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.
1. Evidence is presented for the presence of a stable ubisemiquinone pair in the vicinity of iron-sulphur centre S-3, based on its thermodynamic and spin relaxation properties. 2. These semiquinones are coupled by dipolar interaction; quantitative analysis of the signals of the spin-coupled semiquinones (at pH 7.4) gives midpoint redox potentials E1 (oxidized to semiquinone state) and E2 (semiquinone to fully reduced state) of 140 and 80mV, respectively, for individual ubiquinones. 3. Values of pKS (pK of the semiquinone form) below 6.5 and pKR (pK of the fully reduced ubiquinone) of about 8.0 or above were estimated from the pH-dependence of the midpoint potentials of the spin coupled signals. Thus the ubisemiquinone associated with succinate dehydrogenase (designated as SQS) functions mostly in the anionic form of the physiological pH range. 4. Theonyltrifluoroacetone, a specific inhibitor of the succinate-ubiquinone reductase segment of the respiratory chain, destabilized the intermediate redox state; thus it quenches both the g = 2.00 signal and ubisemiquinone (SQS) and split signals from the spin coupled pair. This inhibitor has no significant effect on another bound ubisemiquinone species present in the cytochrome bc1 region (designated as SQC). 5. The possible function and location of these stabilized ubisemiquinone species were discussed in connection with Site-II energy transduction.  相似文献   

2.
The ubiquinol-cytochrome c oxidoreductase (cytochrome bc1) complex from Paracoccus denitrificans exhibits a thermodynamically stable ubisemiquinone radical detectable by EPR spectroscopy. The radical is centered at g = 2.004, is sensitive to antimycin, and has a midpoint potential at pH 8.5 of +42 mV. These properties are very similar to those of the stable ubisemiquinone (Qi) previously characterized in the cytochrome bc1 complexes of mitochondria. The micro-environment of the Rieske iron-sulfur cluster in the Paracoccus cytochrome bc1 complex changes in parallel with the redox state of the ubiquinone pool. This change is manifested as shifts in the gx, gy, and gz values of the iron-sulfur cluster EPR signal from 1.80, 1.89, and 2.02 to 1.76, 1.90, and 2.03, respectively, as ubiquinone is reduced to ubiquinol. The spectral shift is accompanied by a broadening of the signal and follows a two electron reduction curve, with a midpoint potential at pH 8.5 of +30 mV. A hydroxy analogue of ubiquinone, UHDBT, which inhibits respiration in the cytochrome bc1 complex, shifts the gx, gy, and gz values of the iron-sulfur cluster EPR signal to 1.78, 1.89, and 2.03, respectively, and raises the midpoint potential of the iron-sulfur cluster at pH 7.5 from +265 to +320 mV. These changes in the micro-environment of the Paracoccus Rieske iron-sulfur cluster are like those elicited in mitochondria. These results indicate that the cytochrome bc1 complex of P. denitrificans has a binding site for ubisemiquinone and that this site confers properties on the bound ubisemiquinone similar to those in mitochondria. In addition, the line shape of the Rieske iron-sulfur cluster changes in response to the oxidation-reduction status of ubiquinone, and the midpoint of the iron-sulfur cluster increases in the presence of a hydroxyquinone analogue of ubiquinone. The latter results are also similar to those observed in the mitochondrial cytochrome bc1 complex. However, unlike the mitochondrial complexes, which contain eight to 11 polypeptides and are thought to contain distinct quinone binding proteins, the Paracoccus cytochrome bc1 complex contains only three polypeptide subunits, cytochromes b, c1, and iron-sulfur protein. The ubisemiquinone binding site and the site at which ubiquinone and/or ubiquinol bind to affect the Rieske iron-sulfur cluster in Paracoccus thus exist in the absence of any distinct quinone binding proteins and must be composed of domains contributed by the cytochromes and/or iron-sulfur protein.  相似文献   

3.
The redox midpoint potential (E (m)) of the primary quinone of bacterial reaction centers, Q(A), in native membranes (chromatophores) measured by redox potentiometry is reported to be pH dependent (-60 mV/pH) up to a highly distinctive pK ( a ) (9.8 in Rba. sphaeroides) for the reduced state. In contrast, the E (m) of Q(A) in isolated RCs of Rba. sphaeroides, although more variable, has been found to be essentially pH-independent by both redox potentiometry and by delayed fluorescence, which determines the free energy (DeltaG (P*A)) of the P(+)Q (A) (-) state relative to P*. Delayed fluorescence was used here to determine the free energy of P(+)Q (A) (-) in chromatophores. The emission intensity in chromatophores is two orders of magnitude greater than from isolated RCs largely due to the entropic effect of antenna pigments "drawing out" the excitation from the RC. The pH dependence of DeltaG (P*A) was almost identical to that of isolated RCs, in stark contrast with potentiometric redox titrations of Q(A). We considered that Q(A) might be reduced by disproportionation with QH(2) through the Q(B) site, so the titration actually reflects the quinone pool, giving the -60 mV/pH unit dependence expected for the Q/QH(2) couple. However, the parameters necessary to achieve a strong pH-dependence are not in good agreement with expected properties of Q(A) and Q(B). We also consider the possibility that the time scale of potentiometric titrations allows the reduced state (Q (A) (-) ) to relax to a different conformation that is accompanied by stoichiometric H(+) binding. Finally, we discuss the choice of parameters necessary for determining the free energy level of P(+)Q (A) (-) from delayed fluorescence emission from chromatophores of Rba. sphaeroides.  相似文献   

4.
Absorbance changes at 450 nm of the semiquinone form of the secondary electron acceptor were studied in chromatophores of Rhodospirillum rubrum. When chromatophores are illuminated by a series of single turnover flashes ubisemiquinone is formed and destroyed on alternate flashes at ambient redox potential from 100 to 250 mV. A simple kinetic model of the binary oscillations is suggested. On the base of the model it is shown that the rate constant of electron transfer from primary to secondary quinone after the first flash is larger that after the second flash. Cooperativity in electron transfer from primary to secondary quinone can be explained by electrostatic interactions of charged carriers.  相似文献   

5.
We present an electron nuclear double resonance (ENDOR) study of the bound Qc.- ubisemiquinone in the mitochondrial quinol cytochrome c reductase complex. An ENDOR probe specifically modified for insertion into our electron paramagnetic resonance cavity was used for this study. We observed strongly hyperfine-coupled protons whose exchangeable nature indicated they were hydrogen-bonded to the quinone oxygen(s). It is thought that such hydrogen bonds are critical in binding the ubiquinone to protein, in stabilizing its semiquinone form, and in modulating the thermodynamic properties of the bound ubiquinone in the mitochondrial quinol cytochrome c reductase complex. Additional ENDOR features were assigned to protons of the quinone ring itself and to weakly coupled protons that may be associated with nearby amino acids. From very weakly hyperfine-coupled, distant, exchangeable protons there was also ENDOR evidence to suggest proximity and accessibility of the ubiquinone site to the solvent.  相似文献   

6.
(1) The kinetics of the reduction by duroquinol of the prosthetic groups of QH2:cytochrome c oxidoreductase and of the formation of ubisemiquinone have been studied using a combination of the freeze-quench technique, low-temperature diffuse-reflectance spectroscopy, EPR and stopped flow. (2) The formation of the antimycin-sensitive ubisemiquinone anion parallels the reduction of both high-potential and low-potential cytochrome b-562. (3) The rates of reduction of both the [2Fe-2S] clusters and cytochromes (c + c1) are pH dependent. There is, however, a pH-dependent discrepancy between their rate of reduction, which can be correlated with the difference in pH dependencies of their midpoint potentials. (4) Lowering the pH or the Q content results in a slower reduction of part of the [2Fe-2S] clusters. It is suggested that one cluster is reduced by a quinol/semiquinone couple and the other by a semiquinone/quinone couple. (5) Myxothiazol inhibits the reduction of the [2Fe-2S] clusters, cytochrome c1 and high-potential cytochrome b-562. (6) The results are consistent with a Q-cycle model describing the pathway of electrons through a dimeric QH2:cytochrome c oxidoreductase.  相似文献   

7.
Oxidation-reduction potentiometry was carried out on Rhodopseudomonas viridis chromatophores. Measurements of e.p.r. signals of the semiquinone-iron type at g=1.82 have revealed a more complex situation than previously reported. The presence of three different components is indicated. The midpoint potential (E(m)) of the primary acceptor quinone/semiquinone couple was found to be approx. -165mV at pH10, with a pK being reached at around pH7.5. The primary acceptor also accepts a second electron with an E(m) of -525mV, but this redox transition exhibits a hysteresis effect. Interaction effects indicate the presence of another component with E(m) values at pH10 of approx. -165mV (pK reached at around pH7.5) for single reduction and -350mV (pK at pH10 or greater) for double reduction. It is suggested that this component is the secondary acceptor. Another semiquinone-iron-type component which gives a g=1.82 signal is also present. This component is distinguishable from the primary acceptor by its e.p.r. spectrum, which shows a double peak at g=1.82 and a g(x) line at g=1.76. This component has E(m) values at pH10 for single and double reduction of -15mV and approx. -150mV respectively. Both of these E(m) values are pH-dependent. The presence of an interaction between this component and the photoreduced primary acceptor indicates the close proximity of these components. However, the midpoint potential of this component indicates a function as a secondary electron-transport component rather than an electron acceptor in the reaction centre. The dependence of the bacteriopheophytin intermediate (I) doublet e.p.r. signal on the presence of the semiquinone-iron form of the primary acceptor is demonstrated. The midpoint potential of the I/I(-) couple is estimated to be lower than -600mV.  相似文献   

8.
Kobayashi K  Tagawa S  Mogi T 《Biochemistry》2000,39(50):15620-15625
To elucidate a unique mechanism for the quinol oxidation in the Escherichia coli cytochrome bo, we applied pulse radiolysis technique to the wild-type enzyme with or without a single bound ubiquinone-8 at the high-affinity quinone binding site (Q(H)), using N-methylnicotinamide (NMA) as an electron mediator. With the ubiquinone bound enzyme, the reduction of the oxidase occurred in two phases as judged from kinetic difference spectra. In the faster phase, the transient species with an absorption maximum at 440 nm, a characteristic of the formation of ubisemiquinone anion radical, appeared within 10 micros after pulse radiolysis. In the slower phase, a decrease of absorption at 440 nm was accompanied by an increase of absorption at 428 and 561 nm, characteristic of the reduced form. In contrast, with the bound ubiquinone-8-free wild-type enzyme, NMA radicals directly reduced hemes b and o, though the reduction yield was low. These results indicate that a pathway for an intramolecular electron transfer from ubisemiquinone anion radical at the Q(H) site to heme b exists in cytochrome bo. The first-order rate constant of this process was calculated to be 1.5 x 10(3) s(-1) and is comparable to a turnover rate for ubiquinol-1. The rate constant for the intramolecular electron transfer decreased considerably with increasing pH, though the yields of the formation of ubisemiquinone anion radical and the subsequent reduction of the hemes were not affected. The pH profile was tightly linked to the stability of the bound ubisemiquinone in cytochrome bo [Ingledew, W. J., Ohnishi, T., and Salerno, J. C. (1995) Eur. J. Biochem. 227, 903-908], indicating that electron transfer from the bound ubisemiquinone at the Q(H) site to the hemes slows down at the alkaline pH where the bound ubisemiquinone can be stabilized. These findings are consistent with our previous proposal that the bound ubiquinone at the Q(H) site mediates electron transfer from the low-affinity quinol oxidation site in subunit II to low-spin heme b in subunit I.  相似文献   

9.
Oxidation-reduction thermodynamic equilibria involving the quinone-acceptor complex have been examined in whole-membrane fragments from Chloroflexus aurantiacus. The primary quinone acceptor was titrated by monitoring the amount of cytochrome c554 photooxidized by a flash of light as a function of the redox potential. In contrast to previous data obtained in purified plasma membranes, in which the primary quinone acceptor exhibited a midpoint potential equal to -50 mV at pH 8.2, in whole-membrane fragments it titrated at -210 mV (pH 8.0), with a pH dependence of -60 mV/pH up to a pK value of 9.3. o-Phenanthroline, an inhibitor of electron transfer from the primary to the secondary quinone acceptor, shifted the Em/pH curve of the primary acceptor to higher redox potentials. The midpoint potential of the secondary quinone acceptor and its dependence on pH has been determined by comparing the kinetics of the charge recombination processes within the reaction center complex in the presence and in the absence of o-phenanthroline. It is concluded that both the primary and the secondary quinone acceptors interact with a proton, with pK values of 9.3 and of approximately 10.2 respectively. At physiological pH the electron appears to be stabilized on the secondary with respect to the primary quinone acceptor by approximately 60 meV.  相似文献   

10.
The putative oxidation of ubiquinol by the cytochrome bo3 terminal oxidase of Escherichia coli in sequential one-electron steps requires stabilization of the semiquinone. ENDOR spectroscopy has recently been used to study the native ubisemiquinone radical formed in the cytochrome bo3 quinone-binding site [Veselov, A.V., Osborne, J.P., Gennis, R.B. & Scholes, C.P. (2000) Biochemistry 39, 3169-3175]. Comparison of these spectra with those from the decyl-ubisemiquinone radical in vitro indicated that the protein induced large changes in the electronic structure of the ubisemiquinone radical. We have used quinone-substitution experiments to obtain ENDOR spectra of ubisemiquinone, phyllosemiquinone and plastosemiquinone anion radicals bound at the cytochrome bo3 quinone-binding site. Large changes in the electronic structures of these semiquinone anion radicals are induced on binding to the cytochrome bo3 oxidase. The changes in electronic structure are, however, independent of the electronic structures of these semiquinones in vitro. Thus it is shown to be the structure of this binding site in the protein, not the covalent structure of the bound quinone, that determines the electronic structure of the protein-bound semiquinone.  相似文献   

11.
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   

12.
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   

13.
Electron transfer between purified reaction centers from Rhodopseudomonas sphaeroides and exogenous ubiquinone has been studied in the presence of electron donors by measurements of light-induced absorbance changes following a sequence of short actinic light flashes. Each odd flash promotes the formation of a molecule of ubisemiquinone; after each even flash the semiquinone disappears and a molecule of the fully reduced quinone appears. We interpret these results by means of a model where a specialized molecule of ubiquinone is reduced by the primary electron acceptor in a one-electron transfer reaction after each flash, and is reoxidized by a molecule of the ubiquinone pool in a two-electron transfer reaction every two flashes.  相似文献   

14.
Respiratory processes often use quinone oxidoreduction to generate a transmembrane proton gradient, making the 2H+/2e quinone chemistry important for ATP synthesis. There are a variety of quinones used as electron carriers between bioenergetic proteins, and some respiratory proteins can functionally interact with more than one quinone type. In the case of complex II homologs, which couple quinone chemistry to the interconversion of succinate and fumarate, the redox potentials of the biologically available ubiquinone and menaquinone aid in driving the chemical reaction in one direction. In the complex II homolog quinol:fumarate reductase, it has been demonstrated that menaquinol oxidation requires at least one proton shuttle, but many of the remaining mechanistic details of menaquinol oxidation are not fully understood, and little is known about ubiquinone reduction. In the current study, structural and computational studies suggest that the sequential removal of the two menaquinol protons may be accompanied by a rotation of the naphthoquinone ring to optimize the interaction with a second proton shuttling pathway. However, kinetic measurements of site-specific mutations of quinol:fumarate reductase variants show that ubiquinone reduction does not use the same pathway. Computational docking of ubiquinone followed by mutagenesis instead suggested redundant proton shuttles lining the ubiquinone-binding site or from direct transfer from solvent. These data show that the quinone-binding site provides an environment that allows multiple amino acid residues to participate in quinone oxidoreduction. This suggests that the quinone-binding site in complex II is inherently plastic and can robustly interact with different types of quinones.  相似文献   

15.
Cytochrome bo3 is the major respiratory oxidase located in the cytoplasmic membrane of Escherichia coli when grown under high oxygen tension. The enzyme catalyzes the 2-electron oxidation of ubiquinol-8 and the 4-electron reduction of dioxygen to water. When solubilized and isolated using dodecylmaltoside, the enzyme contains one equivalent of ubiquinone-8, bound at a high affinity site (QH). The quinone bound at the QH site can form a stable semiquinone, and the amino acid residues which hydrogen bond to the semiquinone have been identified. In the current work, it is shown that the tightly bound ubiquinone-8 at the QH site is not displaced by ubiquinol-1 even during enzyme turnover. Furthermore, the presence of high affinity inhibitors, HQNO and aurachin C1–10, does not displace ubiquinone-8 from the QH site. The data clearly support the existence of a second binding site for ubiquinone, the QL site, which can rapidly exchange with the substrate pool. HQNO is shown to bind to a single site on the enzyme and to prevent formation of the stable ubisemiquinone, though without displacing the bound quinone. Inhibition of the steady state kinetics of the enzyme indicates that aurachin C1–10 may compete for binding with quinol at the QL site while, at the same time, preventing formation of the ubisemiquinone at the QH site. It is suggested that the two quinone binding sites may be adjacent to each other or partially overlap.  相似文献   

16.
Hellwig P  Yano T  Ohnishi T  Gennis RB 《Biochemistry》2002,41(34):10675-10679
During turnover of cytochrome bo(3) from Escherichia coli, a semiquinone radical is stabilized in a high-affinity binding site. To identify binding partners of this radical, site-directed mutants have been designed on the basis of a recently modeled quinone binding site (Abramson et al., 2000). The R71H, H98F, D75H, and I102W mutant enzymes were found to show very little or no quinol oxidase activity. The thermodynamic and EPR spectroscopic properties of semiquinone radicals in these mutants were characterized. For the H98F and the R71H mutants, no EPR signal of the semiquinone radical was observed in the redox potential range from -100 to 250 mV. During potentiometric titration of the D75H mutant enzyme, a semiquinone signal was detected in the same potential range as that of the wild-type enzyme. However, the EPR spectrum of the D75H mutant lacks the characteristic hyperfine structure of the semiquinone radical signal observed in the wild-type oxidase, indicating that D75 or the introduced His, interacts with the semiquinone radical. For the I102W mutant, a free radical signal was observed with a redox midpoint potential downshifted by about 200 mV. On the basis of these observations, it is suggested that R71, D75, and H98 residues are involved in the stabilization of the semiquinone state in the high-affinity binding site. Details of the possible binding motif and mechanistic implications are discussed.  相似文献   

17.
The effects of pH and inhibitors on the spectra and redox properties of the haems b of the bc1 complex of beef heart submitochondrial particles were investigated. The major findings were: (1) both haems have a weakly redox-linked protonatable group with pKox and pKred of around 6 and 8; (2) at pH values above 7, haem bH becomes heterogeneous in its redox behaviour. This heterogeneity is removed by the Qi site inhibitors antimycin A, funiculosin and HQNO, but not by the Qo site inhibitors myxothiazol or stigmatellin; (3) of all inhibitors tested only funiculosin had a large effect on the Em/pH profile of either haem b. In all cases where definite effects were found, the haem most affected was that thought to be closest to the site of inhibitor binding; (4) spectral shifts of haem groups caused by inhibitor binding were usually, but not always, of the haem group closest to the binding site; (5) titrations with succinate/fumarate were in reasonable agreement with redox-mediated data provided that strict anaerobiosis was maintained. Apparent large shifts of haem midpoint potentials with antimycin A and myxothiazol could be produced in aerobic succinate/fumarate titrations in the presence of cyanide, as already reported in the literature, but these were artefactual; (6) the heterogeneous haem bH titration behaviour can be simulated with a model similar to that proposed by Salerno et al. (J. Biol. Chem. (1989) 264, 15398-15403) in which there is redox interaction between haem bH and ubiquinone species bound at the Qi site. Simulations closely fit both the haem bH data and known semiquinone data only if it is assumed that semiquinone bound to oxidised haem bH is EPR-silent.  相似文献   

18.
The mechanism of ubiquinone homologs reduction by different preparations of mitochondrial NADH dehydrogenase: complex I within submitochondrial particles, isolated NADH-ubiquinone oxidoreductase and soluble low molecular weight NADH dehydrogenase, has been investigated. It has been shown that NADH oxidation via the rotenone-insensitive reaction is associated with one-electron reduction of low molecular weight ubiquinone homologs (Q0, Q1, Q2) to semiquinone with subsequent fast oxidation of the latter by atmospheric oxygen to form a superoxide radical. The two-electron ubiquinone reduction to quinol in the rotenone-sensitive reaction is unaccompanied by the semiquinone release from the enzyme active center into the surrounding solution.  相似文献   

19.
Following addition of myxothiazol to antimycin-treated chromatophores from Rhodobacter sphaeroides poised at an ambient redox potential (E(h)) of approximately 300 mV, the amplitude of the flash-induced cytochrome c(1) oxidation in the ms range increased, indicating a decrease in the availability of electrons from the immediate donor to c(1), the Rieske iron-sulfur protein (ISP). Because the effect was seen only over the limited E(h) range, we conclude that it is due to a decrease in the apparent midpoint redox potential (E(m)) of the ISP by about 40 mV on addition of myxothiazol. This is in line with the change in E(m) previously seen in direct redox titrations. Our results show that the reduced ISP binds with quinone at the Q(o) site with a higher affinity than does the oxidized ISP. The displacement of ubiquinone by myxothiazol leads to elimination of this preferential binding of the ISP reduced form and results in a shift in the midpoint potential of ISP to a more negative value. A simple hypothesis to explain this effect is that myxothiazol prevents formation of hydrogen bond of ubiquinone with the reduced ISP. We conclude that all Q(o) site occupants (ubiquinone, UHDBT, stigmatellin) that form hydrogen bonds with the reduced ISP shift the apparent E(m) of the ISP in the same direction to more positive values. Inhibitors that bind in the domain of the Q(o) site proximal to heme b(L) (myxothiazol, MOA-stilbene) and displace ubiquinone from the site cause a decrease in E(m) of ISP. We present a new formalism for treatment of the relation between E(m) change and the binding constants involved, which simplifies analysis. Using this formalism, we estimated that binding free energies for hydrogen bond formation with the Q(o) site occupant, range from the largest value of approximately 23 kJ mol(-1) in the presence of stigmatellin (appropriate for the buried hydrogen bond shown by structures), to a value of approximately 3.5 kJ mol(-1) in the native complex. We discuss this range of values in the context of a model in which the native structure constrains the interaction of ISP with the Q(o) site occupant so as to favor dissociation and the faster kinetics of unbinding necessary for rapid turnover.  相似文献   

20.
To determine the effect of the redox state of the Rieske protein on ligand binding to the quinol oxidation site of the bc(1) complex, we measured the binding rate constants (k(1)) for stigmatellin and myxothiazol, at different concentrations of decylbenzoquinone or decylbenzoquinol, in the bovine bc(1) complex with the Rieske protein in the oxidized or reduced state. Stigmatellin and myxothiazol bound tightly and competitively with respect to quinone or quinol, independently of the redox state of the Rieske protein. In the oxidized bc(1) complex, the k(1) values for stigmatellin ( approximately 2.6 x 10(6) m(-1)s(-1)) and myxothiazol ( approximately 8 x 10(5) m(-1)s(-1)), and the dissociation constant (K(d)) for quinone, were similar between pH 6.5 and 9, indicating that ligand binding is independent of the protonation state of histidine 161 of the Rieske protein (pK(a) approximately 7.6). Reduction of the Rieske protein increased the k(1) value for stigmatellin and decreased the K(d) value for quinone by 50%, without modifying the k(1) for myxothiazol. These results indicate that reduction of the Rieske protein and protonation of histidine 161 do not induce a strong stabilization of ligand binding to the quinol oxidation site, as assumed in models that propose the existence of a highly stabilized semiquinone as a reaction intermediate during quinol oxidation.  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