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 22 毫秒
1.
J. Butler  G.G. Jayson  A.J. Swallow 《BBA》1975,408(3):215-222

1. 1. The superoxide anion radical (O2) reacts with ferricytochrome c to form ferrocytochrome c. No intermediate complexes are observable. No reaction could be detected between O2 and ferrocytochrome c.

2. 2. At 20 °C the rate constant for the reaction at pH 4.7 to 6.7 is 1.4 · 106 M−1 · s−1 and as the pH increases above 6.7 the rate constant steadily decreases. The dependence on pH is the same for tuna heart and horse heart cytochrome c. No reaction could be demonstrated between O2 and the form of cytochrome c which exists above pH ≈ 9.2. The dependence of the rate constant on pH can be explained if cytochrome c has pKs of 7.45 and 9.2, and O2 reacts with the form present below pH 7.45 with k = 1.4 · 106 M−1 · s−1, the form above pH 7.45 with k = 3.0 · 105 M−1 · s−1, and the form present above pH 9.2 with k = 0.

3. 3. The reaction has an activation energy of 20 kJ mol−1 and an enthalpy of activation at 25 °C of 18 kJ mol−1 both above and below pH 7.45. It is suggested that O2 may reduce cytochrome c through a track composed of aromatic amino acids, and that little protein rearrangement is required for the formation of the activated complex.

4. 4. No reduction of ferricytochrome c by HO2 radicals could be demonstrated at pH 1.2–6.2 but at pH 5.3, HO2 radicals oxidize ferrocytochrome c with a rate constant of about 5 · 105–5 · 106 M−1 · s−1

.  相似文献   


2.
Direct evidence obtained by means of the technique of pulse radiolysis-kinetic spectrometry, with measurements in the time range 10−6 to 1 s, is presented that, consequent upon reaction of a single H-atom with a single molecule of ferricytochrome c, a reducing equivalent is transmitted via the protein structure to the ferriheme moiety. Such transmission accounts for at least 70% of the total reduction of the ferri to the ferro state of cytochrome c. The remainder of the total reduction takes place without stages resolvable on the time scale of these experiments. Reduction brought about by H atoms appears to follow a different course than reduction by hydrated electrons. In the latter case, intramolecular transmission of reducing equivalents could not be demonstrated (Lichtin, N. N., Shafferman, A. and Stein, G. (1973) Biochim. Biophys. Acta 314, 117–135).

Not every H-atom reacts with ferricytochrome c at a site which results in conversion of the Fe(III) state to the Fe(II) state. Approximately half of reacting H-atoms do not produce reduction.

The following second order rate constants have been determined in solutions of low ionic strength at 20±2 °C: k[H+ferricytochrome c] = (1.0±0.2) · 1010 M−1 · s−1 at pH 3.0 and 6.7; k[H+ferrocytochrome c] = (1.3±0.2) · 1010 M−1 · s−1 at pH 3.0; k[eaq + ferrocytochrome c] = (1.9±0.4) · 1010 M−1 · s−1 at pH 6.7.  相似文献   


3.
Richard Maskiewicz  Benon H.J. Bielski   《BBA》1982,680(3):297-303
It has been shown by the pulse radiolysis technique that radiation-generated NADP free radicals (NADP·) first combine with ferredoxin-NADP reductase and then transfer the odd electron by a fast intramolecular process to the enzyme flavin moiety yielding the semiquinone (ferredoxin-NADP reductase, FNR-FADH·). The corresponding first-order rate constant k15 varies with ionic strength from 2.6·103 s−1 at I = 0.66 M to 2.3·104 s−1 at I = 0.005 M In the presence of ferredoxin-NADP reductase-bound oxidized ferredoxin, the electron cascades, thus further reducing the ferredoxin. The transfer of the electron from the flavin semiquinone (ferredoxin-NADP reductase, FNR-FADH·) to the bound oxidized ferredoxin proceeds at a rate of k18 = 2.36 s−1. This process approaches an equilibrium condition which is in favor of the reverse reaction suggesting that k−18 > k18.  相似文献   

4.
P.Muir Wood 《BBA》1974,357(3):370-379
The rate of electron transfer between reduced cytochrome ƒ and plastocyanin (both purified from parsley) has been measured as k = 3.6 · 107 M−1 · s−1, at 298 °K and pH 7.0, with activation parameters ΔH = 44 kJ · mole−1 and ΔS = +46 J · mole−1 · °K−1. Replacement of cytochrome ƒ with red algal cytochrome c-553, Pseudomonas cytochrome c-551 and mammalian cytochrome c gave rates at least 30 times slower: k = 5 · 105, 7.5 · 105 and 1.0 · 106 M−1 · s−1, respectively.

Similar measurements made with azurin instead of plastocyanin gave k = 6 · 106 and approx. 2 · 107 M−1 · s−1 for reaction of reduced azurin with cytochrome ƒ and algal cytochrome respectively.

Rate constants of 115 and 80 M−1 · s−1 were found for reduction of plastocyanin by ascorbate and hydroquinone at 298 °K and pH 7.0. The rate constants for the oxidation of plastocyanin, cytochrome ƒ, Pseudomonas cytochrome c-551 and red algal cytochrome c-553 by ferricyanide were found to be between 3 · 104 and 8 · 104 M−1 · s−1.

The results are discussed in relation to photosynthetic electron transport.  相似文献   


5.
H.F. Kauffman  B.F. Van Gelder 《BBA》1973,314(3):276-283
1. Cyanide causes a slow disappearance of the oxidized band (648 nm) of cytochrome d in particles of Azotobacter vinelandii and inhibits the appearance of the reduced band (631 nm). No effect of cyanide is found on the reduced band of cytochrome d.

2. The kinetics of the disappearance of the 648-nm band of cytochrome d with excess cyanide deviates from first-order kinetics at lower temperatures (22 °C) indicating that at least two conformations of the enzyme are involved. At higher temperatures (32 °C) the observed kinetics of the cyanide reaction are first order with a kon = 0.7 M−1·s−1 and with an estimated koff of approximately 5·10−5 s−1.

3. The value of the koff (7·10−4−14·10−4 s−1 at 32 °C) determined from the rate of reduction of cyanocytochrome d by Na2S2O4 or NADH is one order of magnitude larger than the koff value found when the enzyme is in its oxidized state.

4. No effect of cyanide is found on the spectrum of cytochrome a1.  相似文献   


6.
Roger N.F. Thorneley 《BBA》1974,333(3):487-496
1. Single reduced methyl viologen (MV.+) acts as an electron donor in a number of enzyme systems. The large changes in extinction coefficient upon oxidation (λmax 600 nm; MV.+, = 1.3 · 104 M−1 · cm−1; oxidised form of methyl viologen (MV2+), = 0.0) make it ideally suited to kinetic studies of electron transfer reactions using stopped-flow and standard spectrophotometric techniques.

2. A convenient electrochemical preparation of large amounts of MV.+ has been developed.

3. A commercial stopped-flow apparatus was modified in order to obtain a high degree of anaerobicity.

4. The reaction of MV.+ with O2 produced H2O2 (k > 5 · 106 M−1 · s−1, pH 7.5, 25 °C). H2O2 subsequently reacted with excess MV.+ (k = 2.3 · 103 M−1 · s−1, pH 7.5, 25 °C) to produce water. The kinetics of this reaction were complex and have only been interpreted over a limited range of concentrations.

5. The results support the theory that the herbicidal action of methyl viologen (Paraquat, Gramoxone) is due to H2O2 (or radicals derived from H2O2) induced damage of plant cell membrane.  相似文献   


7.
The kinetics of the reaction of hydrated electron (eaq) and carboxyl anion radical (CO2) with Pseudomonas aeruginosa ferricytochrome c-551 were studied by pulse radiolysis. The rate of reaction of eaq with the negatively charged ferricytochrome c-551 (17 nM−1 · s−1) is significantly slower than the larger positively charged horse heart ferricytochrome c (70 nM · s). This difference cannot be explained solely by electrostatic effects on the diffusion-controlled reactions. After the initial encounter of eaq with the protein, ferricytochrome c-551 is less effective in transferring an electron to the heme which may be due to the negative charge on the protein. The charge on ferricytochrome c-551 is estimated to be −5 at pH 7 from the effect of ionic strength on the reaction rate. A slower relaxation (2 · 104 s−1) observed after fast eaq reduction is attributed to a small conformational change. The rate of reaction of CO2 with ferricytochrome c-551 (0.7 nM−1 · s) is, after electrostatic correction, the same as ferricytochrome c, indicating that the steric requirements for reaction are similar. This reaction probably takes place through the exposed heme edge.  相似文献   

8.
Oxygenation of [CuII(fla)(idpa)]ClO4 (fla=flavonolate; IDPA=3,3′-iminobis(N,N-dimethylpropylamine)) in dimethylformamide gives [CuII(idpa)(O-bs)]ClO4 (O-bs=O-benzoylsalicylate) and CO. The oxygenolysis of [CuII(fla)(idpa)]ClO4 in DMF was followed by electronic spectroscopy and the rate law −d[{CuII(fla)(idpa)}ClO4]/dt=kobs[{CuII(fla)(idpa)}ClO4][O2] was obtained. The rate constant, activation enthalpy and entropy at 373 K are kobs=6.13±0.16×10−3 M−1 s−1, ΔH=64±5 kJ mol−1, ΔS=−120±13 J mol−1 K−1, respectively. The reaction fits a Hammett linear free energy relationship and a higher electron density on copper gives faster oxygenation rates. The complex [CuII(fla)(idpa)]ClO4 has also been found to be a selective catalyst for the oxygenation of flavonol to the corresponding O-benzoylsalicylic acid and CO. The kinetics of the oxygenolysis in DMF was followed by electronic spectroscopy and the following rate law was obtained: −d[flaH]/dt=kobs[{CuII(fla)(idpa)}ClO4][O2]. The rate constant, activation enthalpy and entropy at 403 K are kobs=4.22±0.15×10−2 M−1 s−1, ΔH=71±6 kJ mol−1, ΔS=−97±15 J mol−1 K−1, respectively.  相似文献   

9.
To clarify the radical-scavenging activity of butylated hydroxytoluene (BHT), a food additive, stoichiometric factors (n) and inhibition rate constants (kinh) were determined for 2,6-di-tert-butyl-4-methylphenol (BHT) and its metabolites 2,6-di-tert-butyl-p-benzoquinone (BHT-Q), 3,5-di-tert-butyl-4-hydroxybenzaldehyde (BHA-CHO) and 3,5-di-tert-butyl-4-hydroperoxy-4-methyl-2,5-cyclohexadiene-1-one (BHT-OOH). Values of n and kinh were determined from differential scanning calorimetry (DSC) monitoring of the polymerization of methyl methacrylate (MMA) initiated by 2,2′-azobis(isobutyronitrile) (AIBN) or benzoyl peroxide (BPO) at 70 °C in the presence or absence of antioxidants (BHT-related compounds). The n values declined in the order BHT (1–2) > BHT-CHO, BHT-OOH (0.1–0.3) > BHT-Q (0). The n value for BHT with AIBN was approximately 1.0, suggesting dimerization of BHT. The kinh values declined in the order BHT-Q ((3.5–4.6)×104 M−1 s−1) > BHT-OOH (0.7–1.9×104 M−1 s−1) > BHT-CHO ((0.4–1.7)×104 M−1 s−1) > BHT ((0.1–0.2)×104 M−1 s−1). The kinh for metabolites was greater than that for the parent BHT. Growing MMA radicals initiated by BPO were suppressed much more efficiently by BHT or BHT-Q compared with those initiated by AIBN. BHT was effective as a chain-breaking antioxidant.  相似文献   

10.
1. Rate constants for reduction of paraquat ion (1,1′-dimethyl-4,4′-bipyridy-lium, PQ2+) to paraquat radical (PQ+·) by eaq and CO2· have been measured by pulse radiolysis. Reduction by eaq is diffusion controlled (k = 8.4·1010 M−1·s−1) and reduction by CO2· is also very fast k = 1.5·1010 M−1·s−1).

2. The reaction of paraquat radical with oxygen has been analysed to give rate constants of 7.7·108 M−1·s−1 and 6.5·108 M−1·s−1 for the reactions of paraquat radical with O2 and O2·, respectively. The similarity in these rate constants is in marked contrast to the difference in redox potentials of O2 and O2· (− 0.59 V and + 1.12 V, respectively).

3. These rate constants, together with that for the self-reaction of O2·, have been used to calculate the steady-state concentration of O2· under conditions thought to apply at the site of reduction of paraquat in the plant cell. On the basis of these calculations the decay of O2· appears to be governed almost entirely by its self-reaction, and the concentration 5 μm away from the thylakoid is still 90% of that at the thylakoid itself. Thus, O2· persists long enough to diffuse as far as the chloroplast envelope and tonoplast, which are the first structures to be damaged by paraquat treatment. O2· is therefore sufficiently long-lived to be a candidate for the phytotoxic product formed by paraquat in plants.  相似文献   


11.
1. The reduction of cytochrome c oxidase by hydrated electrons was studied in the absence and presence of cytochrome c.

2. Hydrated electrons do not readily reduce the heme of cytochrome c oxidase. This observation supports our previous conclusion that heme a is not directly exposed to the solvent.

3. In a mixture of cytochrome c and cytochrome c oxidase, cytochrome c is first reduced by hydrated electrons (k = 4 · 1010 M−1 · s−1 at 22 °C and pH 7.2) after which it transfers electrons to cytochrome c oxidase with a rate constant of 6 · 107 M−1 · s−1 at 22 °C and pH 7.2.

4. It was found that two equivalents of cytochrome c are oxidized initially per equivalent of heme a reduced, showing that one electron is accepted by a second electron acceptor, probably one of the copper atoms of cytochrome c oxidase.

5. After the initial reduction, redistribution of electrons takes place until an equilibrium is reached similar to that found in redox experiments of Tiesjema, R. H., Muijsers, A. O. and Van Gelder, B. F. (1973) Biochim. Biophys. Acta 305, 19–28.  相似文献   


12.
We have used picosecond absorption spectroscopy with low intensity (5 · 1011–5 · 1012 photons · pulse−1 · cm−2) continuously tunable infrared (800–900 nm) pulses to study the energy transfer dynamics in the isolated B800–850 pigment-protein complex of Rhodobacter sphaeroides. Our results suggest the following picture of the energy transfer dynamics: (i) a fast transfer, within approx. 1 ps, from BChl 800 to BChl 850; (ii) transfer among different BChl 800's with a rate which is at the most of the same order of magnitude as that of BChl 800 → BChl 850 transfer; (iii) very fast transfer (k > 1 · 1012 s−1) between BChl 850 molecules. Assuming Förster type of energy transfer maximum distances of about 22 and 15 Å are obtained for the BChl 800–BChl 850 and BChl 850–BChl 850 separations, respectively.  相似文献   

13.
The kinetics and equilibria of complex formation by Ga(III) with NCS in aqueous solution have been measured over a range of acidities and temperatures, the contributing paths to the reaction resolved, and their rate constants and activation parameters determined. The hydrolysis equilibria required to carry out this resolution of kinetic behaviour have also been measured.

Unlike the other reported complexation reactions of Ga(III) in aqueous solution, the separate reaction pathways can be assigned with no ambiguity. At 25 °C and ionic strength 0.5 M, the observed forward rate constant for the complex formation is described by {k1 + k2K1h/[H+] + k3K1hK2h/[H+]2} M−1 s−1. For these conditions, the first and second successive hydrolysis constants of Ga(H2O)63+ are given by pK1h = 3.69 ± 0.01 and pK2h = 3.74 ± 0.04. The rate constants corresponding to the reactions of the species Ga(H2O)63+, Ga(H2O)5(OH)2+ and Ga(H2O)4(OH)2+ with NCS are k1 = 57 ± 4 M−1 −1, k2 = (1.08 ± 0.01) × 105 M−1 s−1 and k3 = 3 × 106 M−1 s−1 respectively. The complexation equilibrium quotient [GaNCS2+]/([Ga3+][NCS]) has been independently determined by spectrophotometric titration to be 20.8 ± 0.3 M−1 at 25 °C and ionic strength 0.5 M.

These kinetic results lead to an interpretation of the data, and a reinterpretation of other data for aquo-Ga(III) complex formation kinetics from the literature which support the assignment of a dissociative interchange mechanism for these reactions rather than the associative activation mode sometimes proposed.  相似文献   


14.
The reaction between a cytochrome oxidase from Pseudomonas aeruginosa and oxygen has been studied by a rapid mixing technique. The data indicate that the heme d1 moiety of the ascorbate-reduced enzyme is oxidized faster than the heme c component. The oxidation of heme d1 is accurately second order with respect to oxygen and has a rate constant of 5.7 · 104 M−1 · s−1 at 20 °C. The oxidation of the heme c has a first-order rate constant of about 8 s−1 at infinite concentration of O2. The results indicate that the rate-limiting step is the internal transfer of electrons from heme c to heme d1. These more rapid reactions are followed by more complicated but smaller absorbance changes whose origin is still not clear.

The reaction of ascorbate-reduced oxidase with CO has also been studied and is second order with a rate constant of 1.8 · 104 M−1 · s−1. The initial reaction with CO is followed by a slower reaction of significantly less magnitude. The equilibrium constant for the reaction with CO, calculated as a dissociation constant from titrimetric experiments with dithionite-reduced oxidase, is about 2.3 · 10−6 M. From these data a rate constant of 0.041 s−1 can be calculated for the dissociation of CO from the enzyme.  相似文献   


15.
The reaction of peroxynitrous acid with monohydroascorbate, over the concentration range of 250 μM to 50 mM of monohydroascorbate at pH 5.8 and at 25°C, was reinvestigated and the rate constant of the reaction found to be much higher than reported earlier (Bartlett, D.; Church, D. F.; Bounds, P. L.; Koppenol, W. H. The kinetics of oxidation of L-ascorbic acid by peroxynitrite. Free Radic. Biol. Med. 18:85–92; 1995; Squadrito, G. L.; Jin, X.; Pryor, W. A. Stopped-flow kinetics of the reaction of ascorbic acid with peroxynitrite. Arch. Biochem. Biophys. 322:53–59; 1995). The new rate constants at pH 5.8 are k1 = 1 × 106 M−1 s−1 and k−1 = 500 s−1 for 25°C and k1 = 1.5 × 106 M−1 s−1 and k−1 = 1 × 103 s−1 for 37°C. These values indicate that even at low monohydroascorbate concentrations most of peroxynitrous acid forms an adduct with this antioxidant. The mechanism of the reaction involves formation of an intermediate, which decays to a second intermediate with an absorption maximum at 345 nm. At low monohydroascorbate concentrations, the second intermediate decays to nitrate and monohydroascorbate, while at monohydroascorbate concentrations greater than 4 mM, this second intermediate reacts with a second monohydroascorbate to form nitrite, dehydroascorbate, and monohydroascorbate. EPR experiments indicate that the yield of the ascorbyl radical is 0.24% relative to the initial peroxynitrous acid concentration, and that this small amount of ascorbyl radicals is formed concomitantly with the decrease of the absorption at 345 nm. Thus, the ascorbyl radical is not a primary reaction product. Under the conditions of these experiments, no homolysis of peroxynitrous acid to nitrogen dioxide and hydroxyl radical was observed. Aside from monohydroascorbate's ability to “repair” oxidatively modified biomolecules, it may play a role as scavenger of peroxynitrous acid.  相似文献   

16.
Cuaq+ forms stable complexes with carbon monoxide in aqueous solutions. Furthermore it reacts very fast with aliphatic radicals. The reaction of Cu(CO)maq+ with methyl radicals, CH3 was studied using the pulse-radiolysis technique. The results point out that methyl radicals react with Cu(CO)aq+ to form an unstable intermediate with a CuII-C σ bond identified as (CO)CuII-CH3+, k = (1.1±0.2) × 109 M−1 s−1. This intermediate has a strong LMCT charge transfer band (λmax = 385 nm, max = 2500 M−1 cm−1) which is similar to the absorption bands of other transient complexes with CuII-alkyl σ bonds. The coordinated carbon monoxide in (CO)CuII-CH3+ inserts into the copper—carbon bond (or rather the coordinated methyl migrates to the coordinated carbon monoxide ligand) at a rate of (3.0±0.8) × 102 s−1 to form the copperacetyl complex (CO)mCuII-C(CH3)=O+max = 480 nm, max = 2100 M−1 cm−1). The rate of formation of (CO)CuII-CH3+ and of the insertion reaction are pH independent. The complex (CO)mCuII-C(CH3)=O+ is also unstable and decomposes heterolytically to yield acetaldehyde and Cuaq2+ as the final stable products. This reaction is slightly pH dependent. The same reactivity pattern has been observed for the Cu(COnaq+ complexes (n = 2 or 3). The results clearly point out that CO remains coordinated to transient complexes of the type CuII-alkyl.  相似文献   

17.
1. Fluoride is a mixed-type inhibitor of the cytochrome c oxidase activity with a Ki for the free enzyme of 10 mM and a Ki for the cytochrome c-complexed enzyme of 35 mM.

2. Fluoride shifts the γ-band of the enzyme from 423 to 421 nm and the -band from 597 to 598 nm. The difference spectrum (oxidized enzyme in the presence of fluoride minus oxidized enzyme) has peaks at 400, 453, 482, 605 and 638 nm and troughs at 430, 520, 552 and 674 nm. The changes in absorbance are small (about 3% at absorbance maxima) with respect to those of other hemoproteins.

3. On addition of fluoride to isolated cytochrome c oxidase 3 reactions can be distinguished: (I) a bimolecular binding reaction (Kon = 4 M−1 · s−1 and koff = 2.9 · 10−2s−1 at 25 °C, pH 7.4) contributing at 638 nm and 430 nm; (II) a first-order reaction (k = 2.4 · 10−2) s−1 at 22 °C, pH 7.2) visible mainly at 430 nm and (III) a very slow reaction with a half-time in the order of 10 min.

4. The spectroscopic dissociation constants for the fluoride binding, determined from Hill plots using the absorbance changes at 638 and 430 nm, are similar (7 and 10 mM, respectively, at 22 °C, pH 7.2).

5. A mechanism for the reaction is discussed in which the bimolecular binding reaction is followed by a conformational change of the enzyme-fluoride complex.  相似文献   


18.
Proton NMR studies of N,N-diethylformamide (def) exchange on [M(Me6tren)def]2+ where M = Co and Cu yield: kex (298.2K) = 26.3 ± 2.2, 980 ± 70 s−1; ΔH = 58.3 ± 1.7, 36.3 ± 0.9 kJ mol−1; ΔS= −22.2 ± 4.6, −65.9 ± 2.5 J K−1 mol−1; and ΔV = −1.3 ± 0.2, 5.3 ± 0.3 cm3 mol−1 respectively. These data which are consistent with a and d activation modes operating when M = Co and Cu respectively are compared with data for related systems.  相似文献   

19.
Thor Arnason  John Sinclair 《BBA》1976,430(3):517-523
The modulated oxygen polarograph has been used to study the rate-determining steps of photosynthetic oxygen evolution in spinach chloroplasts. The rate constant, k, of the reaction has a value of 218±10 (S.E.) s−1 at 23 °C and an activation energy of 7±2 (S.E.) kcal · mol−1. A kinetic isotope experiment indicated that this step is probably not the water-splitting reaction. These findings resemble previous results with the unicellular alga Chlorella (Sinclair, J. and Arnason, T. (1974) Biochim. Biophys. Acta 368, 393–400). In other experiments we changed the pH, O2 concentration and osmolarity of the medium, and treated the chloroplasts with 1 mM NH4Cl without detecting any significant change in k. These results suggest that the step is irreversible. However, a significantly lower value of k, 110±20 (S.E.) s−1 was obtained when all salts except 1 mM MgCl2 were removed from the medium bathing the chloroplasts.  相似文献   

20.
Electron self-exchange in solutions of the ‘blue’ copper protein plastocyanin is catalysed by the redox-inert multivalent cations Mg2+ or Co(NH3)3+6. Measurements of specific 1H-NMR line broadening with 50% reduced solutions in the presence of these cations show that electron exchange proceeds through encounters of cation-protein complexes which dissociate at high ionic strength. In the presence of 8mM (5 equivalents/total protein) Co(NH3)3+6, with 10 mM cacodylate (pH*6.0) as background electrolyte, the bimolecular rate constant at 25°C is 7 × 104 M−1·s−1. For comparison, the ‘electrostatically screened’ rate constant measured in 0.1 M KCl in the absence of added multivalent cations is ˜ 4 × 103 M1·s−1.

Plastocyanin Electron self-exchange NMR Protein-protein interaction Multivalent cation Blue copper protein  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