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.
Photo-excitation of membrane-bound Rhodobacter sphaeroides reaction centres containing the mutation Ala M260 to Trp (AM260W) resulted in the accumulation of a radical pair state involving the photo-oxidised primary electron donor (P). This state had a lifetime of hundreds of milliseconds and its formation was inhibited by stigmatellin. The absence of the QA ubiquinone in the AM260W reaction centre suggests that this long-lived radical pair state is P+QB, although the exact reduction/protonation state of the QB quinone remains to be confirmed. The blockage of active branch (A-branch) electron transfer by the AM260W mutation implies that this P+QB state is formed by electron transfer along the so-called inactive branch (B-branch) of reaction centre cofactors. We discuss how further mutations may affect the yield of the P+QB state, including a double alanine mutation (EL212A/DL213A) that probably has a direct effect on the efficiency of the low yield electron transfer step from the anion of the B-branch bacteriopheophytin (HB) to the QB ubiquinone.  相似文献   

2.
The selective picosecond excitation of Rhodopseudomonas sphaeroides (R-26) reaction centers (RCs) at 870 nm induces the formation of the transient state within <1 ps followed by the conversion into the state PF (P± Bph±− during 7 ± 2 ps at both 293 K and 110 K. The transient state including the intense bleaching at 800 nm has been shown not to be due: (a) to photon excitation at 870 nm; (b) the excitation of P+; (c) photoselection effects. The transient state is interpreted as the state 1[P+B] in agreement with earlier works. The primary formation of the state 1P+B] and the big effective singlet-triplet splitting in this state correspond to the spectral splitting of the P band at 900 nm in R-26 RCs and at 1000 nm in Rhodopseudomonas viridis RCs found at 4.2 K and attributed to the optical transition to both 1P and 1[P+B] states.  相似文献   

3.
Delayed fluorescence from isolated reaction centers of Rhodopseudomonas sphaeroides was measured to study the energetics of electron transfer from the bacteriochlorophyll complex (P-870, or P) to the primary and secondary quinones (QA and QB). The analysis was based on the assumption that electron transfer between P and Q reaches equilibrium quickly after flash excitation, and stays in equilibrium during the lifetime of the P+Q radical pair. Delayed fluorescence of 1Q reaction centers (reaction centers that contain only QA) has a lifetime of about 0.1 s, which corresponds to the decay of P+QA. 2Q reaction centers (which contain both QA and QB) have a much weaker delayed fluorescence, with a lifetime that corresponds to that of P+QB (about 1 s). In the presence of o-phenanthroline, the delayed fluorescence of 2Q reaction centers becomes similar in intensity and decay kinetics to that of 1Q reaction centers. From comparisons of the intensities of the delayed fluorescence from P+QA and P+QB, the standard free energy difference between P+QA and P+QB is calculated to be 78 ± 8 meV. From a comparison of the intensity of the delayed fluorescence with that of prompt fluorescence, we calculate that P+QA is 0.86 ± 0.02 eV below the excited singlet state of P in free energy, or about 0.52 eV above the ground state PQA. The temperature dependence of the delayed fluorescence indicates that P+QA is about 0.75 eV below the excited singlet state in enthalpy, or about 0.63 eV above the ground state.  相似文献   

4.
Magnetic field-dependent recombination measurements together with magnetic field-dependent triplet lifetimes (Chidsey, E.D., Takiff, L., Goldstein, R.A. and Boxer, S.G. (1985) Proc. Natl. Acad. Sci USA 82, 6850–6854) yield a free energy change ΔG(P+H3P*) = 0.165 eV ±0.008 at 290 K. This does not depend on whether nuclear spin relaxation in the state 3P* is assumed to be fast or slow compared to the lifetime of this state. This value, being (almost) temperature independent, indicates ΔG(P+H3P*) ΔH(P+H3P*) and is consistent with ΔG(1P* − P+H) and ΔH(1P* − 3P*) from previous delayed fluorescence and phosphorescence data, implying ΔG ΔH for all combinations of these states.  相似文献   

5.
Resonance Raman measurements have been performed with solutions of iodine-complexed synthetic amyloses (DP 25–200), malto-oligomers (DP 3–18, and -cylodextrin. Interest was focused on the minimum chain length for helical complex formation and a possible preferred length for the polyiodine chain. Four fundamental vibrations are observed at 164, 112, 52 and 24 cm−1. The 112 cm−1 Raman line was shown to arise both from free I3 (enhanced at 363.8 nm excitation) and from bound iodine (relatively most intense at 457.9 nm excitation). The main signal of complexed iodine at 164 cm−1is enhanced at an excitation wavelength close to the long wavelength absorption maximum. This signal is observed firt with malto-octaose and -cyclodextrin. The less intense signals at 52 and 24−1 are only detected at DP 15 and higher. Raman spectra give no evidence for a preferred length of the polyiodine chain. Significantly identical Raman spectra are obtained when using different molar ratios of I2/KI solution or I2 solution initially free of I ions. The results are discussed in view of previous assignments of the Raman lines to I2, I3/I2, and I5 subunits. Our findings are incompatible with I3 units as the only bound species. They are compatible with both I3/I2 and I3 subunits under certain conditions. In the case of I2 solution used for complexation we favour the polyiodine chain model proposed previously by Cramer35,36. The I3 ions formed could function mainly as chain initiators, as has been suggested by Cesàro and Brant30.  相似文献   

6.
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   


7.
Nitrosylation of Os(H)3ClL2 (L = P1Pr3) affords the known Os(H)2Cl(NO)L2 (2). Soft electrophiles (Ag, Na) react with complex 2 by chloride abstraction to ultimately yield truly 16-electron dihydride Os(H)2(NO)(P1Pr3)2 (4a), characterized by variable-temperature NMR. Complex 4a reversibly binds H2, forming Os(H)2(H2)(NO)(P1Pr3) with an unusually high barrier for intramolecular hydride exchange. Under kinetic conditions, protonation of 2 with strong acids follows the selectivity for chloride abstraction. Thermodynamically, protonation at the hydride is preferred, quantitatively producing cationic OsHCl(NO)L2+, isolated and characterized by X-ray diffraction as the BAr4F− salt (7) (ArF=3,5−(CF3)2C6H3). Structures of isoelectronic OsHCl(NO)(PH3)2 and OsHCl(CO)(PH3)2 were optimized with ab initio DFT (Becke3LYP) methods and compared to show the greater unsaturation of the metal in the cationic species. Both complexes, 4a and 7, are highly electrophilic and reversibly coordinate dichloromethane in solution. The observed reactivity patterns of the synthesized unsaturated hydrides are rationalized in terms of the determining influence of the ‘push-pull’ π-stabilization of the metal center.  相似文献   

8.
A novel nutrient removal/waste heat utilization process was simulated using semicontinuous cultures of the thermophilic cyanobacterium Fischerella. Dissolved inorganic carbon (DIC)-enriched cultures, maintained with 10 mg l−1 daily productivity, diurnally varying temperature (from 55°C to 26–28°C), a 12:12 light cycle (200 μE sec−1 m−2) and 50% biomass recycling into heated effluent at the beginning of each light period, removed > 95% of NO3 + NO2−N, 71% of NH3-N, 82% of PO43− −P, and 70% of total P from effluent water samples containing approximately 400 μg l−1 combined N and 60 μg l−1 P. Nutrient removal was not severely impaired by an altered temperature gradient, doubled light intensity, or DIC limitation. Recycling 75% of the biomass at the end of each light period resulted in unimpaired NO3 + NO2 removal, 38–45% P removal and no net NH3 removal. Diurnally varying P removal, averaging 50–60%, and nearly constant > 80% N removal, are therefore projected for a full-scale process with continuous biomass recycling.  相似文献   

9.
We have previously shown that crystals of calcium oxalate (COM) elicit a superoxide (O2) response from mitochondria. We have now investigated: (i) if other microparticles can elicit the same response, (ii) if processing of crystals is involved, and (iii) at what level of mitochondrial function oxalate acts. O2 was measured in digitonin-permeabilized MDCK cells by lucigenin (10 μM) chemiluminescence. [14C]-COM dissociation was examined with or without EDTA and employing alternative chelators. Whereas mitochondrial O2 in COM-treated cells was three- to fourfold enhanced compared to controls, other particulates (uric acid, zymosan, and latex beads) either did not increase O2 or were much less effective (hydroxyapatite +50%, p < 0.01), with all at 28 μg/cm2. Free oxalate (750 μM), at the level released from COM with EDTA (1 mM), increased O2 (+50%, p < 0.01). Omitting EDTA abrogated this signal, which was restored completely by EGTA and partially by ascorbate, but not by desferrioxamine or citrate. Omission of phosphate abrogated O2, implicating phosphate-dependent mitochondrial dicarboxylate transport. COM caused a time-related increase in the mitochondrial membrane potential (Δψm) measured using TMRM fluorescence and confocal microscopy. Application of COM to Fura 2-loaded cells induced rapid, large-amplitude cytosolic Ca2+ transients, which were inhibited by thapsigargin, indicating that COM induces release of Ca2+ from internal stores. Thus, COM-induced mitochondrial O2 requires the release of free oxalate and contributes to a synergistic response. Intracellular dissociation of COM and the mitochondrial dicarboxylate transporter are important in O2 production, which is probably regulated by Δψm.  相似文献   

10.
A new functional macrocyclic ligand, 2,4-dinitrophenylcyclen (= 1-(2,4-dinitrophenyl)-1,4,7,10-tetraazacyclododecane), has been synthesized and isolated as its trihydrochloric acid salt (L·3HCl). The protonation constants (log Kn) for three secondary nitrogens of L were determined by potentiometric pH titration to be 10.10, 7.33 and <2 with I = 0.10 (NaNO3) at 25°C. The 2,4-dinitrophenylaniline chromophore was proven to be a good reporter signaling proton- and metal-binding events in the macrocyclic cavity. The UV absorption band (λmax 370 nm, 8200) of the 2,4-dinitrophenylaniline moiety at pH ≥ 9 becomes quenched as pH is lowered (to pH 3.1, where the major species is L·2H+), due to the strong protonation effect extended to the aniline moiety within the macrocyclic cavity. This is in sharp contrast to the pH-independent UV absorption (λmax 390 nm, 14 000) of a reference compound, N,N-diethyl-2,4-dinitroaniline. The UV absorption band of L is shifted to lower wavelengths with Zn2+max 320 nm), Cd2+max 316 nm) and Pb2+max 317 nm), while it almost disappears with Cu2+ and Ni2+. The 1:1 Zn2+ and Cu2+ complexes with L were isolated and characterized. The Zn2+ complex recognizes 1-methylthymine anion (MT) in aqueous solution at physiological pH to yield a stable ternary complex ZnL-MT. The X-ray crystal structure of ZnL-MT showed that Zn2+ is four-coordinate with three secondary nitrogens of L and the deprotonated imide anion that is cofacial to the 2,4-dinitrophenyl ring.  相似文献   

11.
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   

12.
Philip John  F. R. Whatley 《BBA》1970,216(2):342-352
A procedure is described for preparing particles from cells of Micrococcus denitrificans which were broken osmotically after treatment with lysozyme.

1. 1. The preparations catalysed ATP synthesis coupled to O2 uptake or NO3 reduction. With NADH or succinate as the electron donors the P:O ratios were about 1.5 and 0.5, respectively; and the P:NO3 ratios were about 0.9 and 0.06, respectively.

2. 2. Addition of ADP or Pi to the reaction mixture increased the rates of NADH-dependent O2 uptake and NO3 reduction. Addition of 1 mM 2,4-dinitrophenol, which inhibited phosphorylation by 50–60%, increased the basal rates of electron transport.

3. 3. Evidence derived from spectrophotometry and from the differential inhibition by antimycin A of O2 and NO3 reduction leads to the conclusion that the nitrate reductase interacted with the respiratory chain in the region of the b-type cytochrome, and that the c-type cytochrome present was not involved in the reduction of NO3 to NO2.

Abbreviations: TMPD; tetramethyl-p-phenylenediamine  相似文献   


13.
Hiroshi Seki  Yael A. Ilan  Yigal Ilan  Gabriel Stein   《BBA》1976,440(3):573-586
The reduction of ferricytochrome c by O2 and CO2 was studied in the pH range 6.6–9.2 and Arrhenius as well as Eyring parameters were derived from the rate constants and their temperature dependence. Ionic effects on the rate indicate that the redox process proceeds through a multiply-positively charged interaction site on cytochrome c. It is shown that the reaction with O2 and correspondingly with O2 of ferrocytochrome c) is by a factor of approx. 103 slower than warranted by factors such as redox potential. Evidence is adduced to support the view that this slowness is connected with the role of water in the interaction between O2/O2 and ferri-ferrocytochrome c in the positively charged interaction site on cytochrome c in which water molecules are specifically involved in maintaining the local structure of cytochrome c and participate in the process of electron equivalent transfer.  相似文献   

14.
Due to contradictions in the literature we have redetermined the acid-base properties of riboflavin (=RiFl; vitamin B2), i.e. 7,8-dimethyl-10-ribityl-isoalloxazine, and of flavin mononucleotide (FMN2−), also known as riboflavin 5′-phosphate, via potentiometric pH titrations (I = 0.1 M, NaNO3; 25 °C). In contrast to various claims, the isoalloxazine ring cannot be protonated at pH > 1, a result in agreement with an early study (pKa = −0.2; L. Michaelis, M.P. Schubert and C.V. Smythe, J. Biol. Chem., 116 (1936) 587–607); deprotonation of the ring system occurs in both compounds with pKa 10. The pKa value of 0.7 determined for the deprotonation of H2(FMN) must be attributed to the release of the first proton from the fully protonated phosphate group; its second proton is released with pKa = 6.18 in agreement with the acidity constants of various other monoprotonated monophosphate esters. The stability constants of the 1:1 complexes formed between Mg2+, Ca2+, Sr2+, Ba2+, Mn2+, Co2+, Ni2+, Cu2+, Zn2+ or Cd2+ (---M2+) and FMN2− were determined by potentiometric pH titrations in aqueous solution (I = 0.1 M, NaNO3; 25 °C). The log stability constants of all these M(FMN) complexes are about 0.2 log units higher than expected from the basicity of the phosphate group. This slight stability increase cannot be attributed to the formation of a seven-membered chelate involving the ribit-hydroxy group at C-4′ as the stability constants for the M2+ 1:1 complexes of glycerol 1-phosphate (G1P2−) demonstrate: G1P2− contains the same structural unit which would also allow in this case the formation of the mentioned seven-membered chelate; however, the stability of the M(G1P) complexes is solely determined by the basicity of the phosphate group. Hence, in agreement with earlier conclusions (J. Bidwell, J. Thomas and J. Stuehr, J. Am. Chem. Soc., 108 (1986) 820–825) regarding Ni(FMN) one must conclude that the slight stability increase of the M(FMN) complexes has to be attributed to the isoalloxazine ring. The equality of the stability increase of the complexes for all the mentioned ten metal ions precludes its attribution to an interaction with an N site and makes a specific interaction with an O site also somewhat unlikely. In addition, carbonyl oxygens appear as not very favorable for the formation of macrochelates by a further interaction with already phosphate-coordinated metal ions. Therefore, we propose that the slight but significant stability increase originates from M(FMN) species (with a formation degree of about 30%) in which the hydrophobic flavin residue is close to the metal ion, thereby lowering the ‘effective’ dielectric constant in the microenvironment of the metal ion and thus indirectly promoting the −PO32−/M2+ interaction.  相似文献   

15.
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   


16.
Rates of stepwise anation of cis-Cr(ox)2(H2O2) with SCN/N3, Cr(acac)2(H2O)2+ with SCN and Cr(atda)(H2O)2 with SCN have been investigated in weakly acidic aqueous solutions. Rate constants, kI and kII for the two steps in each system, are composite as kx = kx0+kxX[X] (x = I, II; X = SCN, N3). These rate constants have been evaluated also as the corresponding ΔH and ΔS values. The results obtained and the plausible Id mechanism seem to suggest Cr---OOC bond dissociation (hence a strongly negative ΔS) generating the transition state in each system with outer-sphere association forming the precursor complex in the X dependent paths.  相似文献   

17.
Y. Lam  D. J. D. Nicholas 《BBA》1969,180(3):459-472
The formation of nitrite reductase and cytochrome c in Micrococcus denitrificans was repressed by O2. The purified nitrite reductase utilized reduced forms of cytochrome c, phenazine methosulphate, benzyl viologen and methyl viologen, respectively, as electron donors. The enzyme was inhibited by KCN, NaN3 and NH2OH each at 1 mM, whereas CO and bathocuproin, diethyl dithiocarbamate, o-phenanthroline and ,'-dipyridyl at 1 mM concentrations were relatively ineffective. The purified enzyme contains cytochromes, probably of the c and a2 types, in one complex. A Km of 46 μM for NO2 and a pH optimum of 6.7 were recorded for the enzyme. The molecular weight of the enzyme was estimated to be around 130000, and its anodic mobility was 6.8·10−6 cm2·sec−1·V−1 at pH 4.55.

The most highly purified nitrite reductase still exhibited cytochrome c oxidase activity with a Km of 27 μM for O2. This activity was also inhibited by KCN, NaN3 and NH2OH and by NO2.

A constitutive cytochrome oxidase associated with membranes was also isolated from cells grown anaerobically with NO2. It was inhibited by smaller amounts of KCN, NaN3 and NH2OH than the cytochrome oxidase activity of the nitrite reductase enzyme and also differed in having a pH optimum of about 8 and a Km for O2 of less than 0.1 μM. Spectroscopically, cytochromes b and c were found to be associated with the constitutive oxidase in the particulate preparation. Its activity was also inhibited by NO2.

The physiological role of the cytochrome oxidase activity associated with the purified nitrite reductase is likely to be of secondary importance for the following reasons: (a) it accounts for less than 10% of total cytochrome c oxidase activity of cell extracts; (b) the constitutive cytochrome c oxidase has a smaller Km for O2 and would therefore be expected to function more efficiently especially at low concentrations of O2.  相似文献   


18.
1. Difference spectra, at room and liquid N2 temperatures, of S2O42−-, and NO2-reduced intact cells and cell-free preparations of Nitrobacter agilis demonstrated the presence of cytochromes of the c- and a-types. Reduction of cytochromes by succinate, and to a limited extent, by NADPH also occurred, provided KCN (0.1 mM) was also present.

2. A particulate, heat-labile nitrite oxidase having an absolute requirement for O2 was prepared from N. agilis cells using sonic oscillation and differential centrifugation. The particles also possessed NADH oxidase, succinoxidase, formate oxidase and traces of NADPH oxidase activity. The stoichiometry of the nitrite oxidase reaction approached the theoretical value of 2 moles of NO2 consumed per mole of O2 consumed. The pH optimum of the nitrite oxidase system shifted to progressively more alkaline values as the NO2 concentration was increased, changing from a pH value of 6.8 at 0.6 mM KNO2 to pH 8.0 at 0.01 M KNO2 with apparent Km's of 0.2 and 1.2 mM NO2, respectively. Computations of the HNO2 concentrations present under the above conditions showed an approx. 500-fold greater affinity for HNO2 which was independent of pH, suggesting the involvement of HNO2 as both a substrate and an inhibitor (at higher concentrations) of the nitrite oxidase system. The marked inhibition by NaN3, NaCN and Na2S, as well the light-reversible inhibition by CO, indicated the presence of cytochrome oxidase which was subsequently characterized. NO2 proved to be a competitive inhibitor of the nitrite oxidase system.

3. The particulate preparation also possessed a heat-labile nitrite-cytochrome c reductase activity which was energy independent and routinely measured under anaerobic conditions. As in the case of nitrite oxidase, the affinity of the enzyme for NO3 increased as the pH was lowered, but the pH optimum remained unaffected. In terms of calculated HNO2 concentration an approximately constant Km of about 0.2 μM was estimated at the several pH's examined. The inhibition by NO3 was shown to be competitive. The marked sensitivity of the reductase to several metal-binding agents implicated a metal component in the electron transport chain at the site prior to cytochrome c.

4. The membrane-like composition of the nitrite oxidase system is indicated.  相似文献   


19.
The reactions of Ru(NH3)5py2+, Ru(NH3)4bpy2+, Ru2(NH3)10pz5+, RuRh(NH3)10pz5+ and Ru(NH3)5pz2+ with bromine are first-order in ruthenium and first-order in bromine. The rates decrease with increasing bromide ion concentration and, except for Ru(NH3)5pz2+, are independent of hydrogen ion concentration. The reactions are postulated to proceed via outer-sphere, one-electron transfer from Ru(II) to Br2 with the formation of Br2 as a reactive intermediate. The bromide inhibition is ascribed to the formation of Br3 which is unreactive in outer-sphere reactions because of the barrier imposed by the need to undergo reductive cleavage. The reaction of Ru(NH3)5pz2+ is inhibited by hydrogen ions. The hydrogen ion dependence shows that Ru(NH3)5pzH3+ has a pKa of 2.49 and is at least 500 times less reactive than Ru(NH3)5pz2+. The reaction of Ru2(NH3)10pz4+ with bromine is biphasic. The second phase has a rate identical to that of the Ru2(NH3)10pz5+-Br2 reaction. A detailed analysis shows that the reaction of Ru2(NH3)10pz4+ with bromine proceeds by a sequence of one-electron steps, Br2 being produced as an intermediate. A linear free energy relationship between rate constants and equilibrium constants, obeyed for all the reactions studied, provides an estimate of 1.5 × 102 M−1 s−1 for the self-exchange rate constant of the Br2/Br2 couple.  相似文献   

20.
Om wild-type Escherichia coli, near-ultraviolet radiation (NUV) was only weakly mutagenic. However, in an allelic mutant strain (sodA sodB) that lacks both Mn- and Fe-superoxide dismutase (SOD) and assumed to have excess superoxide anion (O2), NUV induced a 9-fold increase in mutation above the level that normally occurs in this double mutant. When a sodA sodB double mutant contained a plasmid carrying katG+ HP-I catalase), mutation by NUV was reduced to wild-type (sodA+ sodB+) levels. Also, in the sodA sodB xthA triple mutant, which lacks exonuclease III (exoIII) in addition to SOD, the mutations frequency by NUV was reduced to wild-type levels. This synergistic action of NUV and O2 suggested that pre-mutational lesions occur, with exoIII converting these lesions to stable mutants. Exposure to H2O2 induced a 2.8 fold increase in mutations in sodA sodB double mutants, but was reduced to control levels when a plasmid carrying katG+ was introduced. These results suggest that NUV, in addition to its other effects on cells, increases mutations indirectly by increasing the flux of OH. radicals, possibly by generating excess H2O2.  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