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.
The role of the protein environment in determining the redox midpoint potential (E(m)) of Q(A), the primary quinone of bacterial reaction centers, was investigated by mutation of isoleucine at position 265 of the M subunit in Rhodobacter sphaeroides. Isoleucine was changed to threonine, serine, and valine, yielding mutants M265IT, M265IS, and M265IV, respectively. All three mutants, with smaller residues replacing isoleucine, exhibited decreased binding affinities of the Q(A) site for various quinone analogues, consistent with an enlargement or loosening of the headgroup binding domain and a decrease in the van der Waals contact for small quinones. In all other respects, M265IV was like the wild type, but the polar mutants, M265IT and M265IS, had unexpectedly dramatic decreases in the redox midpoint potential of Q(A), resulting in faster rates of P(+)Q(A)(-) charge recombination. For both anthraquinone and native ubiquinone, the in situ E(m) of Q(A) was estimated to be approximately 100 and 85 mV lower in M265IT and M265IS, respectively. The effect on E(m)(Q(A)) indicates destabilization of the semiquinone or stabilization of the quinone. This is suggested to arise from either (i) electrostatic interaction between the partial charges or dipole of the residue hydroxyl group and the charge distribution of quinone and semiquinone states with particular influence near the C4 carbonyl group or (ii) from hydrogen bonding interactions between the hydroxyl oxygen and the N(delta)H of histidine M219, causing a weakening of the hydrogen bond to the C4 carbonyl. The rate of the first electron transfer (k(AB)(()(1)())) in the polar mutants was the same as in the wild type at low pH but decelerated at higher pH with altered pH dependence. The rate of the second electron transfer (k(AB)(()(2)())) was 3-4-fold greater than in the wild type over the whole pH range from 4 to 11, consistent with a larger driving force for electron transfer derived from the lower E(m) of Q(A).  相似文献   

2.
Quinones are naturally occurring isoprenoids that are widely exploited by photosynthetic reaction centers. Protein interactions modify the properties of quinones such that similar quinone species can perform diverse functions in reaction centers. Both type I and type II (oxygenic and nonoxygenic, respectively) reaction centers contain quinone cofactors that serve very different functions as the redox potential of similar quinones can operate at up to 800 mV lower reduction potential when present in type I reaction centers. However, the factors that determine quinone function in energy transduction remain unclear. It is thought that the location of the quinone cofactor, the geometry of its binding site, and the "smart" matrix effects from the surrounding protein environment greatly influence the functional properties of quinones. Photosystem II offers a unique system for the investigation of the factors that influence quinone function in energy transduction. It contains identical plastoquinones in the primary and secondary quinone acceptor sites, Q(A) and Q(B), which exhibit very different functional properties. This study is focused on elucidating the tuning and control of the primary semiquinone state, Q(A)(-), of photosystem II. We utilize high-resolution two-dimensional hyperfine sublevel correlation spectroscopy to directly probe the strength and orientation of the hydrogen bonds of the Q(A)(-) state with the surrounding protein environment of photosystem II. We observe two asymmetric hydrogen bonding interactions of reduced Q(A)(-) in which the strength of each hydrogen bond is affected by the relative nonplanarity of the bond. This study confirms the importance of hydrogen bonds in the redox tuning of the primary semiquinone state of photosystem II.  相似文献   

3.
The B3LYP hybrid density functional method is used to calculate spin densities and hyperfine couplings for the 1,4-naphthosemiquinone anion radical and a model of the phyllosemiquinone anion radical. The effect of hydrogen bonding on the spin density distribution is shown to lead to a redistribution of pi spin density from the semiquinone carbonyl oxygens to the carbonyl carbon atoms. The effect of in plane and out of plane hydrogen bonding is examined. Out of plane hydrogen bonding is shown to give rise to a significant delocalisation of spin density on to the hydrogen bond donor heavy atom. Excellent agreement is observed between calculated and experimental hyperfine couplings. Comparison of calculated hyperfine couplings with experimental determinations for the A1 phyllosemiquinone anion radical present in Photosystem I (PS I) of higher plant photosynthesis indicates that the in vivo radical may have a hydrogen bond to the O4 atom only as opposed to hydrogen bonds to each oxygen atom in alcohol solvents. The hydrogen bonding situation appears to be the reverse of that observed for QA in the bacterial type II reaction centres where the strong hydrogen bond occurs to the quinone O1 oxygen atom. For different types of reaction centre the presence or absence of the non-heme Fe(II) atom may well determine which type of hydrogen bonding situation prevails at the primary quinone site which in turn may influence the direction of subsequent electron transfer.  相似文献   

4.
The vital role of coenzyme Q in mitochondrial electron transfer and its regulation, and in energy conservation, is well established. However, the role of coenzyme Q in free oxyradical formation and as an antioxidant remains controversial. Demonstration of the existence of the semiquinone form of coenzyme Q during electron transport, coupled with recent evidence that hydrogen peroxide (but not molecular oxygen) may act as an oxidant of the semiquinone, suggests that the highly reactive OH. radical may be formed from the semiquinone. On the other hand, data exist implicating the Fe-S species as the source of electron transfer chain, free radical production. Additional data exist suggesting instead that the unpaired electron of the coenzyme Q semiquinone most likely dismutases superoxide radicals. These concepts and those arising from observations at several levels of organization including subcellular systems, intact animals, and human subjects in the clinical setting, supporting the concept of reduced coenzyme Q as an antioxidant, will be presented. The results of recent studies on the interaction between the two-electron quinone reductase--DT diaphorase and coenzyme Q10 will be presented. The possibility that superoxide dismutase may interact with reduced coenzyme Q, in conjunction with DT diaphorase inhibiting its autoxidation, will be described. The regulation of cellular coenzyme Q concentrations during oxidative stress accompanying aerobic exercise, resulting in increased protection from free radical damage, will also be presented.  相似文献   

5.
Madeo J  Gunner MR 《Biochemistry》2005,44(33):10994-11004
Bacterial reaction centers (RCs) catalyze a series of electron-transfer reactions reducing a neutral quinone to a bound, anionic semiquinone. The dissociation constants and association rates of 13 tailless neutral and anionic benzo- and naphthoquinones for the Q(A) site were measured and compared. The K(d) values for these quinones range from 0.08 to 90 microM. For the eight neutral quinones, including duroquinone (DQ) and 2,3-dimethoxy-5-methyl-1,4-benzoquinone (UQ(0)), the quinone concentration and solvent viscosity dependence of the association rate indicate a second-order rate-determining step. The association rate constants (k(on)) range from 10(5) to 10(7) M(-)(1) s(-)(1). Association and dissociation rate constants were determined at pH values above the hydroxyl pK(a) for five hydroxyl naphthoquinones. These negatively charged compounds are competitive inhibitors for the Q(A) site. While the neutral quinones reach equilibrium in milliseconds, anionic hydroxyl quinones with similar K(d) values take minutes to bind or dissociate. These slow rates are independent of ionic strength, solvent viscosity, and quinone concentration, indicating a first-order rate-limiting step. The anionic semiquinone, formed by forward electron transfer at the Q(A) site, also dissociates slowly. It is not possible to measure the association rate of the unstable semiquinone. However, as the protein creates kinetic barriers for binding and releasing anionic hydroxyl quinones without greatly increasing the affinity relative to neutral quinones, it is suggested that the Q(A) site may do the same for anionic semiquinone. Thus, the slow semiquinone dissociation may not indicate significant thermodynamic stabilization of the reduced species in the Q(A) site.  相似文献   

6.
Wells TA  Takahashi E  Wraight CA 《Biochemistry》2003,42(14):4064-4074
In the primary quinone (Q(A)) binding site of Rb. sphaeroides reaction centers (RCs), isoleucine M265 is in extensive van der Waals contact with the ubiquinone headgroup. Substitution of threonine or serine for this residue (mutants M265IT and M265IS), but not valine (mutant M265IV), lowers the redox midpoint potential of Q(A) by about 100 mV (Takahashi et al. (2001) Biochemistry 40, 1020-1028). The unexpectedly large effect of the polar substitutions is not due to reorientation of the methoxy groups as similar redox potential changes are seen for these mutants with either ubiquinone or anthraquinone as Q(A). Using FTIR spectroscopy to compare Q(A)(-)/Q(A) IR difference spectra for wild type and the M265 mutant RCs, we found changes in the polar mutants (M265IT and M265IS) in the quinone C[double bond]O and C[double bond]C stretching region (1600-1660 cm(-1)) and in the semiquinone anion band (1440-1490 cm(-1)), as well as in protein modes. Modeling the mutations into the X-ray structure of the wild-type RC indicates that the hydroxyl group of the mutant polar residues, Thr and Ser, is hydrogen bonded to the peptide C[double bond]O of Thr(M261). It is suggested that the mutational effect is exerted through the extended backbone region that includes Ala(M260), the hydrogen bonding partner to the C1 carbonyl of the quinone headgroup. The resulting structural perturbations are likely to include lengthening of the hydrogen bond between the quinone C1[double bond]O and the peptide NH of Ala(M260). Possible origins of the IR spectroscopic and redox potential effects are discussed.  相似文献   

7.
Mitchell's key insight that all bioenergetic membranes run on the conversion of redox energy into transmembrane electrical and proton gradients took the form 30 years ago of a working model of the Q cycle of cytochrome bc1, which operates reversibly on coupled electron and proton transfers of quinone at two binding sites on opposite membrane faces. His remarkable model still stands today, but he had no structural information to provide understanding into how dangerous short-circuit redox reactions were avoided. Now, it is clear that the Q cycle must be fixed with a special mechanism that avoids semiquinone-mediated short circuits. Either the redox states of the quinone electron-transfer partners double-gate the semiquinone-intermediate stability, or semiquinone is avoided altogether in concerted double-electron transfer.  相似文献   

8.
The cytochrome bo3 ubiquinol oxidase catalyzes the two-electron oxidation of ubiquinol in the cytoplasmic membrane of Escherichia coli, and reduces O2 to water. This enzyme has a high affinity quinone binding site (QH), and the quinone bound to this site acts as a cofactor, necessary for rapid electron transfer from substrate ubiquinol, which binds at a separate site (QL), to heme b. Previous pulsed EPR studies have shown that a semiquinone at the QH site formed during the catalytic cycle is a neutral species, with two strong hydrogen bonds to Asp-75 and either Arg-71 or Gln-101. In the current work, pulsed EPR studies have been extended to two mutants at the QH site. The D75E mutation has little influence on the catalytic activity, and the pattern of hydrogen bonding is similar to the wild type. In contrast, the D75H mutant is virtually inactive. Pulsed EPR revealed significant structural changes in this mutant. The hydrogen bond to Arg-71 or Gln-101 that is present in both the wild type and D75E mutant oxidases is missing in the D75H mutant. Instead, the D75H has a single, strong hydrogen bond to a histidine, likely His-75. The D75H mutant stabilizes an anionic form of the semiquinone as a result of the altered hydrogen bond network. Either the redistribution of charge density in the semiquinone species, or the altered hydrogen bonding network is responsible for the loss of catalytic function.  相似文献   

9.
The quinone chemistry that gives rise to the rather unusual strict bifurcation of electron transfer at the Q(o) site of the cytochrome bc complexes remains controversial. In this article, I review recent ideas and propose a "logic-gated" binding mechanism that combines classical quinone electrochemistry with specific hydrogen bonding requirements and results in a reversible reaction that minimizes unwanted side-reactions that could otherwise undermine the efficiency of the Q-cycle proton/electron coupling mechanism.  相似文献   

10.
Interruption of the menA or menB gene in Synechocystis sp. PCC 6803 results in the incorporation of a foreign quinone, termed Q, into the A(1) site of photosystem I with a number of experimental indicators identifying Q as plastoquinone-9. A global multiexponential analysis of time-resolved optical spectra in the blue region shows the following three kinetic components: 1) a 3-ms lifetime in the absence of methyl viologen that represents charge recombination between P700(+) and an FeS(-) cluster; 2) a 750-microseconds lifetime that represents electron donation from an FeS(-) cluster to methyl viologen; and 3) an approximately 15-microseconds lifetime that represents an electrochromic shift of a carotenoid pigment. Room temperature direct detection transient EPR studies of forward electron transfer show a spectrum of P700(+) Q(-) during the lifetime of the spin polarization and give no evidence of a significant population of P700(+) FeS(-) for t 相似文献   

11.
This review considers the state-of-the-art on mechanisms and alternative pathways of electron transfer in photosynthetic electron transport chains of chloroplasts and cyanobacteria. The mechanisms of electron transport control between photosystems (PS) I and II and the Calvin–Benson cycle are considered. The redistribution of electron fluxes between the noncyclic, cyclic, and pseudocyclic pathways plays an important role in the regulation of photosynthesis. Mathematical modeling of light-induced electron transport processes is considered. Particular attention is given to the electron transfer reactions on the acceptor side of PS I and to interactions of PS I with exogenous acceptors, including molecular oxygen. A kinetic model of PS I and its interaction with exogenous electron acceptors has been developed. This model is based on experimental kinetics of charge recombination in isolated PS I. Kinetic and thermodynamic parameters of the electron transfer reactions in PS I are scrutinized. The free energies of electron transfer between quinone acceptors A1A/A1B in the symmetric redox cofactor branches of PS I and iron–sulfur clusters FX, FA, and FB have been estimated. The second-order rate constants of electron transfer from PS I to external acceptors have been determined. The data suggest that byproduct formation of superoxide radical in PS I due to the reduction of molecular oxygen in the A1 site (Mehler reaction) can exceed 0.3% of the total electron flux in PS I.  相似文献   

12.
Electron transfer flavoprotein-ubiquinone oxidoreductase (ETF-QO) accepts electrons from electron transfer flavoprotein (ETF) and reduces ubiquinone from the ubiquinone pool. It contains one [4Fe-4S] (2+,1+) and one FAD, which are diamagnetic in the isolated oxidized enzyme and can be reduced to paramagnetic forms by enzymatic donors or dithionite. In the porcine protein, threonine 367 is hydrogen bonded to N1 and O2 of the flavin ring of the FAD. The analogous site in Rhodobacter sphaeroides ETF-QO is asparagine 338. Mutations N338T and N338A were introduced into the R. sphaeroides protein by site-directed mutagenesis to determine the impact of hydrogen bonding at this site on redox potentials and activity. The mutations did not alter the optical spectra, EPR g-values, spin-lattice relaxation rates, or the [4Fe-4S] (2+,1+) to FAD point-dipole interspin distances. The mutations had no impact on the reduction potential for the iron-sulfur cluster, which was monitored by changes in the continuous wave EPR signals of the [4Fe-4S] (+) at 15 K. For the FAD semiquinone, significantly different potentials were obtained by monitoring the titration at 100 or 293 K. Based on spectra at 293 K the N338T mutation shifted the first and second midpoint potentials for the FAD from +47 and -30 mV for wild type to -11 and -19 mV, respectively. The N338A mutation decreased the potentials to -37 and -49 mV. Lowering the midpoint potentials resulted in a decrease in the quinone reductase activity and negligible impact on disproportionation of ETF 1e (-) catalyzed by ETF-QO. These observations indicate that the FAD is involved in electron transfer to ubiquinone but not in electron transfer from ETF to ETF-QO. Therefore, the iron-sulfur cluster is the immediate acceptor from ETF.  相似文献   

13.
Ubiquinone-2 (UQ-2) selectively labeled with (13)C (I =(1)/(2)) at either the position 1- or the 4-carbonyl carbon is incorporated into the ubiquinol oxidase bo(3) from Escherichia coli in which the native quinone (UQ-8) has been previously removed. The resulting stabilized anion radical in the high-affinity quinone-binding site (Q(H)(*)(-)) is investigated using multifrequency (9, 34, and 94 GHz) electron paramagnetic resonance (EPR) spectroscopy. The corresponding spectra reveal dramatic differences in (13)C hyperfine couplings indicating a strongly asymmetric spin density distribution over the quinone headgroup. By comparison with previous results on labeled ubisemiquinones in proteins as well as in organic solvents, it is concluded that Q(H)(*)(-) is most probably bound to the protein via a one-sided hydrogen bond or a strongly asymmetric hydrogen-bonding network. This observation is discussed with regard to the function of Q(H) in the enzyme and contrasted with the information available on other protein-bound semiquinone radicals.  相似文献   

14.
The geometry of the hydrogen bonds to the two carbonyl oxygens of the semiquinone Q(A)(. -) in the reaction center (RC) from the photosynthetic purple bacterium Rhodobacter sphaeroides R-26 were determined by fitting a spin Hamiltonian to the data derived from (1)H and (2)H ENDOR spectroscopies at 35 GHz and 80 K. The experiments were performed on RCs in which the native Fe(2+) (high spin) was replaced by diamagnetic Zn(2+) to prevent spectral line broadening of the Q(A)(. -) due to magnetic coupling with the iron. The principal components of the hyperfine coupling and nuclear quadrupolar coupling tensors of the hydrogen-bonded protons (deuterons) and their principal directions with respect to the quinone axes were obtained by spectral simulations of ENDOR spectra at different magnetic fields on frozen solutions of deuterated Q(A)(. -) in H(2)O buffer and protonated Q(A)(. -) in D(2)O buffer. Hydrogen-bond lengths were obtained from the nuclear quadrupolar couplings. The two hydrogen bonds were found to be nonequivalent, having different directions and different bond lengths. The H-bond lengths r(OH) are 1.73 +/- 0.03 Angstrom and 1.60 +/- 0.04 Angstrom, from the carbonyl oxygens O(1) and O(4) to the NH group of Ala M260 and the imidazole nitrogen N(delta) of His M219, respectively. The asymmetric hydrogen bonds of Q(A)(. -) affect the spin density distribution in the quinone radical and its electronic structure. It is proposed that the H-bonds play an important role in defining the physical properties of the primary quinone, which affect the electron transfer processes in the RC.  相似文献   

15.
The bacterial reaction center couples light-induced electron transfer to proton pumping across the membrane by reactions of a quinone molecule Q(B) that binds two electrons and two protons at the active site. This article reviews recent experimental work on the mechanism of the proton-coupled electron transfer and the pathways for proton transfer to the Q(B) site. The mechanism of the first electron transfer, k((1))(AB), Q(-)(A)Q(B)-->Q(A)Q(-)(B), was shown to be rate limited by conformational gating. The mechanism of the second electron transfer, k((2))(AB), was shown to involve rapid reversible proton transfer to the semiquinone followed by rate-limiting electron transfer, H(+)+Q(-)(A)Q(-)(B) ifQ(-)(A)Q(B)H-->Q(A)(Q(B)H)(-). The pathways for transfer of the first and second protons were elucidated by high-resolution X-ray crystallography as well as kinetic studies showing changes in the rate of proton transfer due to site directed mutations and metal ion binding.  相似文献   

16.
The antimycin-sensitive ubisemiquinone radical (QC) of the ubiquinol-cytochrome c oxidoreductase of submitochondrial particles and chromatophores of Rhodopseudomonas sphaeroides Ga has been studied by a combination of redox potentiometry and EPR spectroscopy. This g = 2.005 radical signal appears at physiological pH values and increases in intensity with increasing pH up to pH 7.6 in submitochondrial particles and pH 9.0 in R. sphaeroides after which its intensity remains unchanged. The Em7 (ubiquinone/quinol) of the signal, estimated from redox titration data is 80 mV for submitochondrial particles, and 150 mV in chromatophores. Each of these values is higher than that of the quinone pool by 20 mV in submitochondrial particles and 60 mV in R. sphaeroides. This indicates that the quinone at the binding site is out of equilibrium with the pool, and that binding site preferentially binds quinol over quinone. Analysis of the shapes of the semiquinone titration curves, taken together with the midpoint elevation, indicates a quinone-binding site: cytochrome c1 stoichiometry of 1:1 in both submitochondrial particles and chromatophores. At its maximal intensity, the semiquinone concentration at the binding site is 0.26 in submitochondrial particles (greater than pH 7.6) and 0.4 in chromatophores (greater than pH 9.0). In both systems, the midpoint of the ubiquinone/ubisemiquinone couple is constant as the pH is raised up to the pH of maximal semiquinone formation whereafter it becomes more negative at the rate of -60 mV/pH unit. The midpoint of the ubisemiquinone/quinol couple, on the other hand, varies by -120 mV/pH unit at pH values up to the transition pH, after which it, too, changes by -60 mV/pH unit. This seemingly anomalous behavior may be explained by invoking a protonated group at or near the quinone-binding site whose pK corresponds to the pH transition point in the quinone/semiquinone/quinol redox chemistry when the site is free or when quinone or quinol occupies the site. This pK is elevated to at least pH 9.0 in submitochondrial particles and 10.5 in R. sphaeroides when semiquinone is bound to the site.  相似文献   

17.
In Rhodobacter sphaeroides reaction centers (RCs) containing the mutation Ala M260 to Trp (AM260W), transmembrane electron transfer along the full-length of the A-branch of cofactors is prevented by the loss of the Q(A) ubiquinone, but it is possible to generate the radical pair P(+)H(A)(-) by A-branch electron transfer or the radical pair P(+)Q(B)(-) by B-branch electron transfer. In the present study, FTIR spectroscopy was used to provide direct evidence for the complete absence of the Q(A) ubiquinone in mutant RCs with the AM260W mutation. Light-induced FTIR difference spectroscopy of isolated RCs was also used to probe the neutral Q(B) and the semiquinone Q(B)(-) states in two B-branch active mutants, a double AM260W-LM214H mutant, denoted WH, and a quadruple mutant, denoted WAAH, in which the AM260W, LM214H, and EL212A-DL213A mutations were combined. The data were compared to those obtained with wild-type (Wt) RCs and the double EL212A-DL213A (denoted AA) mutant which exhibit the usual A-branch electron transfer to Q(B). The Q(B)(-)/Q(B) spectrum of the WH mutant is very close to that of Wt RCs indicating similar bonding interactions of Q(B) and Q(B)(-) with the protein in both RCs. The Q(B)(-)/Q(B) spectra of the AA and WAAH mutants are also closely related to one another, but are very different to that of the Wt complex. Isotope-edited IR fingerprint spectra were obtained for the AA and WAAH mutants reconstituted with site-specific (13)C-labeled ubiquinone. Whilst perturbations of the interactions of the semiquinone Q(B)(-) with the protein are observed in the AA and WAAH mutants, the FTIR data show that the bonding interaction of neutral Q(B) in these two mutants are essentially the same as those for Wt RCs. Therefore, it is concluded that Q(B) occupies the same binding position proximal to the non-heme iron prior to reduction by either A-branch or B-branch electron transfer.  相似文献   

18.
The dimeric cytochrome bc(1) complex catalyzes the oxidation-reduction of quinol and quinone at sites located in opposite sides of the membrane in which it resides. We review the kinetics of electron transfer and inhibitor binding that reveal functional interactions between the quinol oxidation site at center P and quinone reduction site at center N in opposite monomers in conjunction with electron equilibration between the cytochrome b subunits of the dimer. A model for the mechanism of the bc(1) complex has emerged from these studies in which binding of ligands that mimic semiquinone at center N regulates half-of-the-sites reactivity at center P and binding of ligands that mimic catalytically competent binding of ubiquinol at center P regulates half-of-the-sites reactivity at center N. An additional feature of this model is that inhibition of quinol oxidation at the quinone reduction site is avoided by allowing catalysis in only one monomer at a time, which maximizes the number of redox acceptor centers available in cytochrome b for electrons coming from quinol oxidation reactions at center P and minimizes the leakage of electrons that would result in the generation of damaging oxygen radicals.  相似文献   

19.
The photoreduction of 1,4-benzoquinone (BQ), 1,4-naphthoquinone (NQ), 9,10-anthraquinone (AQ) and several derivatives, e.g. dimethylBQ, trimethylBQ, duroquinone, bromoNQ, methoxyNQ, methylAQ and dimethylAQ in acetonitrile-water by ascorbate was studied by time-resolved UV-vis spectroscopy using 20 ns laser pulses at 308 nm and continuous 254 nm irradiation. The semiquinone radical (*QH/Q*(-)) is formed after H-atom transfer from ascorbate to the quinone triplet state. The rate constant for quenching is k(q)=(2-9) x 10(9) M(-1) s(-1). Termination of the radicals takes place in the micros-ms range. The results are compared with those initiated by electron transfer from DABCO under similar conditions, where the k(q) values are similar, but the termination of Q*(-) takes place by electron back transfer not yielding hydroquinones. Specific properties of the quinone triplet state, e.g. self-quenching, nucleophilic water addition and the effects of structure are discussed.  相似文献   

20.
《BBA》1986,851(3):416-423
The ferrous ion associated with the electron acceptors in Photosystem II can be oxidized by the unstable semiquinone form of certain high-potential quinones (phenyl-p-benzoquinone, dimethylbenzoquinone and benzoquinone) which are used as electron acceptors. In a flash sequence, alternating oxidation of the iron by the photoreduced semiquinone on odd-numbered flashes is followed by photoreduction of the iron on even-numbered flashes. These reactions are detected by monitoring EPR signals arising from Fe3+. The oxidation of the iron can also occur in the frozen state (−30°C) indicating that the high-potential quinone can occupy the QB site. The reaction also takes place when the exogenous quinone is added in the dark to samples in which QB is already in the semiquinone form. The inhibitors of electron transfer between QA and QB, DCMU and sodium formate, block the photoreductant-induced iron oxidation. It is suggested that the iron oxidation takes place through the QB site. This unexpected photochemistry occurs under experimental conditions routinely used in studies of Photosystem II. Some previously reported phenomena can be reinterpreted on the basis of these new data.  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