrated the participation of the OH radical, especially at higher pH. Presence of metal ions at higher pH values, resulting in increased glycolate synthesis, was accompanied by enhanced hydroxyl radical generation. Observed changes in intensity of DEPMPO-OH signals recorded from dihydroxyacetone and fructose reactions demonstrate a strong correlation with changes in glycolate yield, suggesting that OH radical formation enhances glycolate synthesis. The results presented suggest that different mechanisms are responsible for the cleavage or other reactions (isomerisation, auto- or free-radical-mediated oxidation) of keto-monosaccharides depending of experimental conditions.  相似文献   

14.
Iron and aconitase activity   总被引:5,自引:2,他引:3       下载免费PDF全文
Aconitase activated with Fe(2+), cysteine and ascorbate incorporates 1 g-atom of Fe(2+)/mol. Loss of this Fe(2+) by transfer to ferrozine, a Fe(2+) chelator, results in loss of activity. Ascorbate increases the rate of transfer of the essential Fe(2+) whereas citrate retards the rate of transfer. Transfer of Fe(2+) from inactive aconitase, 2 g-atoms of Fe/mol, can be accomplished in the presence of urea and ascorbate. The correlation of activity with the presence of an added g-atom of Fe(2+)/mol leads to the conclusion that active aconitase has only one active site per mol.  相似文献   

15.
In this study, we attempted to elucidate the metabolic pathway and enzymes actually involved in oxalate formation from glycolate in rat and human liver. In rat liver, the formation of oxalate from glycolate appeared to take place predominantly via glyoxylate. The oxalate formation from glycolate observed with crude enzyme preparations was almost entirely accounted for by the sequential actions of glycolate oxidase and xanthine oxidase (XOD) or lactate dehydrogenase (LDH). Under the conditions used, no significant activity was attributable to glycolate dehydrogenase, an enzyme reported to catalyze the direct oxidation of glycolate to oxalate. Among the three enzymes known to catalyze the oxidation of glyoxylate to oxalate, glycolate oxidase and XOD showed much lower activities (a higher Km and lower Vmax) toward glyoxylate than those with the respective primary substrates. As to LDH, none of the LDH subunit-deficient patients examined showed profoundly lowered urinary oxalate excretion. Based on the results obtained, the presumed efficacies in vivo of individual enzymes, as catalysts of glyoxylate oxidation, and the in vivo conditions assumed to allow their catalysis of oxalate production are discussed.  相似文献   

16.
Oxalate synthesis in human hepatocytes is not well defined despite the clinical significance of its overproduction in diseases such as the primary hyperoxalurias. To further define these steps, the metabolism to oxalate of the oxalate precursors glycolate and glyoxylate and the possible pathways involved were examined in HepG2 cells. These cells were found to contain oxalate, glyoxylate, and glycolate as intracellular metabolites and to excrete oxalate and glycolate into the medium. Glycolate was taken up more effectively by cells than glyoxylate, but glyoxylate was more efficiently converted to oxalate. Oxalate was formed from exogenous glycolate only when cells were exposed to high concentrations. Peroxisomes in HepG2 cells, in contrast to those in human hepatocytes, were not involved in glycolate metabolism. Incubations with purified lactate dehydrogenase suggested that this enzyme was responsible for the metabolism of glycolate to oxalate in HepG2 cells. The formation of 14C-labeled glycine from 14C-labeled glycolate was observed only when cell membranes were permeabilized with Triton X-100. These results imply that peroxisome permeability to glycolate is restricted in these cells. Mitochondria, which produce glyoxylate from hydroxyproline metabolism, contained both alanine:glyoxylate aminotransferase (AGT)2 and glyoxylate reductase activities, which can convert glyoxylate to glycine and glycolate, respectively. Expression of AGT2 mRNA in HepG2 cells was confirmed by RT-PCR. These results indicate that HepG2 cells will be useful in clarifying the nonperoxisomal metabolism associated with oxalate synthesis in human hepatocytes. liver; peroxisomes; hepatocytes; hyperoxaluria; alanine:glyoxylate aminotransferase; glyoxylate reductase  相似文献   

17.
The crystal structure of the iron-free (apo) form of the Haemophilus influenzae Fe(3+)-binding protein (hFbp) has been determined to 1.75 A resolution. Information from this structure complements that derived from the holo structure with respect to the delineation of the process of iron binding and release. A 21 degrees rotation separates the two structural domains when the apo form is compared with the holo conformer, indicating that upon release of iron, the protein undergoes a change in conformation by bending about the central beta-sheet hinge. A surprising finding in the apo-hFbp structure was that the ternary binding site anion, observed in the crystals as phosphate, remained bound. In solution, apo-hFbp bound phosphate with an affinity K(d) of 2.3 x 10(-3) M. The presence of this ternary binding site anion appears to arrange the C-terminal iron-binding residues conducive to complementary binding to Fe(3+), while residues in the N-terminal binding domain must undergo induced fit to accommodate the Fe(3+) ligand. These observations suggest a binding process, the first step of which is the binding of a synergistic anion such as phosphate to the C-terminal domain. Next, iron binds to the preordered half-site on the C-terminal domain. Finally, the presence of iron organizes the N-terminal half-site and closes the interdomain hinge. The use of the synergistic anion and this iron binding process results in an extremely high affinity of the Fe(3+)-binding proteins for Fe(3+) (nFbp K'(eff) = 2.4 x 10(18) M(-1)). This high-affinity ligand binding process is unique among the family of bacterial periplasmic binding proteins and has interesting implications in the mechanism of iron removal from the Fe(3+)-binding proteins during FbpABC-mediated iron transport across the cytoplasmic membrane.  相似文献   

18.
Phosphotriesterase, isolated from the soil-dwelling bacterium Pseudomonas diminuta, catalyzes the detoxification of organophosphate-based insecticides and chemical warfare agents. The enzyme has attracted significant research attention in light of its possible employment as a bioremediation tool. As naturally isolated, the enzyme is dimeric. Each subunit contains a binuclear zinc center that is situated at the C-terminal portion of a "TIM" barrel motif. The two zincs are separated by approximately 3.4 A and coordinated to the protein via the side chains of His 55, His 57, His 201, His 230, Asp 301, and a carboxylated Lys 169. Both Lys 169 and a water molecule (or hydroxide ion) serve to bridge the two zinc ions together. Interestingly, these metals can be replaced with cadmium or manganese ions without loss of enzymatic activity. Here we describe the three-dimensional structures of the Zn(2+)/Zn(2+)-, Zn(2+)/Cd(2+)-, Cd(2+)/Cd(2+)-, and Mn(2+)/Mn(2+)-substituted forms of phosphotriesterase determined and refined to a nominal resolution of 1.3 A. In each case, the more buried metal ion, referred to as the alpha-metal, is surrounded by ligands in a trigonal bipyramidal ligation sphere. For the more solvent-exposed or beta-metal ion, however, the observed coordination spheres are either octahedral (in the Cd(2+)/Cd(2+)-, Mn(2+)/Mn(2+)-, and the mixed Zn(2+)/Cd(2+)-species) or trigonal bipyramidal (in the Zn(2+)/Zn(2+)-protein). By measuring the anomalous X-ray data from crystals of the Zn(2+)/Cd(2+)-species, it has been possible to determine that the alpha-metal ion is zinc and the beta-site is occupied by cadmium.  相似文献   

19.
Hinckley GT  Frey PA 《Biochemistry》2006,45(10):3219-3225
Lysine 2,3-aminomutase (LAM) catalyzes the interconversion of l-lysine and l-beta-lysine by a free radical mechanism. The 5'-deoxyadenosyl radical derived from the reductive cleavage of S-adenosyl-l-methionine (SAM) initiates substrate-radical formation. The [4Fe-4S](1+) cluster in LAM is the one-electron source in the reductive cleavage of SAM, which is directly ligated to the unique iron site in the cluster. We here report the midpoint reduction potentials of the [4Fe-4S](2+/1+) couple in the presence of SAM, S-adenosyl-l-homocysteine (SAH), or 5'-{N-[(3S)-3-aminocarboxypropyl]-N-methylamino}-5'-deoxyadenosine (azaSAM) as measured by spectroelectrochemistry. The reduction potentials are -430 +/- 2 mV in the presence of SAM, -460 +/- 3 mV in the presence of SAH, and -497 +/- 10 mV in the presence of azaSAM. In the absence of SAM or an analogue and the presence of dithiothreitol, dihydrolipoate, or cysteine as ligands to the unique iron, the midpoint potentials are -479 +/- 5, -516 +/- 5, and -484 +/- 3 mV, respectively. LAM is a member of the radical SAM superfamily of enzymes, in which the CxxxCxxC motif donates three thiolate ligands to iron in the [4Fe-4S] cluster and SAM donates the alpha-amino and alpha-carboxylate groups of the methionyl moiety as ligands to the fourth iron. The results show the reduction potentials in the midrange for ferredoxin-like [4Fe-4S] clusters. They show that SAM elevates the reduction potential by 86 mV relative to that of dihydrolipoate as the cluster ligand. This difference accounts for the SAM-dependent reduction of the [4Fe-4S](2+) cluster by dithionite reported earlier. Analogues of SAM have a weakened capacity to raise the potential. We conclude that the midpoint reduction potential of the cluster ligated to SAM is 1.2 V less negative than the half-wave potential for the one-electron reductive cleavage of simple alkylsulfonium ions in aqueous solution. The energetic barrier in the reductive cleavage of SAM may be overcome through the use of binding energy.  相似文献   

20.
The terminal electron acceptor of Photosystem II, PSII, is a linear complex consisting of a primary quinone, a non-heme iron(II), and a secondary quinone, Q(A)Fe(2+)Q(B). The complex is a sensitive site of PSII, where electron transfer is modulated by environmental factors and notably by bicarbonate. Earlier studies showed that NO and other small molecules (CN(-), F(-), carboxylate anions) bind reversibly on the non-heme iron in competition with bicarbonate. In the present study, we report on an unusual new mode of transient binding of NO, which is favored in the light-reduced state (Q(A)(-)Fe(2+)Q(B)) of the complex. The related observations are summarized as follows: (i) Incubation with NO at -30 degrees C, following light-induced charge separation, results in the evolution of a new EPR signal at g = 2.016. The signal correlates with the reduced state Q(A)(-)Fe(2+) of the iron-quinone complex. (ii) Cyanide, at low concentrations, converts the signal to a more rhombic form with g values at 2.027 (peak) and 1.976 (valley), while at high concentrations it inhibits formation of the signals. (iii) Electron spin-echo envelope modulation (ESEEM) experiments show the existence of two protein (14)N nuclei coupled to electron spin. These two nitrogens have been detected consistently in the environment of the semiquinone Q(A)(-) in a number of PSII preparations. (iv) NO does not directly contribute to the signals, as indicated by the absence of a detectable isotopic effect ((15)NO vs (14)NO) in cw EPR. (v) A third signal with g values (2.05, 2.03, 2.01) identical to those of an Fe(NO)(2)(imidazole) synthetic complex develops slowly in the dark, or faster following illumination. (vi) In comparison with the untreated Q(A)(-)Fe(2+) complex, the present signals not only are confined to a narrow spectral region but also saturate at low microwave power. At 11 K the g = 2.016 signal saturates with a P(1/2) of 110 microW and the g = 2.027/1.976 signal with a P(1/2) of 10 microW. (vii) The spectral shape and spin concentration of these signals is successfully reproduced, assuming a weak magnetic interaction (J values in the range 0.025-0.05 cm(-)(1)) between an iron-NO complex with total spin of (1)/(2) and the spin, (1)/(2), of the semiquinone, Q(A)(-). The different modes of binding of NO to the non-heme iron are examined in the context of a molecular model. An important aspect of the model is a trans influence of Q(A) reduction on the bicarbonate ligation to the iron, transmitted via H-bonding of Q(A) with an imidazole ligand to the iron.  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