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 62 毫秒
1.
The reaction of nitrite at pH 5.0-7.0 with the deoxyhaemocyanin of a mollusc, the Roman snail (Helix pomatia), yielded nitrosylhaemocyanin (CuIA.NO+ CuIIB), in contrast with the formation of methaemocyanin with the deoxyhaemocyanin of the crustacean Astacus leptodactylus (mud crayfish). With Helix haemocyanin 1 NO was thereby liberated per active site, as shown by m.s., as against 2 NO with Astacus haemocyanin. Helix nitrosylhaemocyanin was characterized in c.d. by the negative extremum at 336 nm (CuIA.NO+) and by the mononuclear e.p.r. signal at g = 2 (CuIIB). Binuclear e.p.r. signals have been observed after the addition of nitrite to methaemocyanins. With Astacus methaemocyanin, no further reaction occurred, whereas with Helix methaemocyanin the mononuclear e.p.r. signal, characteristic for nitrosylhaemocyanin gradually appeared. This formation of Helix nitrosylhaemocyanin implicates the binding, most likely on CuIIA, of a second nitrite besides a bridging nitrite, so that a dismutation into NO and NO2 can occur there. A further dismutation of NO2 yields nitrite and nitrate. The formation of the latter was demonstrated by Raman spectrometry. The reaction rate of Helix methaemocyanin with nitrite decreased with increasing pH according to the Henderson-Hasselbalch equation with a pKa value of 6.77, attributed to a mu-aquo bridging ligand, which can be exchanged for nitrite, in equilibrium with a mu-hydroxo ligand which cannot. These data also favour the formulation of the final reaction product as nitrosylhaemocyanin instead of semi-methaemocyanin, with or without bound nitrite.  相似文献   

2.
The reaction of hydroxyurea with the oxyhaemocyanins of Astacus leptodactylus and Helix pomatia yielded methaemocyanins which could be regenerated with hydroxylamine. Hydroxyurea did not react with Astacus methaemocyanin, but quantitatively regenerated Helix methaemocyanin under N2. The reaction of hydroxyurea with Helix haemocyanin at pH 5.7 under air thus led to a steady state, with an oxyhaemocyanin/methaemocyanin ratio of 2.05:1.  相似文献   

3.
The rate of the reaction of Astacus leptodactylus methaemocyanin with NO follows the Henderson-Hasselbalch equation with a pKa of 5.85, suggesting that one imidazole ligand of Cu was exchanged for NO. The reaction is blocked by F- as a bridging ligand. The same imidazole residue might be responsible for the decomposition of nitrosylhaemocyanin, [Cu1NO+CuII], with an unlocated binding site for NO, into methaemocyanin and NO, as the rate increase with pH. NO could react preferentially with CuA of Helix pomatia methaemocyanin, CuA'IICuBII, as it possibly has only two histidine ligands instead of the three of CuA in Astacus haemocyanin. This difference might explain the higher formation rate and the much greater stability of Helix nitrosylhaemocyanin. The fast reaction is governed by a pKa of 6.80, probably of a bridging mu-aquo ligand. With F- or a mu-hydroxo bridging ligand a low reaction rate was still observed, in contrast with Astacus methaemocyanin. Helix nitrosylhaemocyanin was transformed by N3- into methaemocyanin with the liberation of N2 and N2O. This methaemocyanin could almost quantitatively be regenerated with H2O2. Helix nitrosylhaemocyanin was only partially regenerated by a direct treatment with H2O2 and almost quantitatively by HONH2 in a similar fairly fast reaction, followed by a much slower one.  相似文献   

4.
Deoxyhaemocyanin, treated with NO under strictly anaerobic conditions, yielded methaemocyanin and N2O in a fast reaction. In a further slow reaction this methaemocyanin lost its triplet electron paramagnetic resonance (EPR) signal at g = 4 and yielded a nitrosyl derivative with a characteristic g = 2 Cu(II) EPR signal, indicating the binding of a single NO per copper pair. Thus under strictly anaerobic conditions deoxyhaemocyanin and methaemocyanin, treated with NO, gave the same derivative as shown by circular dichroism and EPR spectra. Methaemocyanin yielded, moreover, reversibly a nitrite derivative, characterized by a triplet signal at g = 4 with 7 hyperfine lines.  相似文献   

5.
During the reaction of oxyhemoglobin (HbO2) with nitrite, the concentration of residual nitrite, nitrate, oxygen, and methemoglobin (Hb+) was determined successively. The results obtained at various pH values indicate the following stoichiometry for the overall reaction: 4HbO2 + 4NO2- 4H+ leads to 4Hb+ + 4NO3- + O2 + 2H2 O (Hb denotes hemoglobin monomer). NO2- binds with methemoglobin noncooperatively with a binding constant of 340 M-1 at pH 7.4 and 25 degrees C. Thus, the major part of Hb+ produced is aquomethemoglobin, not methemoglobin nitrite, when less than 2 equivalents of nitrite is used for the oxidation.  相似文献   

6.
The active site amino acid residues of lignin peroxidase are homologous to those of other peroxidases; however, in contrast to other peroxidases, no pH dependence is observed for the reaction of ferric lignin peroxidase with H2O2 to form compound I (Andrawis, A., Johnson, K.A., and Tien, M. (1988) J. Biol. Chem. 263, 1195-1198). Chloride binding is used in the present study to investigate this reaction further. Chloride binds to lignin peroxidase at the same site as cyanide and hydrogen peroxide. This is indicated by the following. 1) Chloride competes with cyanide in binding to lignin peroxidase. 2) Chloride is a competitive inhibitor of lignin peroxidase with respect to H2O2. The inhibition constant (Ki) is equal to the dissociation constant (Kd) of chloride at all pH values studied. Chloride binding is pH dependent: chloride binds only to the protonated form of lignin peroxidase. Transient-state kinetic studies demonstrate that chloride inhibits lignin peroxidase compound I formation in a pH-dependent manner with maximum inhibition at low pH. An apparent pKa was calculated at each chloride concentration; the pKa increased as the chloride concentration increased. Extrapolation to zero chloride concentration allowed us to estimate the intrinsic pKa for the ionization in the lignin peroxidase active site. The results reported here provide evidence that an acidic ionizable group (pKa approximately 1) at the active site controls both lignin peroxidase compound I formation and chloride binding. We propose that the mechanism for lignin peroxidase compound I formation is similar to that of other peroxidases in that it requires the deprotonated form of an ionizable group near the active site.  相似文献   

7.
Q Su  J P Klinman 《Biochemistry》1999,38(26):8572-8581
Glucose oxidase catalyzes the oxidation of glucose by molecular dioxygen, forming gluconolactone and hydrogen peroxide. A series of probes have been applied to investigate the activation of dioxygen in the oxidative half-reaction, including pH dependence, viscosity effects, 18O isotope effects, and solvent isotope effects on the kinetic parameter Vmax/Km(O2). The pH profile of Vmax/Km(O2) exhibits a pKa of 7.9 +/- 0.1, with the protonated enzyme form more reactive by 2 orders of magnitude. The effect of viscosogen on Vmax/Km(O2) reveals the surprising fact that the faster reaction at low pH (1.6 x 10(6) M-1 s-1) is actually less diffusion-controlled than the slow reaction at high pH (1.4 x 10(4) M-1 s-1); dioxygen reduction is almost fully diffusion-controlled at pH 9.8, while the extent of diffusion control decreases to 88% at pH 9.0 and 32% at pH 5.0, suggesting a transition of the first irreversible step from dioxygen binding at high pH to a later step at low pH. The puzzle is resolved by 18O isotope effects. 18(Vmax/Km) has been determined to be 1.028 +/- 0.002 at pH 5.0 and 1.027 +/- 0.001 at pH 9.0, indicating that a significant O-O bond order decrease accompanies the steps from dioxygen binding up to the first irreversible step at either pH. The results at high pH lead to an unequivocal mechanism; the rate-limiting step in Vmax/Km(O2) for the deprotonated enzyme is the first electron transfer from the reduced flavin to dioxygen, and this step accompanies binding of molecular dioxygen to the active site. In combination with the published structural data, a model is presented in which a protonated active site histidine at low pH accelerates the second-order rate constant for one electron transfer to dioxygen through electrostatic stabilization of the superoxide anion intermediate. Consistent with the proposed mechanisms for both high and low pH, solvent isotope effects indicate that proton transfer steps occur after the rate-limiting step(s). Kinetic simulations show that the model that is presented, although apparently in conflict with previous models for glucose oxidase, is in good agreement with previously published kinetic data for glucose oxidase. A role for electrostatic stabilization of the superoxide anion intermediate, as a general catalytic strategy in dioxygen-utilizing enzymes, is discussed.  相似文献   

8.
Many properties of copper-containing nitrite reductase are pH-dependent, such as gene expression, enzyme activity, and substrate affinity. Here we use x-ray diffraction to investigate the structural basis for the pH dependence of activity and nitrite affinity by examining the type 2 copper site and its immediate surroundings in nitrite reductase from Rhodobacter sphaeroides 2.4.3. At active pH the geometry of the substrate-free oxidized type 2 copper site shows a near perfect tetrahedral geometry as defined by the positions of its ligands. At higher pH values the most favorable copper site geometry is altered toward a more distorted tetrahedral geometry whereby the solvent ligand adopts a position opposite to that of the His-131 ligand. This pH-dependent variation in type 2 copper site geometry is discussed in light of recent computational results. When co-crystallized with substrate, nitrite is seen to bind in a bidentate fashion with its two oxygen atoms ligating the type 2 copper, overlapping with the positions occupied by the solvent ligand in the high and low pH structures. Fourier transformation infrared spectroscopy is used to assign the pH dependence of the binding of nitrite to the active site, and EPR spectroscopy is used to characterize the pH dependence of the reduction potential of the type 2 copper site. Taken together, these spectroscopic and structural observations help to explain the pH dependence of nitrite reductase, highlighting the subtle relationship between copper site geometry, nitrite affinity, and enzyme activity.  相似文献   

9.
The reaction of superoxide radical (O2-) with the heme protein indoleamine 2,3-dioxygenase has been investigated by the use of pulse radiolysis. In the absence of the substrate tryptophan (Trp), the ferric enzyme reacted quantitatively with O2- to form the oxygenated enzyme. The rate constant for the reaction (8.0 x 10(6) M-1 s-1 at pH 7.0) increased with a decrease in pH. In the presence of low concentrations of L-Trp (approximately 50 microM), under which the catalytic site of the ferric enzyme is greater than 99% Trp-free at pH 7.0, the only spectral species observed upon O2- binding was L-Trp-bound oxygenated enzyme, the ternary complex. This suggests that under the conditions employed O2- binds first to the ferric enzyme to form the oxygenated enzyme and is followed by rapid binding of L-Trp. It was also found that absorbance changes (delta A) for the enzyme after the pulse were significantly decreased when an increased L-Trp concentration was employed. A 50% decrease in delta A was caused with approximately 50 microM L-Trp at pH 7.0. Similar results were also observed with other indole derivatives with decreasing delta A values in the order of indole, 3-indoleethanol, alpha-methyl-DL-Trp, and D-Trp. These results suggest that there exists a binding site for these compounds in the dioxygenase different from the catalytic site for Trp and, most significantly, that binding of Trp to the effector binding site of the ferric enzyme markedly inhibits its reaction with O2-.  相似文献   

10.
The kinetics of the reaction of Helix pomatia haemocyanin with oxygen have been studied under conditions where ligand binding is co-operative (n = 4.5). The dissociation of oxygen from oxyhaemocyanin in the presence of sodium dithionite and the combination of deoxyhaemocyanin with oxygen were studied by the stopped-flow technique. The combination with oxygen, as well as the dissociation of oxyhaemocyanin, are clearly autocatalytic. The initial rate constant for oxygen combination to the fully deoxygenated state is 0.2 to 0.3 × 106m?1 s?1; during the course of the reaction the rate constant increases to a value higher than 106m?1s?1.The initial rate of oxygen dissociation from fully saturated haemocyanin is 10 s?1, increasing to about 30 s?1 as the reaction proceeds. Thus, both the combination and the dissociation rate constants contribute to the co-operativity of oxygen binding.Temperature-jump relaxation experiments were carried out at fractional oxygen saturations larger than 0.7. The dependence of the relaxation rate upon the concentration of the reactants indicates the presence of one principal bimolecular process. The calculated combination and dissociation rate constants for this process are: 3.8 × 106m?1 s?1 and 10 s?1, respectively. Evidence is presented which shows that the transition from the T-state to the R-state of the protein is relatively slow. Both the T and R-state seem to be largely stabilized at the expense of intermediate states.Under other conditions, where oxygen binding is non-co-operative, temperature-jump and stopped-flow experiments reveal considerable kinetic heterogeneity.  相似文献   

11.
Mitochondrial amidoxime reducing component (mARC) proteins are molybdopterin-containing enzymes of unclear physiological function. Both human isoforms mARC-1 and mARC-2 are able to catalyze the reduction of nitrite when they are in the reduced form. Moreover, our results indicate that mARC can generate nitric oxide (NO) from nitrite when forming an electron transfer chain with NADH, cytochrome b5, and NADH-dependent cytochrome b5 reductase. The rate of NO formation increases almost 3-fold when pH was lowered from 7.5 to 6.5. To determine if nitrite reduction is catalyzed by molybdenum in the active site of mARC-1, we mutated the putative active site cysteine residue (Cys-273), known to coordinate molybdenum binding. NO formation was abolished by the C273A mutation in mARC-1. Supplementation of transformed Escherichia coli with tungsten facilitated the replacement of molybdenum in recombinant mARC-1 and abolished NO formation. Therefore, we conclude that human mARC-1 and mARC-2 are capable of catalyzing reduction of nitrite to NO through reaction with its molybdenum cofactor. Finally, expression of mARC-1 in HEK cells using a lentivirus vector was used to confirm cellular nitrite reduction to NO. A comparison of NO formation profiles between mARC and xanthine oxidase reveals similar Kcat and Vmax values but more sustained NO formation from mARC, possibly because it is not vulnerable to autoinhibition via molybdenum desulfuration. The reduction of nitrite by mARC in the mitochondria may represent a new signaling pathway for NADH-dependent hypoxic NO production.  相似文献   

12.
《BBA》1986,848(1):1-7
An almost stoichiometric conversion of nitrite to nitrous oxide was observed during the nitrite reduction by Paracoccus denitrificans cells in a medium of pH 6.4. The N2O accumulated in the reaction medium and was decomposed only after nitrite had been consumed; when the pH of the medium was higher than 7.3–7.4, nitrous oxide did not accumulate. The activity of N2O reductase was, in the whole range of pH 6.4–9.2, higher than the activity of NO2 reductase, both activities showing the maximum at the pH higher than 8.0. Using an artificial donor, TMPD plus ascorbate, the maximum activity of NO2 reductase, but not N2O reductase was shifted by about two pH units to acidic region. The activity of nitrite reductase declined in the presence of N2O only at higher pH values. Cytochrome c, as a common electron donor for both N2O and NO2 reductase, was more oxidized at pH < 7.3 in the presence of NO2 than in the presence of N2O, the opposite being true at pH > 7.3. The increased flux of electrons to cytochrome c has for a constant pH value (6.4) no effect on their distribution over NO2 and N2O. The results indicate that the distribution of electrons in the terminal part is determined by the different pH optima for NO2 reductase and N2O reductase, and by a mutual dependence of activities of the two reductases due to the competition for redox equivalents from a substrate.  相似文献   

13.
Hemoglobin released in the circulation from ruptured red blood cells can be oxidized by hydrogen peroxide or peroxynitrite to generate the highly oxidizing iron(IV)oxo species HbFe(IV)z=O. Nitrogen monoxide, produced in large amounts by activated inducible nitric oxide synthase, can have indirect cytotoxic effects, mainly through the generation of peroxynitrite from its very fast reaction with superoxide. In the present work we have determined the rate constant for the reaction of HbFe(IV)z=O with NO(*), 2.4 x 10(7) M(-1)s(-1) at pH 7.0 and 20 degrees C. The reaction proceeds via the intermediate HbFe(III)ONO, which then dissociates to metHb and nitrite. As these products are not oxidizing and because of its large rate, the reaction of HbFe(IV)z=O with NO(*) may be important to remove the high valent form of hemoglobin, which has been proposed to be at least in part responsible for oxidative lesions. In addition, we have determined that the rate constant for the reaction of HbFe(IV)z=O with nitrite is significantly lower (7.5 x 10(2) M(-1)s(-1) at pH 7.0 and 20 degrees C), but increases with decreasing pH (1.8 x 10(3) M(-1)s(-1) at pH 6.4 and 20 degrees C). Thus, under acidic conditions as found in ischemic tissues, this reaction may also have a physiological relevance.  相似文献   

14.
The reaction of hydrogen peroxide with the copper-zinc bovine-liver superoxide dismutase at low molar ratios (0.2-20.0) of H2O2/active site between pH 7.3-10.0 leads to the loss of native enzyme as a distinct form monitored by electrophoresis. The pH dependence of the loss of native enzyme between 7.3 and 9.0 indicates the involvement of a conjugate base on the enzyme of pKa of 8.7 +/- 0.1. The rate of loss of the native enzyme is first order with respect to the concentration of both enzyme and hydrogen peroxide between pH 7.3 and 9.0 with no evidence for binding of peroxide. A second-order rate constant of 3.0 +/- 1.0 M-1 s-1 is obtained from these data. At pH 10.0 the reaction is first order with respect to enzyme concentration but saturable in H2O2. All data are consistent with the interpretation that H2O2 reacts with the enzyme at the lower pH where the reaction is dependent upon the conjugate base of a functional group on the enzyme. At the higher pH, the data are consistent with the reaction of HO2- and H2O2 with the dismutase. The dissociation constant for HO2- calculated from the kinetic data at pH 10.0 is between 25-50 microM and the rate constant for the breakdown of the HO2- dismutase complex is 1.10 + 0.05 x 10(-2) s-1. The change in the electrophoretic pattern at all pH values is accompanied by the loss of the ability of the enzyme to bind copper. Weakly bound or free copper can be detected using bathocuproine disulfonate. Furthermore copper-defficient forms of the enzyme can be detected by staining gels of the peroxide-treated dismutase with diethyldithiocarbamate.  相似文献   

15.
The rate of reaction of trioxodinitrate with reduced cytochrome oxidase d in membrane particles from Escherichia coli at pH 7 and 25 degrees C depends linearly upon [HN2O3-] over the concentration range studied (up to 0.05 mM) and is also first-order in cytochrome d. The known rate of decomposition of trioxodinitrate to give NO- and NO2- is about 4.5-times faster than the rate of reaction of reduced cytochrome d with trioxodinitrate, implying that cytochrome d reacts directly with NO-, with a trapping ratio of between 0.20 and 0.25, rather than with trioxodinitrate. The implications of the facile formation of the NO(-)-nitrosyl complex of cytochrome d for the mechanism of denitrification are discussed with particular reference to the mechanism of N-N bond formation. The reaction of reduced cytochrome d with nitrite (a decomposition product of trioxodinitrate) under these conditions is much slower than that with trioxodinitrate. The kinetics show a biphasic dependence of initial rate upon nitrite concentration. The rate data at low [NO2-] are consistent with saturation of a high affinity site for nitrite, having Vmax = 4.29.10(-9) M s-1 and Km = 0.034 mM. The existence of two binding sites for nitrite is consistent with the suggestion that the cytochrome bd complex contains two cytochrome d haems.  相似文献   

16.
Legler PM  Lee HC  Peisach J  Mildvan AS 《Biochemistry》2002,41(14):4655-4668
Escherichia coli GDP-mannose mannosyl hydrolase (GDPMH), a homodimer, catalyzes the hydrolysis of GDP-alpha-D-sugars to yield the beta-D-sugar and GDP by nucleophilic substitution with inversion at the C1' carbon of the sugar [Legler, P. M., Massiah, M. A., Bessman, M. J., and Mildvan, A. S. (2000) Biochemistry 39, 8603-8608]. GDPMH requires a divalent cation for activity such as Mn2+ or Mg2+, which yield similar kcat values of 0.15 and 0.13 s(-1), respectively, at 22 degrees C and pH 7.5. Kinetic analysis of the Mn2+-activated enzyme yielded a K(m) of free Mn2+ of 3.9 +/- 1.3 mM when extrapolated to zero substrate concentration (K(a)Mn2+), which tightened to 0.32 +/- 0.18 mM when extrapolated to infinite substrate concentration (K(m)Mn2+). Similarly, the K(m) of the substrate extrapolated to zero Mn2+ concentration (K(S)(GDPmann) = 1.9 +/- 0.5 mM) and to infinite Mn2+ concentration (K(m)(GDPmann) = 0.16 +/- 0.09 mM) showed an order of magnitude decrease at saturating Mn2+. Such mutual tightening of metal and substrate binding suggests the formation of an enzyme-metal-substrate bridge complex. Direct Mn2+ binding studies, monitoring the concentration of free Mn2+ by EPR and of bound Mn2+ by its enhanced paramagnetic effect on the longitudinal relaxation rate of water protons (PRR), detected three Mn2+ binding sites per enzyme monomer with an average dissociation constant (K(D)) of 3.2 +/- 1.0 mM, in agreement with the kinetically determined K(a)Mn2+. The enhancement factor (epsilon(b)) of 11.5 +/- 1.2 indicates solvent access to the enzyme-bound Mn2+ ions. No cross relaxation was detected among the three bound Mn2+ ions, suggesting them to be separated by at least 10 A. Such studies also yielded a weak dissociation constant for the binary Mn2+-GDP-mannose complex (K1 = 6.5 +/- 1.0 mM) which significantly exceeded the kinetically determined K(m) values of Mn2+, indicating the true substrate to be GDP-mannose rather than its Mn2+ complex. Substrate binding monitored by changes in 1H-15N HSQC spectra yielded a dissociation constant for the binary E-GDP-mannose complex (K(S)(GDPmann)) of 4.0 +/- 0.5 mM, comparable to the kinetically determined K(S) value (1.9 +/- 0.5 mM). To clarify the metal stoichiometry at the active site, product inhibition by GDP, a potent competitive inhibitor (K(I) = 46 +/- 27 microM), was studied. Binding studies revealed a weak, binary E-GDP complex (K(D)(GDP) = 9.4 +/- 3.2 mM) which tightened approximately 500-fold in the presence of Mn2+ to yield a ternary E-Mn2+-GDP complex with a dissociation constant, K3(GDP) = 18 +/- 9 microM, which overlaps with the K(I)(GDP). The tight binding of Mn2+ to 0.7 +/- 0.2 site per enzyme subunit in the ternary E-Mn2+-GDP complex (K(A)' = 15 microM) and the tight binding of GDP to 0.8 +/- 0.1 site per enzyme subunit in the ternary E-Mg2+-GDP complex (K3 < 0.5 mM) indicate a stoichiometry close to 1:1:1 at the active site. The decrease in the enhancement factor of the ternary E-Mn2+-GDP complex (epsilon(T) = 4.9 +/- 0.4) indicates decreased solvent access to the active site Mn2+, consistent with an E-Mn2+-GDP bridge complex. Fermi contact splitting (4.3 +/- 0.2 MHz) of the phosphorus signal in the ESEEM spectrum established the formation of an inner sphere E-Mn2+-GDP complex. The number of water molecules coordinated to Mn2+ in this ternary complex was determined by ESEEM studies in D2O to be two fewer than on the average Mn2+ in the binary E-Mn2+ complexes, consistent with bidentate coordination of enzyme-bound Mn2+ by GDP. Kinetic, metal binding, and GDP binding studies with Mg2+ yielded dissociation constants similar to those found with Mn2+. Hence, GDPMH requires one divalent cation per active site to promote catalysis by facilitating the departure of the GDP leaving group, unlike its homologues the MutT pyrophosphohydrolase, which requires two, or Ap4A pyrophosphatase, which requires three.  相似文献   

17.
Interaction of thiocyanate with horseradish peroxidase (HRP) was investigated by relaxation rate measurements (at 50.68 MHz) of the 15N resonance of thiocyanate nitrogen and by following the hyperfine shifted ring methyl proton resonances (at 500 MHz) of the heme group of SCN-.HRP solutions. At pH 4.0, the apparent dissociation constant (KD) for thiocyanate binding to HRP was deduced to be 158 mM from the relaxation rate measurements. Chemical shift changes of 1- and 8-ring methyl proton resonances in the presence of various amounts of thiocyanate at pH 4.0 yielded KD values of 166 and 136 mM, respectively. From the pH dependence of KD and the 15N resonance line width, it was observed that thiocyanate binds to HRP only under acidic conditions (pH less than 6). The binding was found to be facilitated by protonation of an acid group on the enzyme with pKa 4.0. The pH dependence of the 15N line width as well as the apparent dissociation constant were quantitatively analyzed on the basis of a reaction scheme in which thiocyanate in deprotonated ionic form binds to the enzyme in protonated acidic form. The KD for thiocyanate binding to HRP was also evaluated in the presence of an excess of exogenous substrates such as resorcinol, cyanide, and iodide ions. It was found that the presence of cyanide (which binds to heme iron at the sixth coordination position) and resorcinol did not have any effect on the binding of thiocyanate, indicating that the binding site of the thiocyanate ion is located away from the ferric center as well as from the aromatic donor binding site. The KD in the presence of iodide, however, showed that iodide competes with thiocyanate for binding at the same site. The distance of the bound thiocyanate ion from the ferric center was deduced from the 15N relaxation time measurements and was found to be a 6.8 A. From the distance as well as the change in the chemical shifts and line width of 1- and 8-methyl proton resonances, it is suggested that the binding site of thiocyanate may be located near heme, placed symmetrically with respect to 1- and 8-methyl groups of the heme of HRP. Similarity in the modes of binding of iodide and thiocyanate suggests that the oxidation of thiocyanate ion by H2O2 may also proceed via the two-electron transfer pathway under acidic conditions, as is the case for iodide.  相似文献   

18.
The reaction of bovine heart ferrocytochrome c with nitrite was studied under various conditions. The reaction product was ferricytochrome c at around pH 5, whereas at around pH 3 it was Compound I, characterized by twin peaks at 529 and 563 nm of equal intensity. However, ferrocytochrome c decreased obeying first-order kinetics over the pH range examined, irrespective of the presence or absence of molecular oxygen. The apparent first-order rate constant was proportional to the square of the nitrite concentration at pH 4.4 and it increased as the pH was lowered. At pH 3 the reaction was so rapid that it had to be followed by stopped-flow and rapid-scanning techniques. The apparent rate constant at this pH was found to increase linearly with the nitrite concentration. Based on these results the active species of nitrite was concluded to be dinitrogen trioxide at pH 4.4 and nitrosonium ion, no+, at pH 3. Compound II was formed by reaction of ferrocytochrome c and NO gas at acidic and alkaline pH values. The absorption peaks were at 533 and 563 nm at pH 3, and at 538 and 567 nm at pH 12.9. This compound was also formed by reducing Compound I with reductants. Compound I prepared from ferricytochrome c and NO was stable below pH 6. However, appreciable absorption peaks for ferrocytochrome c appeared between pH 8 and 10, because Compound I was dissociated into ferrocytochrome c and NO+, and because ferrocytochrome c thus formed reacted with NO very slowly in this pH region. Saccharomyces ferricytochrome c under NO gas behaved differently from mammalian cytochrome, indicating the significance of the nature of the heme environment in determing the reactivity. Only at extreme pH values was Compound II formed exclusively and persisted. A model system for dissimilatory nitrite reductase was constructed by using bovine heart cytochrome c, nitrite and NADH plus PMS at pH 3.3, and a scheme involving cyclic turnover of ferrocytochrome c, Compound I and Compound II is presented, with kinetic parameters.  相似文献   

19.
We have identified a binding site for tissue inhibitors of metalloproteinases 2 (TIMP-2) on human 72-kDa gelatinase that is distinct from the active site. 72-kDa progelatinase is found in a complex with TIMP-2 in the medium of cultured cells and can be activated with organomercurial compounds to yield a gelatinolytic proteinase that remains bound to TIMP-2. Removal of TIMP-2 from 72-kDa progelatinase by reverse-phase high performance liquid chromatography, followed by reconstitution of the progelatinase in neutral pH buffer, results in autocatalytic activation. When samples of autoactivated gelatinase were blotted onto nitrocellulose, then probed with 125I-TIMP-2, we found a 29-kDa peptide that was capable of binding TIMP-2. We isolated this fragment and identified it as the region of gelatinase from amino acid 414 to the carboxyl terminus in the primary amino acid sequence of progelatinase. This portion of the molecule does not contain the putative zinc- or gelatin-binding sites and is proteolytically inactive. Incubation of 125I-TIMP-2 with 72-kDa progelatinase-TIMP-2 complexes resulted in a concentration-dependent exchange of labeled TIMP-2 with unlabeled TIMP-2, in both the presence and absence of the metalloproteinase inhibitor 1,10-phenanthroline. Saturation binding kinetics for the active site of 72-kDa gelatinase were measured in pools of the 43-kDa active fragment that results from the autoactivation of 72-kDa progelatinase; this fragment has no carboxyl-terminal TIMP-2 binding capability. Binding of 125I-TIMP-2 to the active site was completely inhibited by 1,10-phenanthroline. Binding kinetics for the putative stabilization site were determined with isolated 72-kDa progelatinase. In the presence of 1,10-phenanthroline, 72-kDa progelatinase bound 125I-TIMP-2 but not 125I-TIMP-1. Scatchard analysis yielded an approximate dissociation constant (Kd) of 0.72 nM for the active site and 0.42 nM for the stabilization site.  相似文献   

20.
Cytochrome cd(1) nitrite reductase from Pseudomonas stutzeri catalyzes the one electron reduction of nitrite to nitric oxide. It is a homodimer, each monomer containing one heme-c and one heme-d(1), the former being the electron uptake site while the latter is the nitrite reduction site. Hence, internal electron transfer between these sites is an inherent element in the catalytic cycle of this enzyme. We have investigated the internal electron transfer reaction employing pulse radiolytically produced N-methyl nicotinamide radicals as reductant which reacts solely with the heme-c in an essentially diffusion controlled process. Following this initial step, the reduction equivalent is equilibrating between the c and d(1) heme sites in a unimolecular process (k=23 s(-1), 298 K, pH 7.0) and an equilibrium constant of 1.0. The temperature dependence of this internal electron transfer process has been determined over a 277-313 K temperature range and yielded both equilibrium standard enthalpy and entropy changes as well as activation parameters of the specific rate constants. The significance of these parameters obtained at low degree of reduction of the enzyme is discussed and compared with earlier studies on cd(1) nitrite reductases from other sources.  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