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.
Rate constants and activation parameters have been determined for the internal electron transfer from type 1 (T1) to type 3 (T3) copper ions in laccase from both the fungus Trametes hirsuta and the lacquer tree Rhus vernicifera, using the pulse radiolysis method. The rate constant at 298 K and the enthalpy and entropy of activation were 25 ± 1 s(-1), 39.7 ± 5.0 kJ·mol(-1) and -87 ± 9 J·mol(-1) ·K(-1) for the fungal enzyme and 1.1 ± 0.1 s(-1), 9.8 ± 0.2 kJ·mol(-1) and -211 ± 3 J·mol(-1) ·K(-1) for the tree enzyme. The initial reduction of the T1 site by pulse radiolytically produced radicals was direct in the case of T. hirsuta laccase, but occured indirectly via a disulfide radical in R. vernicifera. The equilibrium constant that characterizes the electron transfer from T1 to T3 copper ions was 0.4 for T. hirsuta laccase and 1.5 for R. vernicifera laccase, leading to full reduction of the T1 site occurring at 2.9 ± 0.2 electron equivalents for T. hirsuta and 4 electron equivalents for R. vernicifera laccase. These results were compared with each other and with those for the same process in other multicopper oxidases, ascorbate oxidase and Streptomyces coelicolor laccase, using available structural information and electron transfer theory.  相似文献   

2.
Milligan JR  Tran NQ  Ly A  Ward JF 《Biochemistry》2004,43(17):5102-5108
Guanyl radical species are produced in DNA by electron removal caused by ionizing radiation, photoionization, oxidation, or photosensitization. DNA guanyl radicals can be reduced by electron donation from mild reducing agents. Important biologically relevant examples are the redox active amino acids cysteine, cystine, methionine, tryptophan, and tyrosine. We have quantified the reactivity of derivatives of these amino acids with guanyl radicals located in plasmid DNA. The radicals were produced by electron removal using the single electron oxidizing agent (SCN)(2)(*)(-). Disulfides (cystine) are unreactive. Thioethers (methionine), thiols (cysteine), and phenols (tyrosine) react with rate constants in the range 10(4)-10(6), 10(5)-10(6), and 10(5)-10(6) dm(3) mol(-1) s(-1), respectively. Indoles (tryptophan) are the most reactive with rate constants of 10(7)-10(8) dm(3) mol(-1) s(-1). Selenium analogues of amino acids are over an order of magnitude more reactive than their sulfur equivalents. Increasing positive charge is associated with a ca. 10-fold increase in reactivity. The results suggest that amino acid residues located close to DNA (for example, in DNA binding proteins such as histones) might participate in the repair of oxidative DNA damage.  相似文献   

3.
Horseradish peroxidase (HRP) was successfully immobilized on vertically oriented TiO(2) nanotube arrays (NTAs), which was prepared by a seeded-growth mechanism. The nanotubular structure of TiO(2) was characterized by scanning electron microscope (SEM). After encapsulated HRP on TiO(2) nanotube arrays, the direct electron transfer of HRP was observed. Owing to the redox reaction of electroactive center of HRP, the HRP/TiO(2) NTAs modified electrode exhibited a pair of quasi-reversible peaks with the peak-to-peak separation of 70mV and the formal potential of -0.122V (vs. SCE) in 0.2molL(-1) phosphate buffer solution (PBS, pH 7.0). The number of transference electron was 0.84 and the direct electron transfer (ET) constant (k(s)) was 3.82s(-1). The HRP/TiO(2) NTAs modified electrode displayed an excellent electrocatalytic performance for H(2)O(2) and the formal Michaelis-Menten constant (K(m)(app)) was 1.9mmolL(-1). The response currents had a good linear relation with the concentration of H(2)O(2) from 5.0x10(-7)molL(-1) to 1.0x10(-5)molL(-1) and 5.0x10(-5)molL(-1) to 1.0x10(-3)molL(-1), respectively.  相似文献   

4.
The disintegration of the radionuclides (131)I and (125)I and the subsequent charged-particle tracks left behind in water (as a model substance for a biological cell) are simulated by the Monte Carlo track structure simulation code PARTRAC, using new inelastic electron scattering cross sections for condensed water. Every photon and electron emitted was followed in detail, event by event, down to 10 eV. From the spatial information on the track structures, absorbed dose distributions per (131)I and (125)I decay were calculated in and around water spheres simulating micrometastases as well as in the tissue surrounding such metastases. These radionuclides were assumed to be distributed uniformly inside spheres of different diameters (0.01, 0.03, 0.1, 0.3, 1.0 and 3.0 mm). The respective electron degradation spectra, the nearest-neighbor distance distributions between inelastic events, and the distance distributions for all activations for both iodine radionuclides were calculated. The absorbed fractions of the initial electron energies, absorbed doses and energy depositions, and single-event distributions, F(1)(epsilon), inside the six water spheres described above and in the surrounding tissue were also calculated. The absorbed doses per decay inside the six water spheres, i.e., the calculated S values (listed from 0.01 to 3.0 mm), were 6.8 x 10(-4), 7.2 x 10(-5), 5.5 x 10(-6), 4.9 x 10(-7), 3.1 x 10(-8) and 1.8 x 10(-9) Gy Bq(-1) s(-1) for (131)I, and 3.4 x 10(-3), 1.7 x 10(-4), 5.1 x 10(-6), 2.0 x 10(-7), 5.6 x 10(-9) and 2.2 x 10(-10) Gy Bq(-1) s(-1) for (125)I. It is concluded that, in the treatment of thyroid cancer, the geometrical track structure properties of (125)I might be superior to those of (131)I in micrometastases with diameters less than 0.1 mm; however, in this medical context, many other factors also have to be considered.  相似文献   

5.
G J Harnadek  E A Ries  D Njus 《Biochemistry》1985,24(11):2640-2644
The chromaffin vesicle of the adrenal medulla contains a transmembrane electron carrier that may provide reducing equivalents for dopamine beta-hydroxylase in vivo. This electron-transfer system can be assayed by trapping ascorbic acid inside resealed membrane vesicles (ghosts), adding an external electron acceptor such as ferricytochrome c or ferricyanide, and following the reduction of these acceptors spectrophotometrically. Cytochrome c reduction is more rapid at high pH and is proportional to the amount of chromaffin-vesicle ghosts, at least at low ghost concentrations. At pH 7.0, ghosts loaded with 100 mM ascorbic acid reduce 60 microM cytochrome c at a rate of 0.035 +/- 0.010 mu equiv min-1 (mg of protein)-1 and 200 microM ferricyanide at a rate of 2.3 +/- 0.3 mu equiv min-1 (mg of protein)-1. The rate of cytochrome c reduction is accelerated to 0.105 +/- 0.021 mu equiv min-1 (mg of protein)-1 when cytochrome c is pretreated with equimolar ferrocyanide. Pretreatment of cytochrome c with ferricyanide also causes a rapid rate of reduction, but only after an initial delay. The ferrocyanide-stimulated rate of cytochrome c reduction is further accelerated by the protonophore carbonyl cyanide p-(trifluoromethoxy)phenylhydrazone (FCCP), probably because FCCP dissipates the membrane potential generated by electron transfer. These rates of electron transfer are sufficient to account for electron transfer to dopamine beta-hydroxylase in vivo and are consistent with the mediation of electron transfer by cytochrome b-561.  相似文献   

6.
Recent studies of point mutations in photosystem I have suggested that the two kinetic phases of phylloquinone reoxidation represent electron transfer in the two branches of cofactors. This interpretation implies that changes in the relative amplitudes of the two kinetic phases represent a change in the extent of electron transfer in the two branches. Using time-resolved electron paramagnetic resonance (EPR), this issue is investigated in subunit deletion mutants of Synechococcus sp. PCC 7002. The spin-polarized EPR signals of P(700)(+)A(1)(-) and P(700)(+)FeS(-), both at room temperature and in frozen solution, are altered by deletion of PsaF and/or PsaE, and the differences from the wild type are much more pronounced in PS I complexes isolated from the mutants using Triton X-100 rather than n-dodecyl beta-d-maltopyranoside. The changes in the transient EPR data for the mutant complexes are consistent with a significant fraction of reaction centers showing (i) faster electron transfer from A(1)(-) to F(X), (ii) slower forward electron transfer from A(0)(-) to A(1), and (iii) slightly altered quinone hyperfine couplings, possibly as a result of a change in the hydrogen bonding. The fraction of fast electron transfer and its dependence on the isolation procedure are estimated approximately from simulations of the room temperature EPR data. The results are discussed in terms of possible models for the electron transfer. It is suggested that the detergent-induced fraction of fast electron transfer is most likely due to alteration of the environment of the quinone in the PsaA branch of cofactors and is not the result of a change in the directionality of electron transfer.  相似文献   

7.
Soluble methane monooxygenase (sMMO) from Methylococcus capsulatus (Bath) is a three-component enzyme system that catalyzes the conversion of methane to methanol. A reductase (MMOR), which contains [2Fe-2S] and FAD cofactors, facilitates electron transfer from NADH to the hydroxylase diiron active sites where dioxygen activation and substrate hydroxylation take place. By separately expressing the ferredoxin (MMORFd, MMOR residues 1-98) and FAD/NADH (MMOR-FAD, MMOR residues 99-348) domains of the reductase, nearly all biochemical properties of full-length MMOR are retained, except for interdomain electron transfer rates. To investigate the extent to which rapid electron transfer between domains might be restored and further to explore the modularity of MMOR, MMOR-Fd and MMOR-FAD were connected in a non-native fashion. Four different linker sequences were employed to create MMOR reversed-domain (MMOR-RD) constructs, MMOR(99-342)-linker-MMOR(2-98), with a domain connectivity observed in other homologous oxidoreductases. The optical, redox, and electron transfer properties of the four MMOR-RD proteins were characterized and compared with those of wild-type MMOR. The linker sequence plays a key role in controlling solvent accessibility to the FAD cofactor, as evidenced by perturbed flavin optical spectra, decreased FADox/FADsq redox potentials, and increased steady-state oxidase activities in three of the constructs. Stopped-flow optical spectroscopy revealed slow interdomain electron transfer (k < 0.04 s(-1) at 4 degrees C, compared with 90 s(-1) for wild-type MMOR) for all three MMOR-RD proteins with 7-residue linkers. A long (14-residue), flexible linker afforded much faster electron transfer between the FAD and [2Fe-2S] cofactors (k = 0.9 s(-1) at 4 degrees C).  相似文献   

8.
9.
It has been shown that efficient functioning of photosynthesis and respiration in the cyanobacterium Synechocystis PCC 6803 requires the presence of either cytochrome c6 or plastocyanin. In order to check whether the blue copper protein plastocyanin can act as electron donor to cytochrome c oxidase, we investigated the intermolecular electron transfer kinetics between plastocyanin and the soluble CuA domain (i.e. the donor binding and electron entry site) of subunit II of the aa3-type cytochrome c oxidase from Synechocystis. Both copper proteins were expressed heterologously in Escherichia coli. The forward and the reverse electron transfer reactions were studied yielding apparent bimolecular rate constants of (5.1+/-0.2) x 10(4) M(-1) s(-1) and (8.5+/-0.4) x 10(5) M(-1) s(-1), respectively (20 mM phosphate buffer, pH 7). This corresponds to an apparent equilibrium constant of 0.06 in the physiological direction (reduction of CuA), which is similar to Keq values calculated for the reaction between c-type cytochromes and the soluble fragments of other CuA domains. The potential physiological role of plastocyanin in cyanobacterial respiration is discussed.  相似文献   

10.
We present measurements of low-energy (<18 eV) electron-stimulated desorption of anions from acetamide (CH(3)CONH(2)) and dimethyl disulfide [DMDS: (CH(3)S)(2)] films. Electron irradiation of physisorbed CH(3)CONH(2) produces H(-), CH(3)(-) and O(-) anions, whereas the H(-), CH(2)(-), CH(3)(-), S(-), SH(-) and SCH(3)(-) anions are observed to desorb from the DMDS film. Below 12 eV, the dependence of the anion yields on the incident electron energy exhibits structures that indicate that a resonant process (i.e. dissociative electron attachment) is responsible for molecular fragmentation. Within the range of 1-18 eV, it is found that (1.7 and 1.4) x 10(7) H(-) ions/incident electron and (7.8 x 10(-11) and 4.3 x 10(-8)) of the other ions/incident electron are desorbed from acetamide and DMDS films, respectively. These results suggest that, within proteins, the disulfide bond is more sensitive to low-energy electron attack than the peptide bond. In biological cells, some proteins interact closely with nucleic acid. Therefore, the observed fragments, when produced from secondary low-energy electrons generated by high-energy radiation, not only may denature proteins, but may also induce reactions with the nearby nucleic acid and damage DNA.  相似文献   

11.
The recent crystal structure of photosystem I (PSI) from Thermosynechococcus elongatus shows two nearly symmetric branches of electron transfer cofactors including the primary electron donor, P(700), and a sequence of electron acceptors, A, A(0) and A(1), bound to the PsaA and PsaB heterodimer. The central magnesium atoms of each of the putative primary electron acceptor chlorophylls, A(0), are unusually coordinated by the sulfur atom of methionine 688 of PsaA and 668 of PsaB, respectively. We [Ramesh et al. (2004a) Biochemistry 43:1369-1375] have shown that the replacement of either methionine with histidine in the PSI of the unicellular green alga Chlamydomonas reinhardtii resulted in accumulation of A(0)(-) (in 300-ps time scale), suggesting that both the PsaA and PsaB branches are active. This is in contrast to cyanobacterial PSI where studies with methionine-to-leucine mutants show that electron transfer occurs predominantly along the PsaA branch. In this contribution we report that the change of methionine to either leucine or serine leads to a similar accumulation of A(0)(-) on both the PsaA and the PsaB branch of PSI from C. reinhardtii, as we reported earlier for histidine mutants. More importantly, we further demonstrate that for all the mutants under study, accumulation of A(0)(-) is transient, and that reoxidation of A(0)(-) occurs within 1-2 ns, two orders of magnitude slower than in wild type PSI, most likely via slow electron transfer to A(1). This illustrates an indispensable role of methionine as an axial ligand to the primary acceptor A(0) in optimizing the rate of charge stabilization in PSI. A simple energetic model for this reaction is proposed. Our findings support the model of equivalent electron transfer along both cofactor branches in Photosystem I.  相似文献   

12.
Room temperature transient EPR spectra of photosystem I (PS I) particles from Synechocystis 6803 are presented. Native PS I samples and preparations depleted in the A1-acceptor site by solvent extraction and then reconstituted with the quinones (Q) vitamin K1 (VK1), duroquinone (DQ and DQd12) and naphthoquinone (NQ) have been studied. Sequential electron transfer to P700+A1- (FeS) and P700+A1 (FeS)- is recovered only with VK1. With DQ and NQ electron transfer is restored to form the radical pair P700+Q- as specified by a characteristic electron spin polarization (ESP)-pattern, but further electron transfer is either slowed down or blocked. A qualitative analysis of the K-band spectrum suggests that the orientation of reconstituted NQ in PS I is different from the native acceptor A1 = VK1.  相似文献   

13.
A competitive PCR (cPCR) assay targeting 16S ribosomal DNA was developed to enumerate growth of a Dehalococcoides-like microorganism, bacterium VS, from a mixed culture catalyzing the reductive dehalogenation of cis-1,2-dichloroethene (cDCE) and vinyl chloride (VC), with hydrogen being used as an electron donor. The growth of bacterium VS was found to be coupled to the dehalogenation of VC and cDCE, suggesting unique metabolic capabilities. The average growth yield was (5.2 +/- 1.5) x 10(8) copies of the 16S rRNA gene/ micromol of Cl(-) (number of samples, 10), with VC being used as the electron acceptor and hydrogen as the electron donor. The maximum VC utilization rate (q) was determined to be 7.8 x 10(-10) micromol of Cl(-) (copy(-1) day(-1)), indicating a maximum growth rate of 0.4 day(-1). These average growth yield and q values agree well with values found previously for dechlorinating cultures. Decay coefficients were determined with growth (0.05 day(-1)) and no-growth (0.09 day(-1)) conditions. An important limitation of this cPCR assay was its inability to discriminate between active and inactive cells. This is an essential consideration for kinetic studies.  相似文献   

14.
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   

15.
Shewanella putrefaciens strain 200 respires a wide range of compounds as terminal electron acceptor. The respiratory versatility of Shewanella is attributed in part to a set of c-type cytochromes with widely varying midpoint redox potentials (E'(0)). A point mutant of S. putrefaciens, originally designated Urr14 and here renamed CCMB1, was found to grow at wild-type rates on electron acceptors with high E'0 [O2, NO3-, Fe(III) citrate, MnO2, and Mn(III) pyrophosphate] yet was severely impaired for growth on electron acceptors with low E'0 [NO2-, U(VI), dimethyl sulfoxide, TMAO (trimethylamine N-oxide), fumarate, gamma-FeOOH, SO3(2-), and S2O3(2-)]. Genetic complementation and nucleotide sequence analyses indicated that the CCMB1 respiratory mutant phenotype was due to mutation of a conserved histidine residue (H108Y) in a protein that displayed high homology to Escherichia coli CcmB, the permease subunit of an ABC transporter involved in cytochrome c maturation. Although CCMB1 retained the ability to grow on electron acceptors with high E'(0), the cytochrome content of CCMB1 was <10% of that of the wild-type strain. Periplasmic extracts of CCMB1 contained slightly greater concentrations of the thiol functional group (-SH) than did the wild-type strain, an indication that the E(h) of the CCMB1 periplasm was abnormally low. A ccmB deletion mutant was unable to respire anaerobically on any electron acceptor, yet retained aerobic respiratory capability. These results suggest that the mutation of a conserved histidine residue (H108) in CCMB1 alters the redox homeostasis of the periplasm during anaerobic growth on electron acceptors with low (but not high) E'0. This is the first report of the effects of Ccm deficiencies on bacterial respiration of electron acceptors whose E'0 nearly span the entire redox continuum.  相似文献   

16.
To investigate the environment of the phylloquinone secondary electron acceptor A(1) within the photosystem I reaction center, we have carried out site-directed mutagenesis of two tryptophan residues (W693 and W702) in the PsaA subunit of Chlamydomonas reinhardtii. One of these conserved tryptophans (W693) is predicted to be close to the phylloquinone and has been implicated in the interaction of A(1) with an aromatic residue through pi--pi stacking. We find that replacement of W702 with either histidine or leucine has no effect on the electronic structure of A(1)(*-) or on forward electron transfer from A(1)(*-) to the iron--sulfur center F(x). In contrast, the same mutations of W693 alter the electronic structure of the photoaccumulated A(1)(*-) and slow forward electron transfer as measured by the decay of the electron spin-polarized signal arising from the P700(*+)/A(1)(*-) radical pair. These results provide support for the hypothesis that W693 has a role in poising the redox potential of A(1)/A(1)(*-) so it can reduce F(x), and they indirectly provide evidence for electron transfer along the PsaA-side branch of cofactors in PSI.  相似文献   

17.
Dihydroorotate dehydrogenase (DHOD) (EC 1.3.3.1) from the thermoacidophilic archaeon Sulfolobus solfataricus P2 (DSM 1617) was partially purified 3,158-fold, characterized, and the encoding genes identified. Based on enzymological as well as phylogenetic methods, dihydroorotate dehydrogenase from S. solfataricus (DHODS) represents a new type of DHOD, type 1S. Furthermore, it is unable to use any of the (type-specific) natural electron acceptors employed by all other presently known DHODs. DHODS shows optimal activity at 70 degrees C in the pH range 7-8.5. It is capable of using ferricyanide, 2,6-dichlorophenolindophenol (DCIP), Q(0), and molecular oxygen as electron acceptor. Kinetic studies employing ferricyanide indicate a two-site ping-pong mechanism with K(M) values of 44.2+/-1.9 microM for the substrate dihydroorotate and 344+/-21 microM for the electron acceptor ferricyanide, as well as competitive product inhibition with a K(i) of 23.7+/-3.4 microM for the product orotate (OA). The specific activity, as determined from a partially purified sample, is approximately 20 micromol mg(-1) min(-1). DHODS is a heteromeric enzyme comprising a catalytic subunit encoded by pyrD (291 aa; MW=31.1 kDa) and an electron acceptor subunit (208 aa; MW=23.6 kDa), encoded by orf1. DHODS employs a serine as catalytic base, which is unique for a cytosolic DHOD. To our knowledge, this work represents not only the first study on an archaeal DHOD but the first on a nonmesophilic DHOD as well.  相似文献   

18.
Kinetics of perchlorate- and chlorate-respiring bacteria   总被引:2,自引:0,他引:2  
Ten chlorate-respiring bacteria were isolated from wastewater and a perchlorate-degrading bioreactor. Eight of the isolates were able to degrade perchlorate, and all isolates used oxygen and chlorate as terminal electron acceptors. The growth kinetics of two perchlorate-degrading isolates, designated "Dechlorosoma" sp. strains KJ and PDX, were examined with acetate as the electron donor in batch tests. The maximum observed aerobic growth rates of KJ and PDX (0.27 and 0.28 h(-1), respectively) were only slightly higher than the anoxic growth rates obtained by these isolates during growth with chlorate (0.26 and 0.21 h(-1), respectively). The maximum observed growth rates of the two non-perchlorate-utilizing isolates (PDA and PDB) were much higher under aerobic conditions (0.64 and 0.41 h(-1), respectively) than under anoxic (chlorate-reducing) conditions (0.18 and 0.21 h(-1), respectively). The maximum growth rates of PDX on perchlorate and chlorate were identical (0.21 h(-1)) and exceeded that of strain KJ on perchlorate (0.14 h(-1)). Growth of one isolate (PDX) was more rapid on acetate than on lactate. There were substantial differences in the half-saturation constants measured for anoxic growth of isolates on acetate with excess perchlorate (470 mg/liter for KJ and 45 mg/liter for PDX). Biomass yields (grams of cells per gram of acetate) for strain KJ were not statistically different in the presence of the electron acceptors oxygen (0.46 +/- 0.07 [n = 7]), chlorate (0.44 +/- 0.05 [n = 7]), and perchlorate (0.50 +/- 0.08 [n = 7]). These studies provide evidence that facultative microorganisms with the capability for perchlorate and chlorate respiration exist, that not all chlorate-respiring microorganisms are capable of anoxic growth on perchlorate, and that isolates have dissimilar growth kinetics using different electron donors and acceptors.  相似文献   

19.
The synergistic antioxidant mechanism of alpha-tocopherol (vitamin E) with green tea polyphenols, i.e., (-)-epicatechin (EC), (-)-epigallocatechin (EGC), (-)-epicatechin gallate (ECG), (-)-epigallocatechin gallate (EGCG), and gallic acid (GA), was studied by assaying the kinetics of the reaction of alpha-tocopheroxyl radical with green tea polyphenols by stopped-flow electron paramagnetic resonance, the inhibition of linoleic acid peroxidation by these antioxidants, and the decay of alpha-tocopherol during the peroxidation. It was found that the green tea polyphenols could reduce alpha-tocopheroxyl radical to regenerate alpha-tocopherol with rate constants of 0.45, 1.11, 1.31, 1.91, and 0.43 x 10(2) M(-1) s(-1) for EC, EGC, ECG, EGCG, and GA, respectively, in sodium dodecyl sulfate micelles. In addition, these second-order rate constants exhibited a good linear correlation with their oxidation potentials, suggesting that electron transfer might play a role in the reaction.  相似文献   

20.
Cucumber leaf discs were illuminated at room-temperature with far-red light to photo-oxidise P700, the chlorophyll dimer in Photosystem (PS) I. The post-illumination kinetics of P700(+) re-reduction were studied in the presence of inhibitors or cofactors of photosynthetic electron transport. The re-reduction kinetics of P700(+) were well fitted as the sum of three exponentials, each with its amplitude and rate coefficient, and an initial flux (at the instant of turning off far-red light) given as the product of the two. Each initial flux is assumed equal to a steady state flux during far-red illumination. The fast phase of re-reduction, with rate coefficient k (1) approximately 10 s(-1), was completely abolished by a saturating concentration of 3-(3,4-dichlorophenyl)-1,1-dimethylurea (DCMU); it is attributed to electron flow to P700(+) from PS II, which was stimulated to some extent by far-red light. The intermediate phase, with rate coefficient k (1) approximately 1 s(-1), was only partly diminished by methyl viologen (MV) which diverts electron flow to oxygen. The intermediate phase is attributed to electron donation from reduced ferredoxin to the intersystem pool; reduced ferredoxin could be formed: (1) directly by electron donation on the acceptor of PS I; and/or (2) indirectly by stromal reductants, in line with only a partial inhibition of the intermediate phase by MV. Duroquinol enhanced the intermediate phase in the presence of DCMU, presumably through its interaction with thylakoid membrane components leading to the partial reduction of plastoquinone. The slow phase of P700(+) re-reduction, with rate coefficient k (1) approximately 0.1 s(-1), was unaffected by DCMU and only slightly affected by MV; it could be associated with electron donation to either: (1) the intersystem chain by stromal reductants catalysed by NAD(P)H dehydrogenase slowly; or (2) plastocyanin/P700(+) by ascorbate diffusing across the thylakoid membrane to the lumen. It is concluded that a post-illumination analysis of the fluxes to P700(+) can be used to probe the pathways of electron flow to PS I in steady state illumination.  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