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 31 毫秒
1.
Near-ultraviolet irradiation of actively growing yeast cells leads to cell death by two distinct mechanisms. The first type of cell death is evident after low doses of near-ultraviolet light (3 times 10-4 ergs times mm- minus 2) and is due to a reversible inactivation of the respiratory capacity of the cell. In studies with yeast mitochondrial membranes the quinones were identified as the site of inactivation by determining the relative levels of the following oxidase activities after irradiation: exogenous NADH, endogenous NADH (via isocitrate dehydrogenase), succinate, and D-lactate oxidases. A second type of cell death is caused after high doses (1.8 times 10-5 ergs times mm- minus 2) and is irreversible. The mechanism of this inactivation is unknown.  相似文献   

2.
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   


3.
The electron irradiation effect on scleroglucans was investigated using different energy doses. Electron spin resonance spectra revealed radicals that were stable for several days. 1H NMR, 13C NMR and Raman spectra indicated no differences in chemical backbone structure due to irradiation. In contrast, lower viscosities of aqueous solutions were received at higher energy doses. This was caused by polymer degradation. The irradiation also decreased the weight average molar masses observed by gel permeation chromatography and multi-angle light scattering. Beginning from raw materials exceeding 4 · 10Da, a number of main chain scissions of approximately 0.3 · 10−7 mol J−1 was found. But for one scleroglucan quality the scission number decreased with higher doses. In addition, the characterization via asymmetrical flow field-flow fractionation proved the presence of low and high molar mass fractions. The electron irradiation led to a preferred scission of the high molar mass chains and increased the lower molar mass fraction. Due to this effect, the broadness of the molar mass distribution decreased.  相似文献   

4.
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   


5.
We measured the toxicity and mutagenicity induced in human diploid lymphoblasts by various radiation doses of X-rays and two internal emitters. [125I]iododeoxyuridine ([125I]dUrd) and [3H]thymidine ([3H]TdR), incorporated into cellular DNA. [125I]dUrd was more effective than [3H]TdR at killing cells and producing mutations to 6-thioguanine resistance (6TGR). No ouabain-resistant mutants were induced by any of these agents. Expressing dose as total disintegrations per cell (dpc), the D0 for cell killing for [125I]dUrd was 28 dpc and for [3H]TdR was 385 dpc. The D0 for X-rays was 48 rad at 37°C. The slopes of the mutation curves were approximately 75 × 10−8 6TGR mutants per cell per disintegration for [125I]dUrd and 2 × 10−8 for [3H]TdR. X-Rays induced 8 × 10−8 6TGR mutants per cell per rad. Normalizing for survival, [125I]dUrd remained much more mutagenic at low doses (high survival levels) than the other two agents. Treatment of the cells at either 37°C or while frozen at −70°C yielded no difference in cytotoxicity or mutation for [125I]dUrd or [3H]TdR, whereas X-rays were 6 times less effective in killing cells at −70°C.

Assuming that incorporation was random throughout the genome, the mutagenic efficiencies of the radionuclides could be calculated by dividing the mutation rate by the level of incorporation. If the effective target size of the 6TGR locus is 1000–3000 base pairs, then the mutagenic efficiency of [125I]dUrd is 1.0–3.0 and of [3H]TdR is 0.02–0.06 total genomic mutations per cell per disintegration. 125I disintegrations are known to produce localized DNA double-strand breaks. If these breaks are potentially lethal lesions, they must be repaired, since the mean lethal dose (D0) was 28 dpc. The observations that a single dpc has a high probability of producing a mutation (mutagenic efficiency 1.0–3.0) would suggest, however, that this repair is extremely error-prone. If the breaks need not be repaired to permit survival, then lethal lesions are a subset of or are completely different from mutagenic lesions.  相似文献   


6.
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   


7.
Comparative measurements of bacterial total counts and volumes of flow cytometry (FCM), transmission electron (TEM), and epifluorescence microscopy (EFM), were undertaken during a four week mesocosm experiment. Total counts of bacteria measured by TEM, EFM, and FCM were in the range of 1 · 106−6 cells ml−1, 1 · 106−3 · 1016 cells ml−1, and 5 · 105 cells ml−1 respectively. The mean volume of the bacterial community, measured by means of EFM and TEM, increased from 0.12–0.15 μm3 at the start of the experiment to 0.39–0.53 μm3 at the end. Generally, there was good agreement between the two methods and regression analyses gave r = 0.87 (p < < 0.01) for cell volume and r = 0.97 (p < < 0.01) for cell number. DAPI stained bacteria with volumes less than 0.2 μm3 were not detected by flow cytometry and these were generally an order of magnitude lower than counts made by TEM and EFM. For samples where the mean bacterial cell volume was longer than 0.3 μm3, all three methods were in agreement both with respect to counts and volume estimates.  相似文献   

8.
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   

9.
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   


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.
The infrared, visible and nuclear magnetic resonance spectra of protochlorophyll a and vinylprotochlorophyll a in dry non-polar solvents (carbon tetrachloride, chloroform, cyclohexane) are presented and interpreted in terms of dimer interaction.

The infrared spectra in the 1600–1800 cm−1 region clearly show the existence of a coordination interaction between the C-9 ketone oxygen function of one molecule and the central magnesium atom of another molecule. Infrared spectra in the OH stretching region (3200–3800 cm−1) provide a valuable test of the water content in the samples.

The analysis of the absorption and circular dichroism spectra of protochlorophyll a and vinylprotochlorophyll a in carbon tetrachloride demonstrates the existence of a monomer-dimer equilibrium in the concentration range from 10−6 to 5 · 10−4 M. The dimerization constants are (6±2) · 105 1 · M−1 for protochlorophyll a and (4.5±2) · 105 1 · M−1 for vinylprotochlorophyll a at 20 °C. The deconvolution of visible spectra in the red region has been performed in order to obtain quantitative information on the dimer structure. Two models involving a parallel or a perpendicular arrangement of the associated molecules are considered.

From 1H NMR spectra, it appears that the region of overlap occurs near ring V, in agreement with the interpretation of the infrared spectra.  相似文献   


12.
In this paper a number of experiments with the purple bacteria Rhodospirillum rubrum and Rhodopseudomonas capsulata is described in which the total fluorescence yield and/or the total fraction of reaction centers closed after a picosecond laser pulse were measured as a function of the pulse intensity. The conditions were such that the reaction centers were either all in the open or all in the closed state before the pulse arrived. These experiments are analysed using the theoretical formalism discussed in the preceding paper (Den Hollander, W.T.F., Bakker J.G.C., and Van Grondelle, R., Biochim. Biophys. Acta 725, 492–507). From the experimental results the number of connected photosynthetic units, λ, the rate of energy transfer between neighboring antenna molecules, kh, and the rate of trapping by an open reaction center, kot, can be estimated. For R. rubrum it is found that λ = 14−17, kh = (1−2)·1012 s−1 and kot = (4−6)·1011 s−1, for Rps. capsulata λ ≈ 30, kh ≈ 4·1011 s−1 and kot ≈ 3·1011 s−1. The findings are discussed in terms of current models for the structure of the antenna and the kinetic properties of the decay processes occurring in these purple bacteria.  相似文献   

13.
The reduction of spinach ferredoxin by V(II)-EDTA, Eu(II)-DPTA and propylene viologen was monitored electrochemically. The rates of these reactions were found to be 3.0 · 104, 3.2 · 105 and 1.2 · 105 M−1 · s−1, respectively, by the use of chronoamperometry, pulse polarography, differential pulse polarography and rotating-disk voltammetry. These reaction rates were analyzed by tunneling theory for electron transfers, and the comparisons between the theoretical and experimental values were quite good. The tunneling theory also worked quite well with other ferredoxin electron-transfer rates that are available in the literature. In addition to that, the activation enthalpy and entropy compared well with the tunneling theory.  相似文献   

14.
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   


15.
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   


16.
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   


17.
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   


18.
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   

19.
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   

20.
The fluorescence induction and other fluorescence properties of spinach chloroplasts at room temperature were probed utilizing two 30-ps wide laser pulses (530 nm) spaced Δt (s) apart in time (Δt = 5–110 ns). The energy of the first pulse (P1) was varied (1012–1016 photons · cm−2), while the energy of the second (probe) pulse (P2) was held constant (5 · 1013 photons · cm−2). A gated (10 ns) optical multichannel analyzer-spectrograph system allowed for the detection of the fluorescence generated either by P1 alone, or by P2 alone (preceded by P1). The dominant effect observed for the fluorescence yield generated by P1 alone is the usual singlet-singlet exciton annihilation which gives rise to a decrease in the yield at high energies. However, when the fluorescence yield of dark-adapted chloroplasts is measured utilizing P2 (preceded by pulse P1) an increase in this yield is observed. The magnitude of this increase depends on Δt, and is characterized by a time constant of 28 ± 4 ns. This rise in the fluorescence yield is attributed to a reduction of the oxidized (by P1) reaction center P-680+ by a primary donor. At high pulse energies (P1 = 4 · 1014 photons · cm−2) the magnitude of this fluorescence induction is diminished by another quenching effect which is attributed to triplet excited states generated by intense P1 pulses. Assuming that the P1 pulse energy dependence of the fluorescence yield rise reflects the closing of the reaction centers, it is estimated that about 3–4 photon hits per reaction center are required to close completely the reaction centers, and that there are 185–210 chlorophyll molecules per Photosystem II reaction center.  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