共查询到20条相似文献,搜索用时 15 毫秒
1.
Brandt U 《Journal of bioenergetics and biomembranes》1999,31(3):243-250
The unique bifurcated oxidation of ubiquinol at center P (QO) ofthe cytochrome bc
1 complexis the reaction within the Q-cycle reaction scheme that is most critical for the link betweenelectron transfer and vectorial proton translocation. While there is a general consensus aboutthe overall reaction at center P, the nature of the intermediates and the way the reaction iscontrolled to ensure obligatory bifurcation is still controversial. By reducing the reaction toits essential steps, a kinetic net rate model is developed in which the activation barrier isassociated with the deprotonation of ubiquinol, but the steady state rate is kinetically controlledby the occupancy of the ubiquinol anion and the semiquinone state. This concept is used tointerpret experimental data and is discussed in terms of various mechanistic models that areunder discussion. It is outlined how other aspects of the center P mechanism like the proposedprosthetic ubiquinone and the moving domain of the Rieske protein could be incorporatedin the kinetic framework. 相似文献
2.
Hagai Rottenberg Raul Covian Bernard L. Trumpower 《The Journal of biological chemistry》2009,284(29):19203-19210
The mitochondrial cytochrome bc1 complex (ubiquinol/cytochrome c oxidoreductase) is generally thought to generate superoxide anion that participates in cell signaling and contributes to cellular damage in aging and degenerative disease. However, the isolated, detergent-solubilized bc1 complex does not generate measurable amounts of superoxide except when inhibited by antimycin. In addition, indirect measurements of superoxide production by cells and isolated mitochondria have not clearly resolved the contribution of the bc1 complex to the generation of superoxide by mitochondria in vivo, nor did they establish the effect, if any, of membrane potential on superoxide formation by this enzyme complex. In this study we show that the yeast cytochrome bc1 complex does generate significant amounts of superoxide when reconstituted into phospholipid vesicles. The rate of superoxide generation by the reconstituted bc1 complex increased exponentially with increased magnitude of the membrane potential, a finding that is compatible with the suggestion that membrane potential inhibits electron transfer from the cytochrome bL to bH hemes, thereby promoting the formation of a ubisemiquinone radical that interacts with oxygen to generate superoxide. When the membrane potential was further increased, by the addition of nigericin or by the imposition of a diffusion potential, the rate of generation of superoxide was further accelerated and approached the rate obtained with antimycin. These findings suggest that the bc1 complex may contribute significantly to superoxide generation by mitochondria in vivo, and that the rate of superoxide generation can be controlled by modulation of the mitochondrial membrane potential.The mitochondrial oxidative phosphorylation system utilizes the energy derived from the oxidation of metabolic substrates to drive the synthesis of ATP. Electron transport through the NADH dehydrogenase complex, cytochrome bc1 complex, and cytochrome c oxidase complex is coupled to proton translocation across the mitochondrial inner membrane, thus generating a protonmotive force (Δp) consisting of a membrane potential (ΔΨ) and a pH gradient (ΔpH) that drives the synthesis of ATP by the ATP synthase (reviewed in Ref. 1).Several of the mitochondrial electron transport complexes produce free radical intermediates that interact with oxygen to generate superoxide (reviewed in Refs. 2, 3). Superoxide is a highly reactive compound that can lead to the formation of other free radicals and reactive compounds and thus damage directly or indirectly cellular proteins, DNA, and phospholipids. It is also believed that free radical damage is a major cause of aging and contributes to many degenerative diseases (reviewed in Ref. 4).Studies with isolated mitochondria have attempted to evaluate the contributions of the different mitochondrial energy-transducing complexes to this process (5–9). An early study with isolated rat heart mitochondria suggested that the bc1 complex produces large amounts of superoxide, but only when the mitochondrial membrane potential is high (10). This conclusion led to the suggestion that cells modulate the magnitude of the mitochondrial protonmotive force to protect the mitochondria from excess production of superoxide (11).However, it was shown later that with high concentrations of succinate as a substrate and without rotenone (as in Ref. 10), most of the superoxide is generated by reverse electron transport through complex I (8). Moreover, the rate of generation of superoxide by reverse electron transport through complex I was shown to be more strongly dependent on ΔpH than on Δψ (12). It was also suggested that the contribution of the bc1 complex to superoxide generation by mitochondria is negligible compared with that produced by reverse electron transport through complex I (9), but it is not clear whether reverse electron transport is a significant process under most physiological conditions.Several groups have measured superoxide production by the detergent-solubilized bc1 complex isolated from either yeast or beef heart. It was possible to observe superoxide production by the isolated, detergent-solubilized bc1 complex that was mutated in key residues at the ubiquinol oxidation site (13). However, the native enzyme did not produce measurable amounts of superoxide except when inhibited by antimycin or other bc1 complex inhibitors (14–18). The mechanism of the antimycin-induced generation of superoxide by the bc1 complex is fairly well understood within the framework of the Q cycle mechanism, shown in Fig. 1. Following the oxidation of ubiquinol at center P, as electrons recycle through the b hemes, antimycin inhibits reduction of ubiquinone at center N and electrons back up in center P, resulting in the formation of a ubisemiquinone radical, which can interact with oxygen to form superoxide (15, 18). It also can be predicted that the membrane potential would inhibit electron transfer from heme bL to bH and stimulate the production of superoxide by the bc1 complex. However, it is not known whether this prediction actually manifests and, if so, how strong is the dependence of superoxide production by the bc1 complex on the magnitude of membrane potential.Open in a separate windowFIGURE 1.Mechanistic basis for production of superoxide by the reconstituted cytochrome bc1 complex. The figure shows the protonmotive Q cycle mechanism and the leak of electrons to oxygen that is presumably the source of superoxide formation by the reconstituted enzyme. A shows the protonmotive Q cycle mechanism as it normally functions. Ubiquinol (QH2) is oxidized at center P near the outer surface of the membrane or vesicle in a bifurcated reaction that transfers one electron to the Rieske iron-sulfur protein (ISP) and one electron to the bL heme of cytochrome b. The electron on the iron-sulfur protein is then transferred to cytochrome c1, and the electron on the bL heme is transferred to the bH heme, which then reduces ubiquinone (Q) to semiquinone at center N. When a second molecule of ubiquinol is oxidized, the electron that arrives on the bH heme reduces semiquinone to ubiquinol. B shows the formation of superoxide anion that results when electron transfer from the bL to bH heme is inhibited, either by an opposing membrane potential or by antimycin, which blocks reoxidation of the bH heme, causing electrons to accumulate in the bL heme. Superoxide anion is formed by reaction of oxygen with ubisemiquinone, which is formed either by transfer of one electron from ubiquinol to the iron-sulfur protein or by reduction of ubiquinone by the reduced bL heme. In both panels solid arrows indicate electron transfer reactions. Dashed arrows indicate movement of ubiquinone and ubiquinol between reaction centers in the bc1 complex, release and uptake of protons at center P and center N, or changes in redox status of ubiquinone, ubiquinol, and oxygen. Solid bars in B show the opposition of electron transfer from the bL to bH heme by the membrane potential and inhibition of bH reoxidation by antimycin.We have attempted to resolve this issue by reconstitution of the yeast cytochrome bc1 complex into phospholipid vesicles, followed by measuring the rate of superoxide generation in parallel with the magnitude of the membrane potential that is generated by the reconstituted enzyme. Our findings indicate that superoxide anion formation by the bc1 complex in situ depends strongly on membrane potential and can approach values similar to those promoted by antimycin. 相似文献
3.
Casutt MS Nedielkov R Wendelspiess S Vossler S Gerken U Murai M Miyoshi H Möller HM Steuber J 《The Journal of biological chemistry》2011,286(46):40075-40082
Na(+) is the second major coupling ion at membranes after protons, and many pathogenic bacteria use the sodium-motive force to their advantage. A prominent example is Vibrio cholerae, which relies on the Na(+)-pumping NADH:quinone oxidoreductase (Na(+)-NQR) as the first complex in its respiratory chain. The Na(+)-NQR is a multisubunit, membrane-embedded NADH dehydrogenase that oxidizes NADH and reduces quinone to quinol. Existing models describing redox-driven Na(+) translocation by the Na(+)-NQR are based on the assumption that the pump contains four flavins and one FeS cluster. Here we show that the large, peripheral NqrA subunit of the Na(+)-NQR binds one molecule of ubiquinone-8. Investigations of the dynamic interaction of NqrA with quinones by surface plasmon resonance and saturation transfer difference NMR reveal a high affinity, which is determined by the methoxy groups at the C-2 and C-3 positions of the quinone headgroup. Using photoactivatable quinone derivatives, it is demonstrated that ubiquinone-8 bound to NqrA occupies a functional site. A novel scheme of electron transfer in Na(+)-NQR is proposed that is initiated by NADH oxidation on subunit NqrF and leads to quinol formation on subunit NqrA. 相似文献
4.
Prashant K. Singh Maruf Sarwar Elena Maklashina Violetta Kotlyar Sany Rajagukguk Thomas M. Tomasiak Gary Cecchini Tina M. Iverson 《The Journal of biological chemistry》2013,288(34):24293-24301
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
5.
Yuan-Gang Qu Fei Zhou Linda Yu Chang-An Yu 《The Journal of biological chemistry》2013,288(2):1047-1054
Proton transfer involving internal water molecules that provide hydrogen bonds and facilitate proton diffusion has been identified in some membrane proteins. Arg-94 in cytochrome b of the Rhodobacter sphaeroides bc1 complex is fully conserved and is hydrogen-bonded to the heme propionate and a chain of water molecules. To further elucidate the role of Arg-94, we generated the mutations R94A, R94D, and R94N. The wild-type and mutant bc1 complexes were purified and then characterized. The results show that substitution of Arg-94 decreased electron transfer activity and proton pumping capability and increased O2˙̄ production, suggesting the importance of Arg-94 in the catalytic mechanism of the bc1 complex in R. sphaeroides. This also suggests that the transport of H+, O2, and O2˙̄ in the bc1 complex may occur by the same pathway. 相似文献
6.
Oleksandr Kokhan Vladimir P. Shinkarev Colin A. Wraight 《The Journal of biological chemistry》2010,285(29):22513-22521
We have used imidazole (Im) and N-methylimidazole (MeIm) as probes of the heme-binding cavity of membrane-bound cytochrome (cyt) c1 in detergent-solubilized bc1 complex from Rhodobacter sphaeroides. Imidazole binding to cyt c1 substantially lowers the midpoint potential of the heme and fully inhibits bc1 complex activity. Temperature dependences showed that binding of Im (Kd ≈ 330 μm, 25 °C, pH 8) is enthalpically driven (ΔH0 = −56 kJ/mol, ΔS0 = −121 J/mol/K), whereas binding of MeIm is 30 times weaker (Kd ≈ 9.3 mm) and is entropically driven (ΔH0 = 47 kJ/mol, ΔS0° = 197 J/mol/K). The large enthalpic and entropic contributions suggest significant structural and solvation changes in cyt c1 triggered by ligand binding. Comparison of these results with those obtained previously for soluble cyts c and c2 suggested that Im binding to cyt c1 is assisted by formation of hydrogen bonds within the heme cleft. This was strongly supported by molecular dynamics simulations of Im adducts of cyts c, c2, and c1, which showed hydrogen bonds formed between the NδH of Im and the cyt c1 protein, or with a water molecule sequestered with the ligand in the heme cleft. 相似文献
7.
Minoru Kubo Satoru Nakashima Satoru Yamaguchi Takashi Ogura Masao Mochizuki Jiyoung Kang Masaru Tateno Kyoko Shinzawa-Itoh Koji Kato Shinya Yoshikawa 《The Journal of biological chemistry》2013,288(42):30259-30269
X-ray structural and mutational analyses have shown that bovine heart cytochrome c oxidase (CcO) pumps protons electrostatically through a hydrogen bond network using net positive charges created upon oxidation of a heme iron (located near the hydrogen bond network) for O2 reduction. Pumping protons are transferred by mobile water molecules from the negative side of the mitochondrial inner membrane through a water channel into the hydrogen bond network. For blockage of spontaneous proton back-leak, the water channel is closed upon O2 binding to the second heme (heme a3) after complete collection of the pumping protons in the hydrogen bond network. For elucidation of the structural bases for the mechanism of the proton collection and timely closure of the water channel, conformational dynamics after photolysis of CO (an O2 analog)-bound CcO was examined using a newly developed time-resolved infrared system feasible for accurate detection of a single C=O stretch band of α-helices of CcO in H2O medium. The present results indicate that migration of CO from heme a3 to CuB in the O2 reduction site induces an intermediate state in which a bulge conformation at Ser-382 in a transmembrane helix is eliminated to open the water channel. The structural changes suggest that, using a conformational relay system, including CuB, O2, heme a3, and two helix turns extending to Ser-382, CuB induces the conformational changes of the water channel that stimulate the proton collection, and senses complete proton loading into the hydrogen bond network to trigger the timely channel closure by O2 transfer from CuB to heme a3. 相似文献
8.
泛醌-细胞色素c还原酶(QCR)是线粒体呼吸链的三个能量偶联部位之一,它起着将电子从还原型泛醌传递给细胞色素c(Cyt.c)的作用,根据King和Yu提出的泛醌结合蛋白理论[1],泛醌-细胞色素c还原酶中含有泛醌结合蛋白QPc.研究表明,泛醌-细胞色... 相似文献
9.
Elisabeth C. Lowe Sarah Bydder Robert S. Hartshorne Hannah L. U. Tape Elizabeth J. Dridge Charles M. Debieux Konrad Paszkiewicz Ian Singleton Richard J. Lewis Joanne M. Santini David J. Richardson Clive S. Butler 《The Journal of biological chemistry》2010,285(24):18433-18442
Selenate reductase (SER) from Thauera selenatis is a periplasmic enzyme that has been classified as a type II molybdoenzyme. The enzyme comprises three subunits SerABC, where SerC is an unusual b-heme cytochrome. In the present work the spectropotentiometric characterization of the SerC component and the identification of redox partners to SER are reported. The mid-point redox potential of the b-heme was determined by optical titration (Em + 234 ± 10 mV). A profile of periplasmic c-type cytochromes expressed in T. selenatis under selenate respiring conditions was undertaken. Two c-type cytochromes were purified (∼24 and ∼6 kDa), and the 24-kDa protein (cytc-Ts4) was shown to donate electrons to SerABC in vitro. Protein sequence of cytc-Ts4 was obtained by N-terminal sequencing and liquid chromatography-tandem mass spectrometry analysis, and based upon sequence similarities, was assigned as a member of cytochrome c4 family. Redox potentiometry, combined with UV-visible spectroscopy, showed that cytc-Ts4 is a diheme cytochrome with a redox potential of +282 ± 10 mV, and both hemes are predicted to have His-Met ligation. To identify the membrane-bound electron donors to cytc-Ts4, growth of T. selenatis in the presence of respiratory inhibitors was monitored. The specific quinol-cytochrome c oxidoreductase (QCR) inhibitors myxothiazol and antimycin A partially inhibited selenate respiration, demonstrating that some electron flux is via the QCR. Electron transfer via a QCR and a diheme cytochrome c4 is a novel route for a member of the DMSO reductase family of molybdoenzymes. 相似文献
10.
Eiko Nakamaru-Ogiso Mou-Chieh Kao Han Chen Subhash C. Sinha Takao Yagi Tomoko Ohnishi 《The Journal of biological chemistry》2010,285(50):39070-39078
Complex I pumps protons across the membrane by using downhill redox energy. Here, to investigate the proton pumping mechanism by complex I, we focused on the largest transmembrane subunit NuoL (Escherichia coli ND5 homolog). NuoL/ND5 is believed to have H+ translocation site(s), because of a high sequence similarity to multi-subunit Na+/H+ antiporters. We mutated thirteen highly conserved residues between NuoL/ND5 and MrpA of Na+/H+ antiporters in the chromosomal nuoL gene. The dNADH oxidase activities in mutant membranes were mostly at the control level or modestly reduced, except mutants of Glu-144, Lys-229, and Lys-399. In contrast, the peripheral dNADH-K3Fe(CN)6 reductase activities basically remained unchanged in all the NuoL mutants, suggesting that the peripheral arm of complex I was not affected by point mutations in NuoL. The proton pumping efficiency (the ratio of H+/e−), however, was decreased in most NuoL mutants by 30–50%, while the IC50 values for asimicin (a potent complex I inhibitor) remained unchanged. This suggests that the H+/e− stoichiometry has changed from 4H+/2e− to 3H+ or 2H+/2e− without affecting the direct coupling site. Furthermore, 50 μm of 5-(N-ethyl-N-isopropyl)-amiloride (EIPA), a specific inhibitor for Na+/H+ antiporters, caused a 38 ± 5% decrease in the initial H+ pump activity in the wild type, while no change was observed in D178N, D303A, and D400A mutants where the H+ pumping efficiency had already been significantly decreased. The electron transfer activities were basically unaffected by EIPA in both control and mutants. Taken together, our data strongly indicate that the NuoL subunit is involved in the indirect coupling mechanism. 相似文献
11.
Armen Y. Mulkidjanian 《BBA》2005,1709(1):5-34
This review is focused on the mechanism of ubiquinol oxidation by the cytochrome bc1 complex (bc1). This integral membrane complex serves as a “hub” in the vast majority of electron transfer chains. The bc1 oxidizes a ubiquinol molecule to ubiquinone by a unique “bifurcated” reaction where the two released electrons go to different acceptors: one is accepted by the mobile redox active domain of the [2Fe-2S] iron-sulfur Rieske protein (FeS protein) and the other goes to cytochrome b. The nature of intermediates in this reaction remains unclear. It is also debatable how the enzyme prevents short-circuiting that could happen if both electrons escape to the FeS protein. Here, I consider a reaction mechanism that (i) agrees with the available experimental data, (ii) entails three traits preventing the short-circuiting in bc1, and (iii) exploits the evident structural similarity of the ubiquinone binding sites in the bc1 and the bacterial photosynthetic reaction center (RC). Based on the latter congruence, it is suggested that the reaction route of ubiquinol oxidation by bc1 is a reversal of that leading to the ubiquinol formation in the RC. The rate-limiting step of ubiquinol oxidation is then the re-location of a ubiquinol molecule from its stand-by site within cytochrome b into a catalytic site, which is formed only transiently, after docking of the mobile redox domain of the FeS protein to cytochrome b. In the catalytic site, the quinone ring is stabilized by Glu-272 of cytochrome b and His-161 of the FeS protein. The short circuiting is prevented as long as: (i) the formed semiquinone anion remains bound to the reduced FeS domain and impedes its undocking, so that the second electron is forced to go to cytochrome b; (ii) even after ubiquinol is fully oxidized, the reduced FeS domain remains docked to cytochrome b until electron(s) pass through cytochrome b; (iii) if cytochrome b becomes (over)reduced, the binding and oxidation of further ubiquinol molecules is hampered; the reason is that the Glu-272 residue is turned towards the reduced hemes of cytochrome b and is protonated to stabilize the surplus negative charge; in this state, this residue cannot participate in the binding/stabilization of a ubiquinol molecule. 相似文献
12.
Patrick M. McTernan Sanjeev K. Chandrayan Chang-Hao Wu Brian J. Vaccaro W. Andrew Lancaster Qingyuan Yang Dax Fu Greg L. Hura John A. Tainer Michael W. W. Adams 《The Journal of biological chemistry》2014,289(28):19364-19372
The archaeon Pyrococcus furiosus grows optimally at 100 °C by converting carbohydrates to acetate, CO2, and H2, obtaining energy from a respiratory membrane-bound hydrogenase (MBH). This conserves energy by coupling H2 production to oxidation of reduced ferredoxin with generation of a sodium ion gradient. MBH is encoded by a 14-gene operon with both hydrogenase and Na+/H+ antiporter modules. Herein a His-tagged MBH was expressed in P. furiosus and the detergent-solubilized complex purified under anaerobic conditions by affinity chromatography. Purified MBH contains all 14 subunits by electrophoretic analysis (13 subunits were also identified by mass spectrometry) and had a measured iron:nickel ratio of 15:1, resembling the predicted value of 13:1. The as-purified enzyme exhibited a rhombic EPR signal characteristic of the ready nickel-boron state. The purified and membrane-bound forms of MBH both preferentially evolved H2 with the physiological donor (reduced ferredoxin) as well as with standard dyes. The O2 sensitivities of the two forms were similar (half-lives of ∼15 h in air), but the purified enzyme was more thermolabile (half-lives at 90 °C of 1 and 25 h, respectively). Structural analysis of purified MBH by small angle x-ray scattering indicated a Z-shaped structure with a mass of 310 kDa, resembling the predicted value (298 kDa). The angle x-ray scattering analyses reinforce and extend the conserved sequence relationships of group 4 enzymes and complex I (NADH quinone oxidoreductase). This is the first report on the properties of a solubilized form of an intact respiratory MBH complex that is proposed to evolve H2 and pump Na+ ions. 相似文献
13.
Oleksandr Kokhan Vladimir P. Shinkarev Colin A. Wraight 《The Journal of biological chemistry》2010,285(29):22522-22531
The kinetics of imidazole (Im) and N-methylimidazole (MeIm) binding to oxidized cytochrome (cyt) c1 of detergent-solubilized bc1 complex from Rhodobacter sphaeroides are described. The rate of formation of the cyt c1-Im complex exhibited three separated regions of dependence on the concentration of imidazole: (i) below 8 mm Im, the rate increased with concentration in a parabolic manner; (ii) above 20 mm, the rate leveled off, indicating a rate-limiting conformational step with lifetime ∼1 s; and (iii) at Im concentrations above 100 mm, the rate substantially increased again, also parabolically. In contrast, binding of MeIm followed a simple hyperbolic concentration dependence. The temperature dependences of the binding and release kinetics of Im and MeIm were also measured and revealed very large activation parameters for all reactions. The complex concentration dependence of the Im binding rate is not consistent with the popular model for soluble c-type cytochromes in which exogenous ligand binding is preceded by spontaneous opening of the heme cleft, which becomes rate-limiting at high ligand concentrations. Instead, binding of ligand to the heme is explained by a model in which an initial and superficial binding facilitates access to the heme by disruption of hydrogen-bonded structures in the heme domain. For imidazole, two separate pathways of heme access are indicated by the distinct kinetics at low and high concentration. The structural basis for ligand entry to the heme cleft is discussed. 相似文献
14.
Sangmoon Lhee Derrick R. J. Kolling Satish K. Nair Sergei A. Dikanov Antony R. Crofts 《The Journal of biological chemistry》2010,285(12):9233-9248
The rate-determining step in the overall turnover of the bc1 complex is electron transfer from ubiquinol to the Rieske iron-sulfur protein (ISP) at the Qo-site. Structures of the ISP from Rhodobacter sphaeroides show that serine 154 and tyrosine 156 form H-bonds to S-1 of the [2Fe-2S] cluster and to the sulfur atom of the cysteine liganding Fe-1 of the cluster, respectively. These are responsible in part for the high potential (Em,7 ∼300 mV) and low pKa (7.6) of the ISP, which determine the overall reaction rate of the bc1 complex. We have made site-directed mutations at these residues, measured thermodynamic properties using protein film voltammetry to evaluate the Em and pKa values of ISPs, explored the local proton environment through two-dimensional electron spin echo envelope modulation, and characterized function in strains S154T, S154C, S154A, Y156F, and Y156W. Alterations in reaction rate were investigated under conditions in which concentration of one substrate (ubiquinol or ISPox) was saturating and the other was varied, allowing calculation of kinetic terms and relative affinities. These studies confirm that H-bonds to the cluster or its ligands are important determinants of the electrochemical characteristics of the ISP, likely through electron affinity of the interacting atom and the geometry of the H-bonding neighborhood. The calculated parameters were used in a detailed Marcus-Brønsted analysis of the dependence of rate on driving force and pH. The proton-first-then-electron model proposed accounts naturally for the effects of mutation on the overall reaction. 相似文献
15.
Yip CY Harbour ME Jayawardena K Fearnley IM Sazanov LA 《The Journal of biological chemistry》2011,286(7):5023-5033
Modern α-proteobacteria are thought to be closely related to the ancient symbiont of eukaryotes, an ancestor of mitochondria. Respiratory complex I from α-proteobacteria and mitochondria is well conserved at the level of the 14 "core" subunits, consistent with that notion. Mitochondrial complex I contains the core subunits, present in all species, and up to 31 "supernumerary" subunits, generally thought to have originated only within eukaryotic lineages. However, the full protein composition of an α-proteobacterial complex I has not been established previously. Here, we report the first purification and characterization of complex I from the α-proteobacterium Paracoccus denitrificans. Single particle electron microscopy shows that the complex has a well defined L-shape. Unexpectedly, in addition to the 14 core subunits, the enzyme also contains homologues of three supernumerary mitochondrial subunits as follows: B17.2, AQDQ/18, and 13 kDa (bovine nomenclature). This finding suggests that evolution of complex I via addition of supernumerary or "accessory" subunits started before the original endosymbiotic event that led to the creation of the eukaryotic cell. It also provides further confirmation that α-proteobacteria are the closest extant relatives of mitochondria. 相似文献
16.
Jingfeng Chen Chwen-Lih Chen Sharad Rawale Chun-An Chen Jay L. Zweier Pravin T. P. Kaumaya Yeong-Renn Chen 《The Journal of biological chemistry》2010,285(5):3168-3180
Complex I (NQR) is a critical site of superoxide () production and the major host of redox protein thiols in mitochondria. In response to oxidative stress, NQR-derived protein thiols at the 51- and 75-kDa subunits are known to be reversibly S-glutathionylated. Although several glutathionylated domains from NQR 51 and 75 kDa have been identified, their roles in the regulatory functions remain to be explored. To gain further insights into protein S-glutathionylation of complex I, we used two peptides of S-glutathionylated domain (200GAGAYICGEETALIESIEGK219 of 51-kDa protein and 361VDSDTLCTEEVFPTAGAGTDLR382 of 75-kDa protein) as chimeric epitopes incorporating a “promiscuous” T-cell epitope to generate two polyclonal antibodies, AbGSCA206 and AbGSCB367. Binding of AbGSCA206 and AbGSCB367 inhibited NQR-mediated generation by 37 and 57%, as measured by EPR spin-trapping. To further provide an appropriate control, two peptides of non-glutathionylated domain (21SGDTTAPKKTSFGSLKDFDR40 of 51-kDa peptide and 100WNILTNSEKTKKAREGVMEFL120 of 75-kDa peptide) were synthesized as chimeric epitopes to generate two polyclonal antibodies, Ab51 and Ab75. Binding of A51 did not affect NQR-mediated generation to a significant level. However, binding of Ab75 inhibited NQR-mediated generation by 35%. None of AbGSCA206, AbGSCB367, Ab51, or Ab75 showed an inhibitory effect on the electron transfer activity of NQR, suggesting that antibody binding to the glutathione-binding domain decreased electron leakage from the hydrophilic domain of NQR. When heart tissue homogenates were immunoprecipitated with Ab51 or Ab75 and probed with an antibody against glutathione, protein S-glutathionylation was enhanced in post-ischemic myocardium at the NQR 51-kDa subunit, but not at the 75-kDa subunit, indicating that the 51-kDa subunit of flavin subcomplex is more sensitive to oxidative stress resulting from myocardial infarction. 相似文献
17.
We have measured the rates of superoxide anion generation by cytochrome bc1 complexes isolated from bovine heart and yeast mitochondria and by cytochrome bc1 complexes from yeast mutants in which the midpoint potentials of the cytochrome b hemes and the Rieske iron-sulfur cluster were altered by mutations in those proteins. With all of the bc1 complexes the rate of superoxide anion production was greatest in the absence of bc1 inhibitor and ranged from 3% to 5% of the rate of cytochrome c reduction. Stigmatellin, an inhibitor that binds to the ubiquinol oxidation site in the bc1 complex, eliminated superoxide anion formation, while myxothiazol, another inhibitor of ubiquinol oxidation, allowed superoxide anion formation at a low rate. Antimycin, an inhibitor that binds to the ubiquinone reduction site in the bc1 complex, also allowed superoxide anion formation and at a slightly greater rate than myxothiazol. Changes in the midpoint potentials of the cytochrome b hemes had no significant effect on the rate of cytochrome c reduction and only a small effect on the rate of superoxide anion formation. A mutation in the Rieske iron-sulfur protein that lowers its midpoint potential from +285 to +220 mV caused the rate of superoxide anion to decline in parallel with a decline in cytochrome c reductase activity. These results indicate that superoxide anion is formed by similar mechanisms in mammalian and yeast bc1 complexes. The results also show that changes in the midpoint potentials of the redox components that accept electrons during ubiquinol oxidation have only small effects on the formation of superoxide anion, except to the extent that they affect the activity of the enzyme. 相似文献
18.
Szeto SS Reinke SN Oyedotun KS Sykes BD Lemire BD 《The Journal of biological chemistry》2012,287(27):22509-22520
Succinate dehydrogenase (SDH), also known as complex II, is required for respiratory growth; it couples the oxidation of succinate to the reduction of ubiquinone. The enzyme is composed of two domains. A membrane-extrinsic catalytic domain composed of the Sdh1p and Sdh2p subunits harbors the flavin and iron-sulfur cluster cofactors. A membrane-intrinsic domain composed of the Sdh3p and Sdh4p subunits interacts with ubiquinone and may coordinate a b-type heme. In many organisms, including Saccharomyces cerevisiae, possible alternative SDH subunits have been identified in the genome. S. cerevisiae contains one paralog of the Sdh3p subunit, Shh3p (YMR118c), and two paralogs of the Sdh4p subunit, Shh4p (YLR164w) and Tim18p (YOR297c). We cloned and expressed these alternative subunits. Shh3p and Shh4p were able to complement Δsdh3 and Δsdh4 deletion mutants, respectively, and support respiratory growth. Tim18p was unable to do so. Microarray and proteomics data indicate that the paralogs are expressed under respiratory and other more restrictive growth conditions. Strains expressing hybrid SDH enzymes have distinct metabolic profiles that we distinguished by (1)H NMR analysis of metabolites. Surprisingly, the Sdh3p subunit can form SDH isoenzymes with Sdh4p or with Shh4p as well as be a subunit of the TIM22 mitochondrial protein import complex. 相似文献
19.
Nilanjan Pal Chowdhury Amr M. Mowafy Julius K. Demmer Vikrant Upadhyay Sebastian Koelzer Elamparithi Jayamani Joerg Kahnt Marco Hornung Ulrike Demmer Ulrich Ermler Wolfgang Buckel 《The Journal of biological chemistry》2014,289(8):5145-5157
Electron bifurcation is a fundamental strategy of energy coupling originally discovered in the Q-cycle of many organisms. Recently a flavin-based electron bifurcation has been detected in anaerobes, first in clostridia and later in acetogens and methanogens. It enables anaerobic bacteria and archaea to reduce the low-potential [4Fe-4S] clusters of ferredoxin, which increases the efficiency of the substrate level and electron transport phosphorylations. Here we characterize the bifurcating electron transferring flavoprotein (EtfAf) and butyryl-CoA dehydrogenase (BcdAf) of Acidaminococcus fermentans, which couple the exergonic reduction of crotonyl-CoA to butyryl-CoA to the endergonic reduction of ferredoxin both with NADH. EtfAf contains one FAD (α-FAD) in subunit α and a second FAD (β-FAD) in subunit β. The distance between the two isoalloxazine rings is 18 Å. The EtfAf-NAD+ complex structure revealed β-FAD as acceptor of the hydride of NADH. The formed β-FADH− is considered as the bifurcating electron donor. As a result of a domain movement, α-FAD is able to approach β-FADH− by about 4 Å and to take up one electron yielding a stable anionic semiquinone, α-FAD⨪, which donates this electron further to Dh-FAD of BcdAf after a second domain movement. The remaining non-stabilized neutral semiquinone, β-FADH•, immediately reduces ferredoxin. Repetition of this process affords a second reduced ferredoxin and Dh-FADH− that converts crotonyl-CoA to butyryl-CoA. 相似文献
20.
There are now four structures of vertebrate mitochondrial bc
1 complexes available in theprotein databases and structures from yeast and bacterial sources are expected soon. Thisreview summarizes the new information with emphasis on the avian cytochrome bc
1 complex(PDB entries 1BCC and 3BCC). The Rieske iron–sulfur protein is mobile and this has beenproposed to be important for catalysis. The binding sites for quinone have been located basedon structures containing inhibitors and, in the case of the quinone reduction site Qi, thequinone itself. 相似文献