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 46 毫秒
1.
Tyrosine hydroxylase (TyrH) is a pterin-dependent enzyme that catalyzes the hydroxylation of tyrosine to form dihydroxyphenylalanine. The oxidation state of the active site iron atom plays a central role in the regulation of the enzyme. The kinetics of reduction of ferric TyrH by several reductants were determined by anaerobic stopped-flow spectroscopy. Anaerobic rapid freeze-quench EPR confirmed that the change in the near-UV absorbance of TyrH upon adding reductant corresponded to iron reduction. Tetrahydrobiopterin reduces wild-type TyrH following a simple second-order mechanism with a rate constant of 2.8 +/- 0.1 mM(-)(1) s(-)(1). 6-Methyltetrahydropterin reduces the ferric enzyme with a second-order rate constant of 6.1 +/- 0.1 mM(-)(1) s(-)(1) and exhibits saturation kinetics. No EPR signal for a radical intermediate was detected. Ascorbate, glutathione, and 1,4-benzoquinone all reduce ferric TyrH, but much more slowly than tetrahydrobiopterin, suggesting that the pterin is a physiological reductant. E332A TyrH, which has an elevated K(m) for tetrahydropterin in the catalytic reaction, is reduced by tetrahydropterins with the same kinetic parameters as those of the wild-type enzyme, suggesting that BH(4) does not bind in the catalytic conformation during the reduction. Oxidation of ferrous TyrH by molecular oxygen can be described as a single-step second-order reaction, with a rate constant of 210 mM(-)(1) s(-)(1). S40E TyrH, which mimics the phosphorylated state of the enzyme, has oxidation and reduction kinetics similar to those of the wild-type enzyme, suggesting that phosphorylation does not directly regulate the interconversion of the ferric and ferrous forms.  相似文献   

2.
Membrane ghosts derived from chromaffin vesicles of bovine adrenal medullas have been used to examine the mechanism of reduction of dopamine beta-monooxygenase in its compartmentalized state. The rate of the dopamine beta-monooxygenase-catalyzed conversion of dopamine to norepinephrine is greatly stimulated by the presence of ATP, reflecting substrate hydroxylation on the ghost interior subsequent to the active transport of dopamine. We demonstrate a 2-3-fold increase in the turnover rate for ghosts resealed with 0.2-2 mM potassium ferrocyanide, conditions leading to a slight decrease in the rate of dopamine transport. These data provide the first evidence that an intravesicular pool of reductant can activate dopamine beta-monooxygenase, as required by models in which vesicular ascorbate behaves as enzyme reductant. Although there is sufficient catecholamine (endogenous plus substrate) to keep internal ferrocyanide reduced in these experiments, an additional 2-3-fold increase in turnover occurs in the presence of 0.2-2 mM ascorbate on the ghost exterior. The magnitude of this activation is found to be constant at all concentrations of internal ferrocyanide (both below and above saturation), implying that reductants on opposite sides of the membrane behave independently. Replacement of ascorbate by potassium ferrocyanide as external reductant leads to almost identical results, and we are able to rule out an inward transport of dehydroascorbate as the source of activation by external ascorbate. We conclude that external reductants are capable of reducing membrane-bound dopamine beta-monooxygenase from the exterior face of the vesicle, either by direct reduction or through a membrane-bound mediator. It appears that two viable modes for reduction of dopamine beta-monooxygenase may exist in vivo, involving the reduction of membrane-bound enzyme by cytosolic ascorbate as well as the reduction of soluble enzyme by the pool of intravesicular ascorbate present in chromaffin vesicles.  相似文献   

3.
The kinetics of electron transfer between cytochrome-c oxidase and ruthenium hexamine has been characterized using the native enzyme or its cyanide complex either solubilized by detergent (soluble cytochrome oxidase) or reconstituted into artificial phospholipid vesicles (cytochrome oxidase-containing vesicles). Ru(NH3)2+6 (Ru(II] reduces oxidized cytochrome a, following (by-and-large) bimolecular kinetics; the second order rate constant using the cyanide complex of the enzyme is 1.5 x 10(6) M-1 s-1, for the enzyme in detergent, and slightly higher for COV. In the case of COV the kinetics are not affected by the addition of ionophores. Upon mixing fully reduced cytochrome oxidase with oxygen (in the presence of excess reductants), the oxidation leading to the pulsed enzyme is followed by a steady state phase and (eventually) by complete re-reduction. When the concentrations of dioxygen and oxidase are sufficiently low (micromolar range), the time course of oxidation can be resolved by stopped flow at room temperature, yielding an apparent bimolecular rate constant of 5 x 10(7) M-1 s-1. After exhaustion of oxygen and end of steady state, re-reduction of the pulsed enzyme by the excess Ru(II) is observed; the concentration dependence shows that the rate of re-reduction is limited at 3 s-1 in detergent; this limiting value is assigned to the intramolecular electron transfer process from cytochrome a-Cua to the binuclear center. Using the reconstituted enzyme, the internal electron transfer step is sensitive to ionophores, increasing from 2-3 to 7-8 s-1 upon addition of valinomycin and carbonyl cyanide m-chlorophenylhydrazone. This finding indicates for the first time an effect of the electrochemical potential across the membrane on the internal electron transfer rate; the results are compared with expectations based on the hypothesis formulated by Brunori et al. (Brunori, M., Sarti, P., Colosimo, A., Antonini, G., Malatesta, F., Jones, M.G., and Wilson, M.T. (1985) EMBO J. 4, 2365-2368), and their bioenergetic relevance is discussed with reference to the proton pumping activity of the enzyme.  相似文献   

4.
We have investigated the formal reduction potentials (E degrees') of adrenodoxin with and without adrenodoxin reductase in order to elucidate the mechanism of electron transfer from adrenodoxin reductase (a flavoprotein) to adrenodoxin (an iron-sulfur protein). It was found by our spectropotentiostatic method that adrenodoxin showed no variation of E degrees' at different pH's in the absence of adrenodoxin reductase. The average E degrees' was -252 +/- 2 mV in the pH range between 6.0 and 8.3. In the presence of adrenodoxin reductase, adrenodoxin exhibited, on the other hand, a pH dependence of E degrees' at pH higher than 7.2 with a slope of -59 mV per pH unit: Adrenodoxin molecule possesses one protonation site with a pKa of 7.2. Cyclic voltammograms of adrenodoxin additionally revealed that the reoxidation reaction of reduced adrenodoxin is very slow in the absence of adrenodoxin reductase, but that it is readily reoxidized in the presence of adrenodoxin reductase.  相似文献   

5.
Azotobacter vinelandii hydrogenase (EC class 1.12), either purified or membrane-associated, was obtained aerobically in an inactive state. The kinetics of activation by treatment with a reductant (H2 or dithionite) were determined. Three distinct phases of the activation were observed. Aerobically prepared, inactive hydrogenase was insensitive to acetylene inhibition, but could be rendered acetylene-sensitive by reduction with dithionite. These findings indicate that acetylene inhibition of hydrogenase requires catalytically active enzyme.  相似文献   

6.
The reductant dependence of iron mobilization from isolated rabbit reticulocyte endosomes containing diferric transferrin is reported. The kinetic effects of acidification by a H(+)-ATPase are eliminated by incubating the endosomes at pH 6.0 in the presence of 15 microM FCCP to acidify the intravesicular milieu and to dissociate 59Fe(III) from transferrin. In the absence of reductants, iron is not released from the vesicles, and iron leakage is negligible. The second-order dependence of rate constants and amounts of 59Fe mobilized from endosomes using ascorbate, ferrocyanide, or NADH are consistent with reversible mechanisms. The estimated apparent first-order rate constant for mobilization by ascorbate is (2.7 +/- 0.4) x 10(-3) s-1 in contrast to (3.2 +/- 0.1) x 10(-4) s-1 for NADH and (3.5 +/- 0.6) x 10(-4) s-1 for ferrocyanide. These results support models where multiple reactions are involved in complex processes leading to iron transfer and membrane translocation. A type II NADH dehydrogenase (diaphorase) is present on the endosome outer membrane. The kinetics of extravesicular ferricyanide reduction indicate a bimolecular-bimolecular steady-state mechanism with substrate inhibition. Ferricyanide inhibition of 59Fe mobilization is not detected. Significant differences between mobilization and ferricyanide reduction kinetics indicate that the diaphorase is not involved in 59Fe(III) reduction. Sequential additions of NADH followed by ascorbate or vice versa indicate a minimum of two sites of 59Fe(III) residence; one site available to reducing equivalents from ascorbate and a different site available to NADH. Sequential additions using ferrocyanide and the other reductants suggest interactions among sites available for reduction. Inhibition of ascorbate-mediated mobilization by DCCD and enhancement of ferrocyanide and NADH-mediated mobilization suggest a role for a moiety with characteristics of a proton pore similar to that of the H(+)-ATPase. These data provide significant constraints on models of iron reduction, translocation, and mobilization by endocytic vesicles.  相似文献   

7.
The effect of complex formation between ferricytochrome c and cytochrome c peroxidase (Ferrocytochrome-c:hydrogen peroxide oxidoreductase, EC 1.11.1.5) on the reduction of cytochrome c by N,N,N',N'-tetramethyl-p-phenylenediamine (TMPD), reduced N-methylphenazonium methosulfate (PMSH), and ascorbate has been determined at low ionic strength (pH 7) and 25 degrees C. Complex formation with the peroxidase enhances the rate of ferricytochrome c reduction by the neutral reductants TMPD and PMSH. Under all experimental conditions investigated, complex formation with cytochrome c peroxidase inhibits the ascorbate reduction of ferricytochrome c. This inhibition is due to the unfavorable electrostatic interactions between the ascorbate dianion and the negatively charged cytochrome c-cytochrome c peroxidase complex. Corrections for the electrostatic term by extrapolating the data to infinite ionic strength suggest that ascorbate can reduce cytochrome c peroxidase-bound cytochrome c faster than free cytochrome c. Reduction of cytochrome c peroxidase Compound II by dicyanobis(1,10-phenanthroline)iron(II) (Fe(phen)2(CN)2) is essentially unaffected by complex formation between the enzyme and ferricytochrome c at low ionic strength (pH 6) and 25 degrees C. However, reduction of Compound II by the negatively changed tetracyano-(1,10-phenanthroline)iron(II) (Fe(phen)(CN)4) is enhanced in the presence of ferricytochrome c. This enhancement is due to the more favorable electrostatic interactions between the reductant and cytochrome c-cytochrome c peroxidase Compound II complex then for Compound II itself. These studies indicate that complex formation between cytochrome c and cytochrome c peroxidase does not sterically block the electron-transfer pathways from these small nonphysiological reductants to the hemes in these two proteins.  相似文献   

8.
Nitrogenase-catalyzed reactions using Ti(III) were examined under a wide variety of conditions to determine the suitability of Ti(III) to serve as a general nitrogenase reductant. Solutions prepared from H2-reduced TiCl3, aluminum-reduced TiCl3, TiCl2, evaporated TiCl3 from an HCl, solution, and TiF3 were evaluated as reductants. Three general types of reactivity were observed. The first showed that, below Ti(III) concentrations of about 0.50 mM, nitrogenase catalysis utilized Ti(III) in a first-order reaction. The second showed that, above 0.50 mM, the rate of nitrogenase catalysis was zero order in Ti(III), indicating the enzyme was saturated with this reductant. Above 2.0-5.0 mM, nitrogenase catalysis was inhibited by Ti(III) depending on the titanium source used for solution preparation. This inhibition was investigated and found to be independent of the buffer type and pH, while high salt and citrate concentrations caused moderate inhibition. [Ti(IV)] above 2.0-3.0 mM and [Ti(III)] above about 5.0 mM were inhibitory. ATP/2e values were 4-5 for [Ti(III)] at or below 1.0-2.0 mM, 2.0 from 5.0 to 7.0 mM Ti(III) where nitrogenase is not inhibited, and 2.0 above 7.0 mM Ti(III) where severe inhibition occurs. For nitrogenase-catalyzed reactions using Ti(III) as reductant, the potential of the solution changes with time as the Ti(III)/Ti(IV) ratio changes. From the change in the rate of product formation (Ti(III) disappearance) with change in solution potential, the rate of nitrogenase catalysis was determined as a function of solution potential. From such experiments, a midpoint turnover potential of -480 mV was determined for nitrogenase catalysis with an associated n = 2 value.  相似文献   

9.
Triapine (3-aminopyridine-2-carboxaldehyde thiosemicarbazone, 3-AP) is currently the most promising chemotherapeutic compound among the class of α-N-heterocyclic thiosemicarbazones. Here we report further insights into the mechanism(s) of anticancer drug activity and inhibition of mouse ribonucleotide reductase (RNR) by Triapine. In addition to the metal-free ligand, its iron(III), gallium(III), zinc(II) and copper(II) complexes were studied, aiming to correlate their cytotoxic activities with their effects on the diferric/tyrosyl radical center of the RNR enzyme in vitro. In this study we propose for the first time a potential specific binding pocket for Triapine on the surface of the mouse R2 RNR protein. In our mechanistic model, interaction with Triapine results in the labilization of the diferric center in the R2 protein. Subsequently the Triapine molecules act as iron chelators. In the absence of external reductants, and in presence of the mouse R2 RNR protein, catalytic amounts of the iron(III)–Triapine are reduced to the iron(II)–Triapine complex. In the presence of an external reductant (dithiothreitol), stoichiometric amounts of the potently reactive iron(II)–Triapine complex are formed. Formation of the iron(II)–Triapine complex, as the essential part of the reaction outcome, promotes further reactions with molecular oxygen, which give rise to reactive oxygen species (ROS) and thereby damage the RNR enzyme. Triapine affects the diferric center of the mouse R2 protein and, unlike hydroxyurea, is not a potent reductant, not likely to act directly on the tyrosyl radical.  相似文献   

10.
To further the understanding of the biological importance of metal-binding by avian prion proteins, we have investigated the affinity and selectivity of peptides Hx1 [Ac-HNPGYP-nh] and Hx2 [Ac-NPGYPHNPGYPH-nh] with a range of physiological metals via electrospray ionization mass spectrometry and tyrosine fluorescence emission spectroscopy. Both the hexamer Hx1 and the "dimer" peptide Hx2 bind only one equivalent of Cu(II), although only the latter peptide binds copper with significant affinity (Hx1 K(d)=150+/-35 microM; Hx2 K(d)=1.07+/-0.78 microM, pH 7.0 in 3-(N-morpholino)propanesulfonic acid (MOPS) buffer). Both peptides are selective for Cu(II) over divalent Ca, Co, Mg, Mn, Ni, and Zn. Cyclic voltammetry was used to estimate Cu(II/I) solution potentials at pH 6.8, which were very similar for the two peptides (CuHx1 E degrees'=+350 mV, CuHx2 E degrees'=+320 mV vs. normal hydrogen electrode). These results suggest similar binding modes for the two peptides, and relative stabilization of Cu(I) relative to similar His-Gly-rich peptides in the literature.  相似文献   

11.
Co(II), Ni(II), and N-oxalylglycine (NOG) are well-known inhibitors of Fe(II)/alpha-ketoglutarate (alphaKG)-dependent hydroxylases, but few studies describe their kinetics and no spectroscopic investigations have been reported. Using taurine/alphaKG dioxygenase (TauD) as a paradigm for this enzyme family, time-dependent inhibition assays showed that Co(II) and Ni(II) follow slow-binding inhibition kinetics. Whereas Ni(II)-substituted TauD was non-chromophoric, spectroscopic studies of the Co(II)-substituted enzyme revealed a six-coordinate site (protein alone or with alphaKG) that became five-coordinate upon taurine addition. The Co(II) spectrum was not perturbed by a series of anions or oxidants, suggesting the Co(II) is inaccessible and could be used to stabilize the protein. NOG competed weakly (Ki approximately 290 microM) with alphaKG for binding to TauD, with the increased electron density of NOG yielding electronic transitions for NOG-Fe(II)-TauD and taurine-NOG-Fe(II)-TauD at 380 nm (epsilon380 90-105 M(-1) cm(-1)). The spectra of the NOG-bound TauD species did not change significantly upon oxygen exposure, arguing against the formation of an oxygen-bound state mimicking an early intermediate in catalysis.  相似文献   

12.
The bimolecular reduction of the Cu(II)-based enzyme lysyl oxidase with two inorganic reductants, tris bipyridylchromium(II) and (1,3,6,8,10,13,16,19)-octaazabicyclo (6,6,6)eicosanecobalt(II) has been examined at various ionic strength and [H+] conditions. The electrochemical properties of the enzyme have also been examined. The results show that Cu(II) is the redox site in the enzyme and has E 1/2 = 0.05 +/- 0.005 V against SCE. The observed rate constants, kobs, for the reduction of the enzyme by either Cr(bpy)32+ or Co(sep)2+ at any concentration of the reductant increased with the ionic strength of the medium. The ionic strength dependence of kobs has been analyzed in terms of the charge of the active site being 1 +.  相似文献   

13.
The effect of several transition metals on the activity of Desulfovibrio gigas hydrogenase has been studied. Co(II) and Ni(II) at a concentration of 1 mM did not modify the activity of the enzyme; nor did they affect the pattern of activation/deactivation. Cu(II) inhibited the active hydrogenase, prepared by treatment with hydrogen, but had little effect on the 'unready' enzyme unless a reductant such as ascorbate was present, in which case inactivation took place either in air or under argon. Hg(II) also inactivated the enzyme irreversible in the 'unready' state without the requirement for reductants. The reaction of H2 uptake with methyl viologen was much more sensitive to inhibition than the H2/tritium exchange activity. EPR spectra of this preparation showed that the rates of decline were [3Fe-4S] signal greater than H2-uptake activity greater than Ni-A signal. Similar results were obtained when the protein was treated with Hg(II). The results demonstrate that the [3Fe-4S] cluster is not essential for H2-uptake activity with methyl viologen, but the integrity of [4Fe-4S] clusters is probably necessary to catalyze the reduction of methyl viologen with hydrogen. D. gigas hydrogenase was found to be highly resistant to digestion by proteases.  相似文献   

14.
Azotobacter vinelandii hydrogenase (EC class 1.12), either purified or membrane-associated, was obtained aerobically in an inactive state. The kinetics of activation by treatment with a reductant (H2 or dithionite) were determined. Three distinct phases of the activation were observed. Aerobically prepared, inactive hydrogenase was insensitive to acetylene inhibition, but could be rendered acetylene-sensitive by reduction with dithionite. These findings indicate that acetylene inhibition of hydrogenase requires catalytically active enzyme.  相似文献   

15.
The ionic currents of clonal Y-1 adrenocortical cells were studied using the whole-cell variant of the patch-clamp technique. These cells had two major current components: a large outward current carried by K ions, and a small inward Ca current. The Ca current depended on the activity of two populations of Ca channels, slow (SD) and fast (FD) deactivating, that could be separated by their different closing time constants (at -80 mV, SD, 3.8 ms, and FD, 0.13 ms). These two kinds of channels also differed in (a) activation threshold (SD, approximately -50 mV; FD, approximately -20 mV), (b) half-maximal activation (SD, between -15 and -10 mV; FD between +10 and +15 mV), and (c) inactivation time course (SD, fast; FD, slow). The total amplitude of the Ca current and the proportion of SD and FD channels varied from cell to cell. The amplitude of the K current was strongly dependent on the internal [Ca2+] and was almost abolished when internal [Ca2+] was less than 0.001 microM. The K current appeared to be independent, or only slightly dependent, of Ca influx. With an internal [Ca2+] of 0.1 microM, the activation threshold was -20 mV, and at +40 mV the half-time of activation was 9 ms. With 73 mM external K the closing time constant at -70 mV was approximately 3 ms. The outward current was also modulated by internal pH and Mg. At a constant pCa gamma a decrease of pH reduced the current amplitude, whereas the activation kinetics were not much altered. Removal of internal Mg produced a drastic decrease in the amplitude of the Ca-activated K current. It was also found that with internal [Ca2+] over 0.1 microM the K current underwent a time-dependent transformation characterized by a large increase in amplitude and in activation kinetics.  相似文献   

16.
Two isozyme forms of arginase are found in the rat. All arginases are metalloenzymes which require manganese for activity. Many arginases are activated by cobalt and nickel ions and inhibited by heavy metal ions. The purpose of this study was to compare the effect of other heavy metal ions on the rat liver isozyme (arginase I) and the rat kidney isozyme (arginase II). The activation and inhibition of arginase I and II by metal ions were different. However, both isozymes were strongly inhibited by cupric and mercuric ions. The inhibition of arginase I by cupric and mercuric ions was increased greatly by preincubation of the enzyme with the metal ions. However, preincubation of arginase II by cupric and mercuric ions had little effect on the inhibition of the enzyme. Under certain conditions the kinetics of the inhibition of both arginases I and II by cupric and mercuric ions was nonlinear allosteric.  相似文献   

17.
The redox behaviour of the Ni(III)/Ni(II) transition in hydrogenase from Chromatium vinosum is described and compared with the redox behaviour of the nickel ion in the F420-nonreducing hydrogenase from Methanobacterium thermoautotrophicum. Analogous to the situation in the oxidised hydrogenase of Desulfovibrio gigas (Fernandez, V.M., Hatchikian, E.C., Patil, D.S. and Cammack, R. (1986) Biochim. Biophys. Acta 883, 145-154), the C. vinosum enzyme can also exist in two forms: the 'unready' form (EPR characteristics of Ni(III): gx,y,z = 2.32, 2.24, 2.01) and the 'ready' form (EPR characteristics Ni(III): gx,y,z = 2.34, 2.16, 2.01). Like in the oxidised enzyme of M. thermoautotrophicum the Ni(III)/Ni(II) transition for the unready form titrated completely reversible (both at pH 6.0 and pH 8.0). In contrast, the reversibility of the Ni(III)/Ni(II) transition in the ready enzyme was strongly dependent on pH and temperature. At pH 6.0 and 2 degrees C reduction of Ni(III) in ready enzyme was completely irreversible, whereas at pH 8.0 and 30 degrees C Ni(III) in both ready and unready enzyme titrated with E0' = -115 mV (n = 1). Hampered redox equilibration between the ready enzyme and the mediating dyes is interpreted in terms of an obstruction of the electron transfer from nickel at the active site to the artificial electron acceptors in solution. The origin of this obstruction might be related to possible changes in the protein structure induced by the activation process. The E0'-value of the Ni(III)/Ni(II) equilibrium was pH sensitive (-60 mV/delta pH) indicating that reduction of nickel is coupled to a protonation. A similar pH-dependence was observed for the titration of the spin-spin interaction of Ni(III) and a special form of the [3Fe-4S]+ cluster (E0' = +150 mV, pH 8.0, 30 degrees C). Redox equilibration of this coupling was extremely sensitive to pH and temperature. The uncoupled [3Fe-4S]+ cluster titrated pH-independently with E0' = -10 mV (pH 8.0, 30 degrees C).  相似文献   

18.
C E Cooper  P Nicholls 《Biochemistry》1990,29(16):3865-3871
Cytochrome oxidase proteoliposomes were prepared from bovine heart oxidase. Size distributions determined by quasi-elastic light scattering (QELS) showed that there was a small population of large vesicles (120-200-nm diameter) and a large population of small vesicles (50-100-nm diameter). Trapping cytochrome c inside the proteoliposomes did not significantly alter this size distribution. Separation of the vesicles by gel filtration, however, revealed that the cytochrome c/cytochrome a ratio is higher in the larger vesicles. Internally trapped cytochrome c can be reduced by the membrane-permeable reductants 2,3,5,6-tetramethyl-p-phenylenediamine (DAD) or N,N,N',N'-tetramethyl-p-phenylenediamine (TMPD). Respiration on internal cytochrome c generated a membrane potential of 53 mV (positive inside) and a pH gradient of 0.2 (acid inside) as monitored by the optical probes oxonol V and pyranine, respectively. But the true magnitude of these gradients in individual proteoliposomes is complicated by vesicle heterogeneity. The membrane potential increased biphasically with increasing concentration of reductant. Ionophore sensitivity was higher for the "low Km" phase, and respiration became increasingly uncoupled as the reductant concentration was increased. These findings are consistent with a kinetic heterogeneity such that vesicles respiring at lower reductant concentrations generate a higher proton motive force than those with a larger Km. The steady-state internal acidification induced by turnover of the internally facing enzyme is probably maintained by both cytochrome oxidase proton translocation and a TMPD+/H+ antiport present in these vesicles [Cooper, C. E., & Nicholls, P. (1987) FEBS. Lett. 223, 155-160].  相似文献   

19.
A wide range of organic reductants, including many iron chelators, reduce ferryl myoglobin to its ferric states in exponential time courses whose rate constants display double hyperbolic dependencies on the reductant concentration. This concentration dependence is consistent with a mechanism in which electron transfer to the heme takes place at two independent sites where reductants appear to bind. We propose that the low-affinity site is located close to the heme edge, within the heme pocket; the maximum rate of electron transfer is highly variable depending on the nature of the reductant (0.005 to >10 s(-1)). The other site has higher apparent affinity (K(D) 0.2-50 microM) but a low maximum rate of electron transfer (0.005 to 0.01 s(-1)). By examining native and engineered proteins we have determined that the high-affinity pathway represents a through-protein electron transfer pathway that involves a specific tyrosine residue. The low apparent rate constant for electron transfer from the tyrosine to the heme (approximately 5 A) is accounted for by proposing that electron transfer occurs only in a very poorly populated protonated state of ferryl heme and tyrosine. Hemoglobin shows similar kinetics but only one subunit exhibits double rectangular hyperbolic concentration dependency. The consequence of a high-affinity through-protein electron transfer pathway to the cytotoxicity of ferryl heme is discussed.  相似文献   

20.
Redox thermodynamic data provide a detailed insight into control of the reduction potential E degrees' of the [Fe(S-Cys)4] site in rubredoxin. Mutant forms were studied in which specific structural changes were made in both the primary and secondary coordination spheres. Those changes have been probed by resonance Raman spectroscopy. The decrease of approximately 200 mV in E degrees' observed for the [Fe(S-Cys)3(O-Ser)]-/2- couples in the surface ligand mutants C9S and C42S is essentially enthalpic in origin and associated with the substitution of ligand thiolate by ligand olate. However, the pH dependence of the potentials below characteristic pKa(red) approximately equals 7 is an entropic contribution, plausibly associated with increased conformational flexibility induced by a longer Fe(II)-O(H)-Ser bond in the reduced form. The presence of a second surface Ser ligand in the new double mutant protein C9S/C42S affects the enthalpic term primarily for pH>pKa(red) > or = 9.3, but for pHpKa approximately 9: [Fe(III)(S-Cys)3(OH)]- + e- --> [Fe(II)(S-Cys)3(OH)]2-. pH [Fe(II)(S-Cys)3(OH2)]-.  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