ubiquinone has been researched along with quinone* in 57 studies
4 review(s) available for ubiquinone and quinone
Article | Year |
---|---|
The production of coenzyme Q10 in microorganisms.
Coenzyme Q10 has emerged as a valuable molecule for pharmaceutical and cosmetic applications. Therefore, research into producing and optimizing coenzyme Q10 via microbial fermentation is ongoing. There are two major paths being explored for maximizing production of this molecule to commercially advantageous levels. The first entails using microbes that naturally produce coenzyme Q10 as fermentation biocatalysts and optimizing the fermentation parameters in order to reach industrial levels of production. However, the natural coenzyme Q10-producing microbes tend to be intractable for industrial fermentation settings. The second path to coenzyme Q10 production being explored is to engineer Escherichia coli with the ability to biosynthesize this molecule in order to take advantage of its more favourable fermentation characteristics and the well-understood array of genetic tools available for this bacteria. Although many studies have attempted to over-produce coenzyme Q10 in E. coli through genetic engineering, production titres still remain below those of the natural coenzyme Q10-producing microorganisms. Current research is providing the knowledge needed to alleviate the bottlenecks involved in producing coenzyme Q10 from an E. coli strain platform and the fermentation parameters that could dramatically increase production titres from natural microbial producers. Synthesizing the lessons learned from both approaches may be the key towards a more cost-effective coenzyme Q10 industry. Topics: Agrobacterium tumefaciens; Benzoquinones; Escherichia coli; Fermentation; Genetic Engineering; Metabolic Networks and Pathways; Prokaryotic Cells; Ubiquinone | 2012 |
Structure and function of quinone binding membrane proteins.
Topics: Amino Acid Sequence; Animals; Benzoquinones; Binding Sites; Cattle; Cell Membrane; Electron Transport Complex III; Electrons; Mitochondria; Models, Biological; Models, Molecular; Molecular Sequence Data; Mutation; Myocardium; Oxidation-Reduction; Protein Binding; Protons; Ubiquinone | 2003 |
Interactions of quinone with the iron-sulfur protein of the bc(1) complex: is the mechanism spring-loaded?
Since available structures of native bc(1) complexes show a vacant Q(o)-site, occupancy by substrate and product must be investigated by kinetic and spectroscopic approaches. In this brief review, we discuss recent advances using these approaches that throw new light on the mechanism. The rate-limiting reaction is the first electron transfer after formation of the enzyme-substrate complex at the Q(o)-site. This is formed by binding of both ubiquinol (QH(2)) and the dissociated oxidized iron-sulfur protein (ISP(ox)). A binding constant of approximately 14 can be estimated from the displacement of E(m) or pK for quinone or ISP(ox), respectively. The binding likely involves a hydrogen bond, through which a proton-coupled electron transfer occurs. An enzyme-product complex is also formed at the Q(o)-site, in which ubiquinone (Q) hydrogen bonds with the reduced ISP (ISPH). The complex has been characterized in ESEEM experiments, which detect a histidine ligand, likely His-161 of ISP (in mitochondrial numbering), with a configuration similar to that in the complex of ISPH with stigmatellin. This special configuration is lost on binding of myxothiazol. Formation of the H-bond has been explored through the redox dependence of cytochrome c oxidation. We confirm previous reports of a decrease in E(m) of ISP on addition of myxothiazol, and show that this change can be detected kinetically. We suggest that the myxothiazol-induced change reflects loss of the interaction of ISPH with Q, and that the change in E(m) reflects a binding constant of approximately 4. We discuss previous data in the light of this new hypothesis, and suggest that the native structure might involve a less than optimal configuration that lowers the binding energy of complexes formed at the Q(o)-site so as to favor dissociation. We also discuss recent results from studies of the bypass reactions at the site, which lead to superoxide (SO) production under aerobic conditions, and provide additional information about intermediate states. Topics: Benzoquinones; Binding Sites; Electron Transport Complex III; Iron-Sulfur Proteins; Kinetics; Methacrylates; Oxidation-Reduction; Thermodynamics; Thiazoles; Ubiquinone | 2002 |
Structures of quinone-binding sites in bc complexes: functional implications.
Topics: Animals; Bacterial Proteins; Benzoquinones; Binding Sites; Chickens; Electron Transport Complex III; Electrons; Hydrogen Bonding; Models, Chemical; Protons; Ubiquinone | 1999 |
53 other study(ies) available for ubiquinone and quinone
Article | Year |
---|---|
Increased levels of circulating oxidized mitochondrial DNA contribute to chronic inflammation in metabolic syndrome, and MitoQ-based antioxidant therapy alleviates this DNA-induced inflammation.
Here, the aim was to investigate the role of circulating oxidized mitochondrial DNA (ox-mtDNA) in metabolic syndrome (MetS)-associated chronic inflammation and evaluate the effect of Mito-Quinone (MitoQ)-based antioxidant therapy on inflammation. A total of 112 MetS patients and 111 healthy control individuals (HCs) were recruited. Peripheral blood was collected, and mononuclear cells (PBMCs) were separated. In a preclinical study, MitoQ, a mitochondrial-targeted antioxidant, was administered to Sprague-Dawley (SD) rats fed a high-fat diet (HFD). In vitro, H Topics: Animals; Antioxidants; Cell-Free Nucleic Acids; DNA, Mitochondrial; Hydrogen Peroxide; Inflammation; Metabolic Syndrome; NF-kappa B; Rats; Rats, Sprague-Dawley; Toll-Like Receptor 9; Ubiquinone | 2023 |
A Comparative Study on the Reduction Modes for Quinone to Determine Ubiquinone by HPLC with Luminol Chemiluminescence Detection Based on the Redox Reaction.
Ubiquinone (UQ) is considered one of the important biologically active molecules in the human body. Ubiquinone determination in human plasma is important for the investigation of its bioavailability, and also its plasma level is considered an indicator of many illnesses. We have previously developed sensitive and selective chemiluminescence (CL) method for the determination of UQ in human plasma based on its redox cycle with dithiothreitol (DTT) and luminol. However, this method requires an additional pump to deliver DTT as a post-column reagent and has the problems of high DTT consumption and broadening of the UQ peak due to online mixing with DTT. Herein, an HPLC (high-performance liquid chromatography) system equipped with two types of online reduction systems (electrolytic flow cell or platinum catalyst-packed reduction column) that play the role of DTT was constructed to reduce reagent consumption and simplify the system. The newly proposed two methods were carefully optimized and validated, and the analytical performance for UQ determination was compared with that of the conventional DTT method. Among the tested systems, the electrolytic reduction system showed ten times higher sensitivity than the DTT method, with a limit of detection of 3.1 nM. In addition, it showed a better chromatographic performance and the best peak shape with a number of theoretical plates exceeding 6500. Consequently, it was applied to the determination of UQ in healthy human plasma, and it showed good recovery (≥97.9%) and reliable precision (≤6.8%) without any interference from plasma components. Topics: Chromatography, High Pressure Liquid; Dithiothreitol; Humans; Luminescence; Luminescent Measurements; Luminol; Oxidation-Reduction; Quinones; Ubiquinone | 2022 |
Crystallographic investigation of the ubiquinone binding site of respiratory Complex II and its inhibitors.
The quinone binding site (Q-site) of Mitochondrial Complex II (succinate-ubiquinone oxidoreductase) is the target for a number of inhibitors useful for elucidating the mechanism of the enzyme. Some of these have been developed as fungicides or pesticides, and species-specific Q-site inhibitors may be useful against human pathogens. We report structures of chicken Complex II with six different Q-site inhibitors bound, at resolutions 2.0-2.4 Å. These structures show the common interactions between the inhibitors and their binding site. In every case a carbonyl or hydroxyl oxygen of the inhibitor is H-bonded to Tyr58 in subunit SdhD and Trp173 in subunit SdhB. Two of the inhibitors H-bond Ser39 in subunit SdhC directly, while two others do so via a water molecule. There is a distinct cavity that accepts the 2-substituent of the carboxylate ring in flutolanil and related inhibitors. A hydrophobic "tail pocket" opens to receive a side-chain of intermediate-length inhibitors. Shorter inhibitors fit entirely within the main binding cleft, while the long hydrophobic side chains of ferulenol and atpenin A5 protrude out of the cleft into the bulk lipid region, as presumably does that of ubiquinone. Comparison of mitochondrial and Escherichia coli Complex II shows a rotation of the membrane-anchor subunits by 7° relative to the iron‑sulfur protein. This rotation alters the geometry of the Q-site and the H-bonding pattern of SdhB:His216 and SdhD:Asp57. This conformational difference, rather than any active-site mutation, may be responsible for the different inhibitor sensitivity of the bacterial enzyme. Topics: Amino Acid Sequence; Animals; Benzoquinones; Binding Sites; Chickens; Electron Transport Complex II; Multienzyme Complexes; Mutagenesis, Site-Directed; Quinones; Sequence Alignment; Sus scrofa; Ubiquinone | 2021 |
Engineering Isoprenoid Quinone Production in Yeast.
Isoprenoid quinones are bioactive molecules that include an isoprenoid chain and a quinone head. They are traditionally found to be involved in primary metabolism, where they act as electron transporters, but specialized isoprenoid quinones are also produced by all domains of life. Here, we report the engineering of a baker's yeast strain, Topics: Benzoquinones; Metabolic Engineering; Phosphorus-Oxygen Lyases; Saccharomyces cerevisiae; Terpenes; Ubiquinone | 2020 |
Exploring the quinone/inhibitor-binding pocket in mitochondrial respiratory complex I by chemical biology approaches.
NADH-quinone oxidoreductase (respiratory complex I) couples NADH-to-quinone electron transfer to the translocation of protons across the membrane. Even though the architecture of the quinone-access channel in the enzyme has been modeled by X-ray crystallography and cryo-EM, conflicting findings raise the question whether the models fully reflect physiologically relevant states present throughout the catalytic cycle. To gain further insights into the structural features of the binding pocket for quinone/inhibitor, we performed chemical biology experiments using bovine heart sub-mitochondrial particles. We synthesized ubiquinones that are oversized (SF-UQs) or lipid-like (PC-UQs) and are highly unlikely to enter and transit the predicted narrow channel. We found that SF-UQs and PC-UQs can be catalytically reduced by complex I, albeit only at moderate or low rates. Moreover, quinone-site inhibitors completely blocked the catalytic reduction and the membrane potential formation coupled to this reduction. Photoaffinity-labeling experiments revealed that amiloride-type inhibitors bind to the interfacial domain of multiple core subunits (49 kDa, ND1, and PSST) and the 39-kDa supernumerary subunit, although the latter does not make up the channel cavity in the current models. The binding of amilorides to the multiple target subunits was remarkably suppressed by other quinone-site inhibitors and SF-UQs. Taken together, the present results are difficult to reconcile with the current channel models. On the basis of comprehensive interpretations of the present results and of previous findings, we discuss the physiological relevance of these models. Topics: Amiloride; Animals; Benzoquinones; Binding Sites; Catalysis; Cattle; Crystallography, X-Ray; Electron Transport; Electron Transport Complex I; Kinetics; Mitochondria; Photoaffinity Labels; Quinone Reductases; Ubiquinone | 2019 |
The Role of a FAD Cofactor in the Regulation of Acetohydroxyacid Synthase by Redox Signaling Molecules.
Acetohydroxyacid synthase (AHAS) catalyzes the first step of branched-chain amino acid (BCAA) biosynthesis, a pathway essential to the lifecycle of plants and microorganisms. This enzyme is of high interest because its inhibition is at the base of the exceptional potency of herbicides and potentially a target for the discovery of new antimicrobial drugs. The enzyme has conserved attributes from its predicted ancestor, pyruvate oxidase, such as a ubiquinone-binding site and the requirement for FAD as cofactor. Here, we show that these requirements are linked to the regulation of AHAS, in relationship to its anabolic function. Using various soluble quinone derivatives ( Topics: Acetolactate Synthase; Benzoquinones; Flavin-Adenine Dinucleotide; Humans; Models, Molecular; Mycobacterium tuberculosis; Oxidation-Reduction; Saccharomyces cerevisiae; Tuberculosis; Ubiquinone | 2017 |
Coenzyme Q0 regulates NFκB/AP-1 activation and enhances Nrf2 stabilization in attenuation of LPS-induced inflammation and redox imbalance: Evidence from in vitro and in vivo studies.
Coenzyme Q (CoQ) analogs with variable number of isoprenoid units have been demonstrated as anti-inflammatory and antioxidant/pro-oxidant molecules. In this study we used CoQ0 (2,3-dimethoxy-5-methyl-1,4-benzoquinone, zero isoprenoid side-chains), a novel quinone derivative, and investigated its molecular actions against LPS-induced inflammation and redox imbalance in murine RAW264.7 macrophages and mice. In LPS-stimulated macrophages, non-cytotoxic concentrations of CoQ0 (2.5-10 μM) inhibited iNOS/COX-2 protein expressions with subsequent reductions of NO, PGE2, TNF-α and IL-1β secretions. This inhibition was reasoned by suppression of NFκB (p65) activation, and inhibition of AP-1 (c-Jun., c-Fos, ATF2) translocation. Our findings indicated that IKKα-mediated I-κB degradation and MAPK-signaling are involved in regulation of NFκB/AP-1 activation. Furthermore, CoQ0 triggered HO-1 and NQO-1 genes through increased Nrf2 nuclear translocation and Nrf2/ARE-signaling. This phenomenon was confirmed by diminished CoQ0 protective effects in Nrf2 knockdown cells, where LPS-induced NO, PGE2, TNF-α and IL-1β productions remained high. Molecular evidence revealed that CoQ0 enhanced Nrf2 steady-state level at both transcriptional and translational levels. CoQ0-induced Nrf2 activation appears to be regulated by ROS-JNK-signaling cascades, as evidenced by suppressed Nrf2 activation upon treatment with pharmacological inhibitors of ROS (N-acetylcysteine) and JNK (SP600125). Besides, oral administration of CoQ0 (5 mg/kg) suppressed LPS-induced (1 mg/kg) induction of iNOS/COX-2 and TNF-α/IL-1β through tight regulation of NFκB/Nrf2 signaling in mice liver and spleen. Our findings conclude that pharmacological actions of CoQ0 are mediated via inhibition of NFκB/AP-1 activation and induction of Nrf2/ARE-signaling. Owing to its potent anti-inflammatory and antioxidant properties, CoQ0 could be a promising candidate to treat inflammatory disorders. Topics: Animals; Benzoquinones; Cyclooxygenase 2; Gene Expression Regulation; Heme Oxygenase-1; Inflammation; Lipopolysaccharides; Macrophages; Mice; NF-E2-Related Factor 2; NF-kappa B; Nitric Oxide Synthase Type II; Oxidation-Reduction; Reactive Oxygen Species; Signal Transduction; Transcription Factor AP-1; Transcription Factor RelA; Ubiquinone | 2016 |
Mitochondrial function and lifespan of mice with controlled ubiquinone biosynthesis.
Ubiquinone (UQ) is implicated in mitochondrial electron transport, superoxide generation and as a membrane antioxidant. Here we present a mouse model in which UQ biosynthesis can be interrupted and partially restored at will. Global loss of UQ leads to gradual loss of mitochondrial function, gradual development of disease phenotypes and shortened lifespan. However, we find that UQ does not act as antioxidant in vivo and that its requirement for electron transport is much lower than anticipated, even in vital mitochondria-rich organs. In fact, severely depressed mitochondrial function due to UQ depletion in the heart does not acutely impair organ function. In addition, we demonstrate that severe disease phenotypes and shortened lifespan are reversible upon partial restoration of UQ levels and mitochondrial function. This observation strongly suggests that the irreversible degenerative phenotypes that characterize ageing are not secondarily caused by the gradual mitochondrial dysfunction that is observed in aged animals. Topics: Aging; Analysis of Variance; Animals; Benzoquinones; Gene Silencing; Immunoblotting; Integrases; Longevity; Membrane Proteins; Mice; Mice, Knockout; Mitochondria; Mitochondrial Proteins; Mixed Function Oxygenases; Molecular Structure; Myocardium; Reactive Oxygen Species; Ubiquinone | 2015 |
Crystal structure of the entire respiratory complex I.
Complex I is the first and largest enzyme of the respiratory chain and has a central role in cellular energy production through the coupling of NADH:ubiquinone electron transfer to proton translocation. It is also implicated in many common human neurodegenerative diseases. Here, we report the first crystal structure of the entire, intact complex I (from Thermus thermophilus) at 3.3 Å resolution. The structure of the 536-kDa complex comprises 16 different subunits, with a total of 64 transmembrane helices and 9 iron-sulphur clusters. The core fold of subunit Nqo8 (ND1 in humans) is, unexpectedly, similar to a half-channel of the antiporter-like subunits. Small subunits nearby form a linked second half-channel, which completes the fourth proton-translocation pathway (present in addition to the channels in three antiporter-like subunits). The quinone-binding site is unusually long, narrow and enclosed. The quinone headgroup binds at the deep end of this chamber, near iron-sulphur cluster N2. Notably, the chamber is linked to the fourth channel by a 'funnel' of charged residues. The link continues over the entire membrane domain as a flexible central axis of charged and polar residues, and probably has a leading role in the propagation of conformational changes, aided by coupling elements. The structure suggests that a unique, out-of-the-membrane quinone-reaction chamber enables the redox energy to drive concerted long-range conformational changes in the four antiporter-like domains, resulting in translocation of four protons per cycle. Topics: Benzoquinones; Cell Membrane; Crystallography, X-Ray; Electron Transport Complex I; Humans; Hydrophobic and Hydrophilic Interactions; Models, Molecular; NAD; Oxidation-Reduction; Protein Folding; Protein Subunits; Proton-Motive Force; Protons; Thermus thermophilus; Ubiquinone | 2013 |
Ubiquinol (QH(2)) functions as a negative regulator of purine nucleotide inhibition of Acanthamoeba castellanii mitochondrial uncoupling protein.
We compared the influence of different adenine and guanine nucleotides on the free fatty acid-induced uncoupling protein (UCP) activity in non-phosphorylating Acanthamoeba castellanii mitochondria when the membranous ubiquinone (Q) redox state was varied. The purine nucleotides exhibit an inhibitory effect in the following descending order: GTP>ATP>GDP>ADP≫GMP>AMP. The efficiency of guanine and adenine nucleotides to inhibit UCP-sustained uncoupling in A. castellanii mitochondria depends on the Q redox state. Inhibition by purine nucleotides can be increased with decreasing Q reduction level (thereby ubiquinol, QH₂ concentration) even with nucleoside monophosphates that are very weak inhibitors at the initial respiration. On the other hand, the inhibition can be alleviated with increasing Q reduction level (thereby QH₂ concentration). The most important finding was that ubiquinol (QH₂) but not oxidised Q functions as a negative regulator of UCP inhibition by purine nucleotides. For a given concentration of QH₂, the linoleic acid-induced GTP-inhibited H(+) leak was the same for two types of A. castellanii mitochondria that differ in the endogenous Q content. When availability of the inhibitor (GTP) or the negative inhibition modulator (QH₂) was changed, a competitive influence on the UCP activity was observed. QH₂ decreases the affinity of UCP for GTP and, vice versa, GTP decreases the affinity of UCP for QH₂. These results describe the kinetic mechanism of regulation of UCP affinity for purine nucleotides by endogenous QH₂ in the mitochondria of a unicellular eukaryote. Topics: Acanthamoeba castellanii; Adenine Nucleotides; Benzoquinones; Fatty Acids, Nonesterified; Guanine Nucleotides; Homeostasis; Ion Channels; Membrane Potentials; Mitochondria; Mitochondrial Proteins; Oxidation-Reduction; Oxygen Consumption; Purine Nucleotides; Ribonucleotides; Ubiquinone; Uncoupling Protein 1 | 2011 |
Reaction of superoxide radical with quinone molecules.
When the superoxide radical O(2)(•-) is generated on reaction of KO(2) with water in dimethyl sulfoxide, the decay of the radical is dramatically accelerated by inclusion of quinones in the reaction mix. For quinones with no or short hydrophobic tails, the radical product is a semiquinone at much lower yield, likely indicating reduction of quinone by superoxide and loss of most of the semiquinone product by disproportionation. In the presence of ubiquinone-10, a different species (I) is generated, which has the EPR spectrum of superoxide radical. However, pulsed EPR shows spin interaction with protons in fully deuterated solvent, indicating close proximity to the ubinquinone-10. We discuss the nature of species I, and possible roles in the physiological reactions through which ubisemiquinone generates superoxide by reduction of O(2) through bypass reactions in electron transfer chains. Topics: Benzoquinones; Chemistry, Physical; Dimethyl Sulfoxide; Electron Spin Resonance Spectroscopy; Electron Transport; Oxidation-Reduction; Oxygen; Protons; Solutions; Superoxides; Ubiquinone | 2011 |
Photolabile ubiquinone analogues for identification and characterization of quinone binding sites in proteins.
Quinones are essential components in most cell and organelle bioenergetic processes both for direct electron and/or proton transfer reactions but also as means to regulate various bioenergetic processes by sensing cell redox states. To understand how quinones interact with proteins, it is important to have tools for identifying and characterizing quinone binding sites. In this work three different photo-reactive azidoquinones were synthesized, two of which are novel compounds, and the methods of synthesis was improved. The reactivity of the azidoquinones was first tested with model peptides, and the adducts formed were analyzed by mass spectrometry. The added mass detected was that of the respective azidoquinone minus N(2). Subsequently, the biological activity of the three azidoquinones was assessed, using three enzyme systems of different complexity, and the ability of the compounds to inactivate the enzymes upon illumination with long wavelength UV light was investigated. The soluble flavodoxin-like protein WrbA could only use two of the azidoquinones as substrates, whereas respiratory chain Complexes I and II could utilize all three compounds as electron acceptors. Complex II, purified in detergent, was very sensitive to illumination also in the absence of azidoquinones, making the 'therapeutic window' in that enzyme rather narrow. In membrane bound Complex I, only two of the compounds inactivated the enzyme, whereas illumination in the presence of the third compound left enzyme activity essentially unchanged. Since unspecific labeling should be equally effective for all the compounds, this demonstrates that the observed inactivation is indeed caused by specific labeling. Topics: Benzoquinones; Binding Sites; Light; Models, Molecular; Photochemistry; Protein Binding; Quinones; Ubiquinone | 2010 |
Conformational control of the Q(A) to Q(B) electron transfer in bacterial reaction centers: evidence for a frozen conformational landscape below -25 degrees C.
The competition between the P(+)Q(A)(-) --> PQ(A) charge recombination (P, bacteriochlorophyll pair acting as primary photochemical electron donor) and the electron transfer to the secondary quinone acceptor Q(A)(-)Q(B) --> Q(A)Q(B)(-) (Q(A) and Q(B), primary and secondary electron accepting quinones) was investigated in chromatophores of Rb. capsulatus, varying the temperature down to -65 degrees C. The analysis of the flash-induced pattern for the formation of P(+)Q(A)Q(B)(-) shows that the diminished yield, when lowering the temperature, is not due to a homogeneous slowing of the rate constant k(AB) of the Q(A)(-)Q(B) --> Q(A)Q(B)(-) electron transfer but to a distribution of conformations that modulate the electron transfer rate over more than 3 orders of magnitude. This distribution appears "frozen", as no dynamic redistribution was observed over time ranges > 10 s (below -25 degrees C). The kinetic pattern was analyzed to estimate the shape of the distribution of k(AB), showing a bell-shaped band on the high rate side and a fraction of "blocked" reaction centers (RCs) with very slow k(AB). When the temperature is lowered, the high rate band moves to slower rate regions and the fraction of blocked RCs increases at the expense of the high rate band. The RCs that recombine from the P(+)Q(A)Q(B)(-) state appear temporarily converted to a state with rapid k(AB), indicating that the stabilized state described by Kleinfeld et al. (Biochemistry 1984, 23, 5780-5786) is still accessible at -60 degrees C. Topics: Bacteriochlorophylls; Benzoquinones; Cold Temperature; Cytochromes c2; Electrons; Kinetics; Photosynthetic Reaction Center Complex Proteins; Protein Conformation; Rhodobacter capsulatus; Spectrometry, Fluorescence; Thermodynamics; Ubiquinone | 2008 |
CABC1 gene mutations cause ubiquinone deficiency with cerebellar ataxia and seizures.
Coenzyme Q(10) (CoQ(10)) plays a pivotal role in oxidative phosphorylation (OXPHOS) in that it distributes electrons between the various dehydrogenases and the cytochrome segments of the respiratory chain. Primary coenzyme Q(10) deficiency represents a clinically heterogeneous condition suggestive of genetic heterogeneity, and several disease genes have been previously identified. The CABC1 gene, also called COQ8 or ADCK3, is the human homolog of the yeast ABC1/COQ8 gene, one of the numerous genes involved in the ubiquinone biosynthesis pathway. The exact function of the Abc1/Coq8 protein is as yet unknown, but this protein is classified as a putative protein kinase. We report here CABC1 gene mutations in four ubiquinone-deficient patients in three distinct families. These patients presented a similar progressive neurological disorder with cerebellar atrophy and seizures. In all cases, enzymological studies pointed to ubiquinone deficiency. CoQ(10) deficiency was confirmed by decreased content of ubiquinone in muscle. Various missense mutations (R213W, G272V, G272D, and E551K) modifying highly conserved amino acids of the protein and a 1 bp frameshift insertion c.[1812_1813insG] were identified. The missense mutations were introduced into the yeast ABC1/COQ8 gene and expressed in a Saccharomyces cerevisiae strain in which the ABC1/COQ8 gene was deleted. All the missense mutations resulted in a respiratory phenotype with no or decreased growth on glycerol medium and a severe reduction in ubiquinone synthesis, demonstrating that these mutations alter the protein function. Topics: Adolescent; Adult; Amino Acid Sequence; Benzoquinones; Brain; Cerebellar Ataxia; Female; Haplotypes; Humans; Magnetic Resonance Imaging; Male; Molecular Sequence Data; Muscle, Skeletal; Mutation, Missense; Pedigree; Seizures; Ubiquinone | 2008 |
Theoretical study of the energetics of the reactions of triplet dioxygen with hydroquinone, semiquinone, and their protonated forms: relation to the mechanism of superoxide generation in the respiratory chain.
One-electron reduction of the dioxygen molecule by the reduced form of mitochondrial ubiquinones (Q) of the NADH dehydrogenase (complex I) and mitochondrial cytochrome bc1 (complex III) is believed to be the main source of the superoxide anion radical O2*- and the hydroperoxide radical OOH*. In this work, we modeled the energetics of four possible reactions of the triplet ((3)Sigma(g)) dioxygen-molecule reduction by fully reduced and protonated ubiquinone (QH2; reaction 1), its deprotonated form (QH-; reaction 2), the semiquinone radical (QH*; reaction 3), and the semiquinone anion radical (Q*-; reaction 4), by means of ab initio calculations with the 6-31G(d) and 6-31+G(d) basis set in the restricted open-shell Hartree-Fock (ROHF), unrestricted Hartree-Fock (UHF), and complete active space self-consistent field (CASSCF) with dynamic correlation [at the second-order Møller-Plesset (MP2) or multiple reference Møller-Plesset (MRMP), respectively] schemes and the basis set superposition error (BSSE) correction included, as well as semiempirical AM1 and PM3 calculations in the UHF and ROHF schemes. 2-Butene-1,4-dione and p-benzoquinone were selected as model compounds. For the reduced forms of both compounds, reaction 1 turned out to be energetically unfavorable at all levels of theory, this agreeing with the experimentally observed diminished reductive properties of hydroquinone derivatives at low pH. For 2-butene-1,4-dione treated at the most advanced MRMP/CASSCF/6-31+G(d) level, the energies of reactions 1-4 are 4.7, -34.3, -15.0, and -4.1 kcal/mol, respectively. This finding suggests that reactions 2 and 3 are the most likely mechanisms of electron transfer to molecular oxygen in aprotic environments and that proton transfer is involved in this process. Nearly the same energies of reactions 2 and 3 were calculated at the MRMP/CASSCF/6-31+G(d) level for reduced forms of p-benzoquinone. Inclusion of diffuse functions in the basis set and dynamic correlation at the CASSCF level appears essential. Because deprotonated ubiquinol is unlikely to exist in physiological environments, reaction 3 appears to be the most likely mechanism of one-electron reduction of oxygen; however, if oxygen can penetrate cytochrome bc1 as far as the Q(o) center where ubiquinol can be deprotonated, reaction 2 can also come into play. The energies of reactions 2 and 3 calculated at the MRMP/CASSCF/6-31+G(d) level are most closely reproduced in the ab initio and semiempirical UHF PM3 c Topics: Benzoquinones; Butanones; Hydroquinones; Mitochondria; Models, Chemical; Models, Molecular; Oxidation-Reduction; Oxygen; Protons; Thermodynamics; Ubiquinone | 2007 |
Structural and computational analysis of the quinone-binding site of complex II (succinate-ubiquinone oxidoreductase): a mechanism of electron transfer and proton conduction during ubiquinone reduction.
The transfer of electrons and protons between membrane-bound respiratory complexes is facilitated by lipid-soluble redox-active quinone molecules (Q). This work presents a structural analysis of the quinone-binding site (Q-site) identified in succinate:ubiquinone oxidoreductase (SQR) from Escherichia coli. SQR, often referred to as Complex II or succinate dehydrogenase, is a functional member of the Krebs cycle and the aerobic respiratory chain and couples the oxidation of succinate to fumarate with the reduction of quinone to quinol (QH(2)). The interaction between ubiquinone and the Q-site of the protein appears to be mediated solely by hydrogen bonding between the O1 carbonyl group of the quinone and the side chain of a conserved tyrosine residue. In this work, SQR was co-crystallized with the ubiquinone binding-site inhibitor Atpenin A5 (AA5) to confirm the binding position of the inhibitor and reveal additional structural details of the Q-site. The electron density for AA5 was located within the same hydrophobic pocket as ubiquinone at, however, a different position within the pocket. AA5 was bound deeper into the site prompting further assessment using protein-ligand docking experiments in silico. The initial interpretation of the Q-site was re-evaluated in the light of the new SQR-AA5 structure and protein-ligand docking data. Two binding positions, the Q(1)-site and Q(2)-site, are proposed for the E. coli SQR quinone-binding site to explain these data. At the Q(2)-site, the side chains of a serine and histidine residue are suitably positioned to provide hydrogen bonding partners to the O4 carbonyl and methoxy groups of ubiquinone, respectively. This allows us to propose a mechanism for the reduction of ubiquinone during the catalytic turnover of the enzyme. Topics: Amino Acid Sequence; Benzoquinones; Binding Sites; Catalysis; Computational Biology; Crystallography, X-Ray; Electron Transport; Electron Transport Complex II; Electrons; Escherichia coli; Histidine; Hydrogen Bonding; Ligands; Models, Chemical; Models, Molecular; Molecular Sequence Data; Oxygen; Phenotype; Protein Binding; Protein Conformation; Protein Structure, Tertiary; Protons; Pyridones; Quinones; Sequence Homology, Amino Acid; Serine; Succinate Dehydrogenase; Ubiquinone | 2006 |
Structural and biochemical evidence for an enzymatic quinone redox cycle in Escherichia coli: identification of a novel quinol monooxygenase.
Naturally synthesized quinones perform a variety of important cellular functions. Escherichia coli produce both ubiquinone and menaquinone, which are involved in electron transport. However, semiquinone intermediates produced during the one-electron reduction of these compounds, as well as through auto-oxidation of the hydroxyquinone product, generate reactive oxygen species that stress the cell. Here, we present the crystal structure of YgiN, a protein of hitherto unknown function. The three-dimensional fold of YgiN is similar to that of ActVA-Orf6 monooxygenase, which acts on hydroxyquinone substrates. YgiN shares a promoter with "modulator of drug activity B," a protein with activity similar to that of mammalian DT-diaphorase capable of reducing mendione. YgiN was able to reoxidize menadiol, the product of the "modulator of drug activity B" (MdaB) enzymatic reaction. We therefore refer to YgiN as quinol monooxygenase. Modulator of drug activity B is reported to be involved in the protection of cells from reactive oxygen species formed during single electron oxidation and reduction reactions. The enzymatic activities, together with the structural characterization of YgiN, lend evidence to the possible existence of a novel quinone redox cycle in E. coli. Topics: Benzoquinones; Binding Sites; Crystallography, X-Ray; Dimerization; Drug Resistance, Neoplasm; Electron Transport; Electrons; Escherichia coli; Escherichia coli Proteins; Mixed Function Oxygenases; Models, Chemical; Models, Molecular; Oxidation-Reduction; Promoter Regions, Genetic; Protein Conformation; Protein Folding; Protein Structure, Secondary; Quinones; Reactive Oxygen Species; Recombinant Proteins; Spectrophotometry; Time Factors; Ubiquinone; Vitamin K 2 | 2005 |
Coq3 and Coq4 define a polypeptide complex in yeast mitochondria for the biosynthesis of coenzyme Q.
Coenzyme Q (Q) is a redox active lipid essential for aerobic respiration in eukaryotes. In Saccharomyces cerevisiae at least eight mitochondrial polypeptides, designated Coq1-Coq8, are required for Q biosynthesis. Here we present physical evidence for a coenzyme Q-biosynthetic polypeptide complex in isolated mitochondria. Separation of digitonin-solubilized mitochondrial extracts in one- and two-dimensional Blue Native PAGE analyses shows that Coq3 and Coq4 polypeptides co-migrate as high molecular mass complexes. Similarly, gel filtration chromatography shows that Coq1p, Coq3p, Coq4p, Coq5p, and Coq6p elute in fractions higher than expected for their respective subunit molecular masses. Coq3p, Coq4p, and Coq6p coelute with an apparent molecular mass exceeding 700 kDa. Coq3 O-methyltransferase activity, a surrogate for Q biosynthesis and complex activity, also elutes at this high molecular mass. We have determined the quinone content in lipid extracts of gel filtration fractions by liquid chromatography-tandem mass spectrometry and find that demethoxy-Q(6) is enriched in fractions with Coq3p. Co-precipitation of biotinylated-Coq3 and Coq4 polypeptide from digitonin-solubilized mitochondrial extracts shows their physical association. This study identifies Coq3p and Coq4p as defining members of a Q-biosynthetic Coq polypeptide complex. Topics: Benzoquinones; Biotinylation; Chemical Phenomena; Chemical Precipitation; Chemistry, Physical; Chromatography, Gel; Chromatography, Liquid; Digitonin; Mass Spectrometry; Methyltransferases; Mitochondria; Mitochondrial Proteins; Molecular Weight; Saccharomyces cerevisiae; Saccharomyces cerevisiae Proteins; Solubility; Ubiquinone | 2005 |
Absence of large-scale displacement of quinone QB in bacterial photosynthetic reaction centers.
Photosynthesis transforms light into chemical energy by coupling electron transfer to proton uptake at the quinone Q(B). The possibility of initiating this process with a brief pulse of light and the known X-ray structure makes the photosynthetic bacterial reaction center a paradigm for studying coupled electron-proton transfer in biology. It has been established that electron transfer from the primary quinone Q(A) to Q(B) is gated by a protein conformational change. On the basis of a dramatic difference in the location of Q(B) in structures derived from crystals cooled to 90 K either under illumination or in the dark, a functional model for the gating mechanism was proposed whereby neutral Q(B) moves 4.5 A before receiving the electron from Q(A)(-) [Stowell, M. H. B., McPhillips, T. M., Rees, D. C., Soltis, S. M., Abresch, E., and Feher, G. (1997) Science 276, 812-816]. Isotope-edited FTIR difference spectroscopy of Q(B) photoreduction at 290 and 85 K is used to investigate whether Q(B) moves upon reduction. We show that the specific interactions of the carbonyl groups of Q(B) and Q(B)(-) with the protein at a single binding site remain identical at both temperatures. Therefore, the different locations of Q(B) reported in many X-ray crystal structures probably are unrelated to functional electron transfer from Q(A)(-) to Q(B). Topics: Benzoquinones; Binding Sites; Crystallography, X-Ray; Electron Transport; Freezing; Hydrogen Bonding; Light; Oxidation-Reduction; Photolysis; Photosynthetic Reaction Center Complex Proteins; Rhodobacter sphaeroides; Spectroscopy, Fourier Transform Infrared; Temperature; Ubiquinone | 2004 |
Demethoxy-Q, an intermediate of coenzyme Q biosynthesis, fails to support respiration in Saccharomyces cerevisiae and lacks antioxidant activity.
Caenorhabditis elegans clk-1 mutants cannot produce coenzyme Q(9) and instead accumulate demethoxy-Q(9) (DMQ(9)). DMQ(9) has been proposed to be responsible for the extended lifespan of clk-1 mutants, theoretically through its enhanced antioxidant properties and its decreased function in respiratory chain electron transport. In the present study, we assess the functional roles of DMQ(6) in the yeast Saccharomyces cerevisiae. Three mutations designed to mirror the clk-1 mutations of C. elegans were introduced into COQ7, the yeast homologue of clk-1: E233K, predicted to disrupt the di-iron carboxylate site considered essential for hydroxylase activity; L237Stop, a deletion of 36 amino acid residues from the carboxyl terminus; and P175Stop, a deletion of the carboxyl-terminal half of Coq7p. Growth on glycerol, quinone content, respiratory function, and response to oxidative stress were analyzed in each of the coq7 mutant strains. Yeast strains lacking Q(6) and producing solely DMQ were respiratory deficient and unable to support (6)either NADH-cytochrome c reductase or succinate-cytochrome c reductase activities. DMQ(6) failed to protect cells against oxidative stress generated by H(2)O(2) or linolenic acid. Thus, in the yeast model system, DMQ does not support respiratory activity and fails to act as an effective antioxidant. These results suggest that the life span extension observed in the C. elegans clk-1 mutants cannot be attributed to the presence of DMQ per se. Topics: Alleles; alpha-Linolenic Acid; Antioxidants; Benzoquinones; Binding Sites; Blotting, Western; Chromatography, High Pressure Liquid; Electron Transport; Flow Cytometry; Glycerol; Hydrogen Peroxide; Mass Spectrometry; Mitochondria; Models, Chemical; Mutation; NADH Dehydrogenase; Oxidative Stress; Oxygen Consumption; Peptides; Plasmids; Saccharomyces cerevisiae; Superoxides; Time Factors; Ubiquinone | 2004 |
Structural basis for the quinone reduction in the bc1 complex: a comparative analysis of crystal structures of mitochondrial cytochrome bc1 with bound substrate and inhibitors at the Qi site.
Cytochrome bc(1) is an integral membrane protein complex essential to cellular respiration and photosynthesis. The Q cycle reaction mechanism of bc(1) postulates a separated quinone reduction (Q(i)) and quinol oxidation (Q(o)) site. In a complete catalytic cycle, a quinone molecule at the Q(i) site receives two electrons from the b(H) heme and two protons from the negative side of the membrane; this process is specifically inhibited by antimycin A and NQNO. The structures of bovine mitochondrial bc(1) in the presence or absence of bound substrate ubiquinone and with either the bound antimycin A(1) or NQNO were determined and refined. A ubiquinone with its first two isoprenoid repeats and an antimycin A(1) were identified in the Q(i) pocket of the substrate and inhibitor bound structures, respectively; the NQNO, on the other hand, was identified in both Q(i) and Q(o) pockets in the inhibitor complex. The two inhibitors occupied different portions of the Q(i) pocket and competed with substrate for binding. In the Q(o) pocket, the NQNO behaves similarly to stigmatellin, inducing an iron-sulfur protein conformational arrest. Extensive binding interactions and conformational adjustments of residues lining the Q(i) pocket provide a structural basis for the high affinity binding of antimycin A and for phenotypes of inhibitor resistance. A two-water-mediated ubiquinone protonation mechanism is proposed involving three Q(i) site residues His(201), Lys(227), and Asp(228). Topics: Amino Acid Sequence; Animals; Antimycin A; Benzoquinones; Cattle; Crystallography, X-Ray; Cytochrome b Group; Electron Transport Complex III; Enzyme Inhibitors; Hydroxyquinolines; Mitochondria, Heart; Models, Molecular; Molecular Sequence Data; Oxidation-Reduction; Protein Binding; Protein Conformation; Protein Subunits; Structure-Activity Relationship; Substrate Specificity; Ubiquinone | 2003 |
Reproductive fitness and quinone content of Caenorhabditis elegans clk-1 mutants fed coenzyme Q isoforms of varying length.
Caenorhabditis elegans clk-1 mutants lack coenzyme Q9 and accumulate the biosynthetic intermediate demethoxy-Q9. A dietary source of ubiquinone (Q) is required for larval growth and development of the gonad and germ cells. We considered that uptake of the shorter Q8 isoform present in the Escherichia coli food may contribute to the Clk phenotypes of slowed development and reduced brood size observed when the animals are fed Q-replete E. coli. To test the effect of isoprene tail length, N2 and clk-1 animals were fed E. coli engineered to produce Q7, Q8, Q9, or Q10. Wild-type nematodes showed no change in reproductive fitness regardless of the Qn isoform fed. clk-1(e2519) fed the Q9 diet showed increased egg production; however, this diet did not improve reproductive fitness of the clk-1(qm30) animals. Furthermore, animals with the more severe clk-1(qm30) allele become sterile and their progeny inviable when fed Q7-containing bacteria. The content of Q7 in the mitochondria of clk-1 animals was decreased relative to Q8, suggesting less effective transport of Q7 to the mitochondria, impaired retention, or decreased stability. Additionally, regardless of E. coli diet, clk-1(qm30) animals contain a dysfunctional dense form of mitochondria. The gonads of clk-1(qm30) worms fed Q7-containing food were severely shrunken and disordered. The differential fertility of clk-1 mutant nematodes fed Q isoforms may result from changes in Q localization, altered recognition by Q-binding proteins, and/or potential defects in mitochondrial function resulting from the mutant CLK-1 polypeptide itself. Topics: Animals; Benzoquinones; Caenorhabditis elegans; Caenorhabditis elegans Proteins; Mitochondria; Mutation; Phenotype; Protein Isoforms; Reproduction; Structure-Activity Relationship; Ubiquinone | 2003 |
Assay of coenzyme Q(10) in plasma by a single dilution step.
A new method is described for determining coenzyme Q(10) (CoQ(10)) in plasma. The method is based on oxidation of CoQ(10) in the sample by treating it with para-benzoquinone followed by extraction with 1-propanol and direct injection into the HPLC apparatus. This method achieves a linear detector response for peak area measurements over the concentration range of 0.05-3.47 microM. Diode array analysis of the peak was consistent with CoQ(10) spectrum. Supplementation of the samples with known amounts of CoQ(10) yielded a quantitative recovery of 96-98.5%; the method showed a level of quantitation of 1.23 nmol per HPLC injection (200 microl of propanol extract containing 33.3 microl of plasma). A correlation of r = 0.99 (P < 0.0001) was found with a reference electrochemical detection method. Within run precision showed a CV% of 1.6 for samples approaching normal values (1.02 microM). Day-to-day precision was also close to 2%. Topics: Benzoquinones; Calibration; Centrifugation; Chromatography, High Pressure Liquid; Coenzymes; Electrochemistry; Enzyme Stability; Humans; Indicator Dilution Techniques; Linear Models; Oxidation-Reduction; Reproducibility of Results; Sensitivity and Specificity; Spectrophotometry, Ultraviolet; Ubiquinone | 2002 |
Reduction of 1,4-quinone and ubiquinones by hydrogen atom transfer under UVA irradiation.
1,4-Benzoquinone, coenzyme Q0 and Q10 were reacted with a series of hydrogen donors in the ESR cavity in the presence or absence of UVA irradiation. The signals of the radicals generated from the hydrogen donors or of those of the semiquinones were detected. The reaction mechanism was interpreted by a hydrogen atom transfer instead of the usual electron transfer mechanism on the basis of the redox potentials of the reactants and the Marcus theory. The hydrogen atom transfer is explained by the excited triplet state of quinones, which, on the basis of quantum mechanic calculations, may be reached even under visible light. In some cases, hydrogen atom transfer was also observed without irradiation, although to a lesser extent. Topics: Benzoquinones; Coenzymes; Electron Spin Resonance Spectroscopy; Free Radicals; Hydrogen; Hydrogen Bonding; Hydrogen Peroxide; Oxidation-Reduction; Protons; Thermodynamics; Ubiquinone; Ultraviolet Rays | 2002 |
Disulfide bonds are generated by quinone reduction.
The chemistry of disulfide exchange in biological systems is well studied. However, very little information is available concerning the actual origin of disulfide bonds. Here we show that DsbB, a protein required for disulfide bond formation in vivo, uses the oxidizing power of quinones to generate disulfides de novo. This is a novel catalytic activity, which to our knowledge has not yet been described. This catalytic activity is apparently the major source of disulfides in vivo. We developed a new assay to characterize further this previously undescribed enzymatic activity, and we show that quinones get reduced during the course of the reaction. DsbB contains a single high affinity quinone-binding site. We reconstitute oxidative folding in vitro in the presence of the following components that are necessary in vivo: DsbA, DsbB, and quinone. We show that the oxidative refolding of ribonuclease A is catalyzed by this system in a quinone-dependent manner. The disulfide isomerase DsbC is required to regain ribonuclease activity suggesting that the DsbA-DsbB system introduces at least some non-native disulfide bonds. We show that the oxidative and isomerase systems are kinetically isolated in vitro. This helps explain how the cell avoids oxidative inactivation of the disulfide isomerization pathway. Topics: Bacterial Proteins; Benzoquinones; Catalysis; Chromatography, High Pressure Liquid; Disulfides; Kinetics; Membrane Proteins; Oxidation-Reduction; Oxidoreductases; Protein Denaturation; Protein Disulfide-Isomerases; Protein Folding; Ribonuclease, Pancreatic; Ubiquinone | 2000 |
The Saccharomyces cerevisiae succinate-ubiquinone oxidoreductase. Identification of Sdh3p amino acid residues involved in ubiquinone binding.
Succinate dehydrogenase (SDH) participates in the mitochondrial electron transport chain by oxidizing succinate to fumarate and transferring the electrons to ubiquinone. In yeast, it is composed of a catalytic dimer, comprising the Sdh1p and Sdh2p subunits, and a membrane domain, comprising two smaller hydrophobic subunits, Sdh3p and Sdh4p, which anchor the enzyme to the mitochondrial inner membrane. To investigate the role of the Sdh3p anchor polypeptide in enzyme assembly and catalysis, we isolated and characterized seven mutations in the SDH3 gene. Two mutations are premature truncations of Sdh3p with losses of one or three transmembrane segments. The remaining five are missense mutations that are clustered between amino acids 103 and 117, which are proposed to be located in transmembrane segment II or the matrix-localized loop connecting segments II and III. Three mutations, F103V, H113Q, and W116R, strongly but specifically impair quinone reductase activities but have only minor effects on enzyme assembly. The clustering of the mutations strongly suggests that a ubiquinone-binding site is associated with this region of Sdh3p. In addition, the biphasic inhibition of quinone reductase activity by a dinitrophenol inhibitor supports the hypothesis that two distinct quinone-binding sites are present in the yeast SDH. Topics: Amino Acid Sequence; Benzoquinones; Binding Sites; Cytochrome b Group; Enzyme Stability; Molecular Sequence Data; Mutagenesis; NAD(P)H Dehydrogenase (Quinone); Oxidation-Reduction; Saccharomyces cerevisiae; Succinate Dehydrogenase; Ubiquinone | 1999 |
Effect of arylthiolated 2,3-dimethoxy-1,4-benzoquinones on respiratory activity and lipid peroxidation in bovine heart mitochondria.
A series of arylthiolated 2,3-dimethoxy-1,4-benzoquinones was synthesized and tested for the effect on the respiratory system and the lipid peroxidation in bovine heart mitochondria (BHM). These quinones showed intense inhibitory activities on the respiratory system in BHM. Their inhibitory activity in the succinate oxidase system was greater than that in the NADH oxidase system. No difference between the difference spectra, with and without these quinones, of the reduced minus oxidized forms of cytochromes (cyt.) suggested that these quinones inhibit at the site after cyt. a+a3 in the respiratory chain. Moreover, these quinones were as efficient as exogenous ubiquinone-10 (UQ-10) for the inhibition of lipid peroxidation. 5- And 5,6-di-arylthio groups on the quinone ring were found to be favorable for inhibition of the respiratory system and lipid peroxidation. Our results suggest that arylthiolated 2,3-dimethoxy-1,4-benzoquinones act as antioxidants by increasing the amount of endogenous reduced UQ-10 in BHM. Topics: Animals; Benzoquinones; Cattle; Electron Transport; Lipid Peroxidation; Mitochondria, Heart; Multienzyme Complexes; NADH, NADPH Oxidoreductases; Oxidoreductases; Ubiquinone | 1998 |
Quenching of chlorophyll fluorescence by quinones.
Quinones caused quenching of Chl a fluorescence in native and model systems. Menadione quenched twofold the fluorescence of Chl a and BChl a in pea chloroplasts, chromatophores of purple bacteria, and liposomes at concentrations of 50-80 microM. To obtain twofold quenching in Triton X-100 micelles and in ethanol, the addition of 1.3 mM and 11 mM menadione was required, respectively. A proportional decrease in the lifetime and yield of Chl a fluorescence in chloroplasts, observed as the menadione concentration increased, is indicative of the efficient excitation energy transfer from bulk Chl to menadione. The decrease in the lifetime and yield of fluorescence was close to proportional in liposomes, but not in detergent micelles. The insensitivity of the menadione quenching effect to DCMU in chloroplasts, and similarity of its action in chloroplasts and liposomes indicate that menadione in chloroplasts interacts with antenna Chl, i.e., nonphotochemical quenching of fluorescence occurs. Topics: Bacterial Chromatophores; Bacteriochlorophylls; Benzoquinones; Chlorophyll; Chlorophyll A; Chloroplasts; Diuron; Fluorescence; Liposomes; Micelles; Pisum sativum; Quinones; Rhodobacter sphaeroides; Rhodospirillum rubrum; Spectrometry, Fluorescence; Ubiquinone; Vitamin K | 1998 |
The quinone-binding site in succinate-ubiquinone reductase from Escherichia coli. Quinone-binding domain and amino acid residues involved in quinone binding.
When purified ubiquinone (Q)-depleted succinate-ubiquinone reductase from Escherichia coli is photoaffinity-labeled with 3-azido-2-methyl-5-methoxy-[3H]6-geranyl-1,4-benzoquinone ([3H]azido-Q) followed by SDS-polyacrylamide gel electrophoresis, radioactivity is found in the SdhC subunit, indicating that this subunit is responsible for ubiquinone binding. An [3H]azido-Q-linked peptide, with a retention time of 61.7 min, is obtained by high performance liquid chromatography of the protease K digest of [3H]azido-Q-labeled SdhC obtained from preparative SDS-polyacrylamide gel electrophoresis on labeled reductase. The partial N-terminal amino acid sequence of this peptide is NH2-TIRFPITAIASILHRVS-, corresponding to residues 17-33. The ubiquinone-binding domain in the proposed structural model of SdhC, constructed based on the hydropathy plot of the deduced amino acid sequence of this protein, is located at the N-terminal end toward the transmembrane helix I. To identify amino acid residues responsible for ubiquinone binding, substitution mutations at the putative ubiquinone-binding region of SdhC were generated and characterized. E. coli NM256 lacking genomic succinate-Q reductase genes was constructed and used to harbor the mutated succinate-Q reductase genes in a low copy number pRKD418 plasmid. Substitution of serine 27 of SdhC with alanine, cysteine, or threonine or substitution of arginine 31 with alanine, lysine, or histidine yields cells unable to grow aerobically in minimum medium with succinate as carbon source. Furthermore, little succinate-ubiquinone reductase activity and [3H]azido-Q uptake are detected in succinate-ubiquinone reductases prepared from these mutant cells grown aerobically in LB medium. These results indicate that the hydroxyl group, the size of the amino acid side chain at position 27, and the guanidino group at position 31 of SdhC are critical for succinate-ubiquinone reductase activity, perhaps by formation of hydrogen bonds with carbonyl groups of the 1,4-benzoquinone ring of the quinone molecule. The hydroxyl group, but not the size of the amino acid side chain, at position 33 of SdhC is also important, because Ser-33 can be substituted with threonine but not with alanine. Topics: Affinity Labels; Amino Acid Sequence; Amino Acid Substitution; Azides; Base Sequence; Benzoquinones; Binding Sites; Chromosome Mapping; Chromosomes, Bacterial; Electron Transport Complex II; Escherichia coli; Kinetics; Models, Molecular; Molecular Sequence Data; Multienzyme Complexes; Mutagenesis, Site-Directed; Oligodeoxyribonucleotides; Operon; Oxidoreductases; Peptide Fragments; Point Mutation; Protein Structure, Secondary; Recombinant Proteins; Succinate Dehydrogenase; Ubiquinone | 1998 |
The COQ5 gene encodes a yeast mitochondrial protein necessary for ubiquinone biosynthesis and the assembly of the respiratory chain.
Saccharomyces cerevisiae is a facultative anaerobe capable of meeting its energy requirements by fermentation and is thus an ideal system for studying the biogenesis of respiring mitochondria. We have isolated a respiration-deficient mutant exhibiting a pleiotropic loss of the mitochondrial electron transport chain. The corresponding wild-type gene, COQ5, was cloned, sequenced, and able to restore respiratory growth. Deletion of the chromosomal COQ5 gene results in a respiration deficiency and reduced levels of respiratory protein components. Exogenously added decylubiquinone can partially restore electron transport chain function to mitochondrial membranes from the deletion mutant. The COQ5 nucleotide sequence predicts a polypeptide of 307 amino acids containing a mitochondrial targeting signal. COQ5p is 43% identical to the polypeptide predicted by the Escherichia coli open reading frame, o251 (1). The COQ5 gene, when introduced into E. coli, complements the respiratory deficiency of an ubiE mutant that maps near o251, suggesting that it is the yeast homolog of the ubiE gene product. We conclude that the COQ5 gene encodes the mitochondria-localized 2-hexaprenyl-6-methoxy-1,4-benzoquinone methyltransferase of the yeast ubiquinone biosynthetic pathway. Topics: Amino Acid Sequence; Benzoquinones; Electron Transport; Fungal Proteins; Genes, Fungal; Methyltransferases; Mitochondria; Models, Chemical; Molecular Sequence Data; Mutagenesis, Site-Directed; Oxidoreductases; Oxygen Consumption; Saccharomyces cerevisiae; Saccharomyces cerevisiae Proteins; Sequence Alignment; Sequence Deletion; Ubiquinone | 1997 |
Saturation kinetics of coenzyme Q in NADH oxidation: rate enhancement by incorporation of excess quinone.
In beef heart mitochondria it has been found that the Km for coenzyme Q10 of the NADH oxidation system is in the range of the membrane concentration of the quinone; this is contrary to succinate oxidation which is in Vmax with respect to quinone content. The same proportional difference between the two systems is maintained in their affinities for the exogenous acceptor CoQ1 in non-extracted mitochondria. The Km of succinate- coenzyme Q reductase for CoQ1 is reversibly lowered in CoQ-depleted mitochondria; while in contrast the Km for NADH-coenzyme Q reductase is reversibly increased by CoQ extraction. Incorporation of exogenous quinones by co-sonication with submitochondrial particles, as evidenced by fluorescence quenching of pyrene, enhances NADH-cytochrome c reductase activity in accordance with the lack of saturation of the former system. Topics: Animals; Antioxidants; Benzoquinones; Cattle; Electron Transport Complex II; Kinetics; Mitochondria, Heart; Multienzyme Complexes; NAD; NAD(P)H Dehydrogenase (Quinone); NADH Dehydrogenase; Oxidation-Reduction; Oxidoreductases; Sonication; Succinate Dehydrogenase; Ubiquinone | 1997 |
The specificity of mitochondrial complex I for ubiquinones.
We report the first detailed study on the ubiquinone (coenzyme Q; abbreviated to Q) analogue specificity of mitochondrial complex I, NADH:Q reductase, in intact submitochondrial particles. The enzymic function of complex I has been investigated using a series of analogues of Q as electron acceptor substrates for both electron transport activity and the associated generation of membrane potential. Q analogues with a saturated substituent of one to three carbons at position 6 of the 2,3-dimethoxy-5-methyl-1,4-benzoquinone ring have the fastest rates of electron transport activity, and analogues with a substituent of seven to nine carbon atoms have the highest values of association constant derived from NADH:Q reductase activity. The rate of NADH:Q reductase activity is potently but incompletely inhibited by rotenone, and the residual rotenone-insensitive rate is stimulated by Q analogues in different ways depending on the hydrophobicity of their substituent. Membrane potential measurements have been undertaken to evaluate the energetic efficiency of complex I with various Q analogues. Only hydrophobic analogues such as nonyl-Q or undecyl-Q show an efficiency of membrane potential generation equivalent to that of endogenous Q. The less hydrophobic analogues as well as the isoprenoid analogue Q-2 are more efficient as substrates for the redox activity of complex I than for membrane potential generation. Thus the hydrophilic Q analogues act also as electron sinks and interact incompletely with the physiological Q site in complex I that pumps protons and generates membrane potential. Topics: Animals; Benzoquinones; Binding Sites; Cattle; Membrane Potentials; Mitochondria; NAD; NAD(P)H Dehydrogenase (Quinone); Oxidation-Reduction; Rotenone; Sensitivity and Specificity; Substrate Specificity; Ubiquinone | 1996 |
Stabilization of a semiquinone radical at the high-affinity quinone-binding site (QH) of the Escherichia coli bo-type ubiquinol oxidase.
Reaction of ubiquinone in the high-affinity quinone-binding site (QH) in bo-type ubiquinol oxidase from Escherichia coli was revealed by EPR and optical studies. In the QH site, ubiquinol was shown to be oxidized to ubisemiquinone and to ubiquinone, while no semiquinone signal was detected in the oxidase isolated from mutant cells that cannot synthesize ubiquinone. The QH site highly stabilized ubisemiquinone radical with a stability constant of 1-4 at pH 8.5 and the stability became lower at the lower pH. Midpoint potential of QH2/Q couple was -2 mV at pH 8.5 and showed -60 mV/pH dependence indicative of 2H+/2e- reaction. The Em was more negative than that of low-spin heme b above pH 7.0. We conclude that the QH mediates intramolecular electron transfer from ubiquinol in the low-affinity quinol oxidation site (QL) to low-spin heme b. Unique roles of the quinone-binding sites in the bacterial ubiquinol oxidase are discussed. Topics: Benzoquinones; Binding Sites; Coenzymes; Electron Spin Resonance Spectroscopy; Electron Transport; Electron Transport Complex IV; Enzyme Stability; Escherichia coli; Heme; Hydrogen-Ion Concentration; Oxidation-Reduction; Potentiometry; Spectrum Analysis; Ubiquinone | 1995 |
The relationship between electron flux and the redox poise of the quinone pool in plant mitochondria. Interplay between quinol-oxidizing and quinone-reducing pathways.
The dependence of electron flux through quinone-reducing and quinol-oxidizing pathways on the redox state of the ubiquinone (Q) pool was investigated in plant mitochondria isolated from potato (Solanum tuberosum cv. Bintje, fresh tissue and callus), sweet potato (Ipomoea batatas) and Arum italicum. We have determined the redox state of the Q pool with two different methods, the Q-electrode and Q-extraction techniques. Although results from the two techniques agree well, in all tissues tested (with the exception of fresh potato) an inactive pool of QH2 was detected by the extraction technique that was not observed with the electrode. In potato callus mitochondria, an inactive Q pool was also found. An advantage of the extraction method is that it permits determination of the Q redox state in the presence of substances that interfere with the Q-electrode, such as benzohydroxamate and NADH. We have studied the relation between rate and Q redox state for both quinol-oxidizing and quinone-reducing pathways under a variety of metabolic conditions including state 3, state 4, in the presence of myxothiazol, and benzohydroxamate. Under state 4 conditions or in the presence of myxothiazol, a non-linear dependence of the rate of respiration on the Q-redox state was observed in potato callus mitochondria and in sweet potato mitochondria. The addition of benzohydroxamate, under state 4 conditions, removed this non-linearity confirming that it is due to activity of the cyanide-resistant pathway. The relation between rate and Q redox state for the external NADH dehydrogenase in potato callus mitochondria was found to differ from that of succinate dehydrogenase. It is suggested that the oxidation of cytoplasmic NADH in vivo uses the cyanide-resistant pathway more than the pathway involving the oxidation of succinate. A model is used to predict the kinetic behaviour of the respiratory network. It is shown that titrations with inhibitors of the alternative oxidase cannot be used to demonstrate a pure overflow function of the alternative oxidase. Topics: Benzoquinones; Electron Transport; Hydroquinones; Kinetics; Methacrylates; Mitochondria; NAD; NADH Dehydrogenase; Oxidation-Reduction; Plants; Solanum tuberosum; Succinate Dehydrogenase; Succinates; Succinic Acid; Thiazoles; Ubiquinone | 1994 |
Comparison of structure of quinone redox site in the mitochondrial cytochrome-bc1 complex and photosystem II (QB site).
A series of nitrophenolic electron-transport inhibitors (2-substituted 4,6-dinitrophenols) of rat liver mitochondrial cytochrome-bc1 complex and of photosystem II (QB site) of spinach thylakoids was synthesized. The structure/inhibitory-activity relationship was examined to elucidate differences in the three-dimensional structure of the quinone redox site in the two systems. These inhibitors occupy the ubiquinone redox site of cytochrome-bc1 complex competitively with natural ubiquinol, probably at a Qo reaction center. The inhibitory activity tended to increase with the length of the 2-substituent, which may correspond to the isoprenoid side chain of ubiquinone and plastoquinone, increased in both experimental systems. However, the strict structural requirements of the 2-substituent for binding to the ubiquinone or plastoquinone redox site were not identical. The alkyl substituents with a branching structure at the alpha-position to the benzene ring were favorable for inhibition of the cytochrome-bc1 complex, but not of photosystem II. Molecular-orbital calculations indicated that the main chain of 2-substituents with an alpha-branching structure was almost perpendicular to the benzene-ring plane because of steric congestion between the alpha-methyl and phenolic OH groups. The main chain of 2-substituents without an alpha-branching structure was flexible. Molecular-orbital studies indicated that ubiquinol was most stable when the portion of the isoprenoid side chain adjacent to the quinol ring was perpendicular to the quinol-ring plane, because of steric congestion by the vicinal OH and methyl groups. The side chain of plastoquinol was flexible because of the lack of a vicinal methyl group. Thus, the difference in the inhibitory activities between the two systems seemed to reflect the difference in the configuration of the isoprenoid side chain of ubiquinone and plastoquinone. These results suggested that the quinone redox site of the cytochrome-bc1 complex may recognize the configuration of the side chain near the quinone ring in the strict sense, whereas that of photosystem II (QB site) may recognize it in a loose sense. Topics: Animals; Benzoquinones; Binding Sites; Binding, Competitive; Chloroplasts; Dinitrophenols; Electron Transport Complex III; Male; Mitochondria, Liver; Molecular Conformation; Oxidation-Reduction; Photosynthetic Reaction Center Complex Proteins; Photosystem II Protein Complex; Plants; Plastoquinone; Rats; Rats, Wistar; Structure-Activity Relationship; Ubiquinone | 1992 |
Matrix NADH dehydrogenases of plant mitochondria and sites of quinone reduction by complex I.
In order to distinguish the pathways involved in the oxidation of matrix NADH in plant mitochondria, the oxidation of NADH and nicotinamide hypoxanthine dinucleotide (reduced form) was investigated in submitochondrial particles prepared from beetroot (Beta vulgaris L. cv. Derwent Globe) and soybeans (Glycine max L. cv. Bragg). Nicotinamide-hypoxanthine-dinucleotide(reduced form)-oxidase activity was more strongly inhibited by rotenone than the NADH-oxidase activity but both of the rotenone-inhibited activities could be stimulated by adding ubiquinone-1. The corresponding ubiquinone-1-reductase activities were inhibited by rotenone (to 69%) and further inhibited by N,N'-dicyclohexylcarbodiimide (to 79%), whilst the K3Fe(CN)6-reductase activities were not sensitive to either rotenone or N,N'-dicyclohexylcarbodiimide. Immunological analysis of mitochondrial proteins using an antiserum raised against purified beetroot complex I indicated very few differences between soybean and fresh and aged beetroot mitochondria, despite their varying sensitivities to rotenone. We confirm that there are two dehydrogenases capable of oxidising internal NADH and that only one of these, namely complex I, is inhibited by rotenone. Further, we conclude that complex I has two potential sites of quinone reduction, both sensitive to N,N'-dicyclohexycarbodiimide inhibition but only one of which is sensitive to rotenone inhibition. Topics: Benzoquinones; Dicyclohexylcarbodiimide; Glycine max; NAD; NAD(P)H Dehydrogenase (Quinone); NADH Dehydrogenase; Oxidation-Reduction; Plants; Rotenone; Submitochondrial Particles; Ubiquinone | 1992 |
Stimulation of insulin release from pancreatic islets by quinones.
Coenzyme Q (CoQ0) and other quinones were shown to be potent insulin secretagogues in the isolated pancreatic islet. The order of potency was CoQ0 congruent to benzoquinone congruent to hydroquinone-menadione. CoQ6 and CoQ10 (ubiquinone), duroquinone and durohydroquinone did not stimulate insulin release. CoQ0's insulinotropism was enhanced in calcium-free medium and CoQ0 appeared to stimulate only the second phase of insulin release. CoQ0 inhibited inositol mono-, bis- and trisphosphate formation. Inhibitors of mitochondrial respiration (rotenone, antimycin A, FCCP and cyanide) and the calcium channel blocker verapamil, did not inhibit CoQ0-induced insulin release. Dicumarol, an inhibitor of quinone reductase, did not inhibit CoQ0-induced insulin release, but it did inhibit glucose-induced insulin release suggesting that the enzyme and quinones play a role in glucose-induced insulin release. Quinones may stimulate insulin release by mimicking physiologically-occurring quinones, such as CoQ10, by acting on the plasma membrane or in the cytosol. Exogenous quinones may bypass the quinone reductase reaction, as well as many reactions important for exocytosis. Topics: Animals; Benzoquinones; Calcium; Dicumarol; Glucose; Inositol Phosphates; Insulin; Insulin Secretion; Islets of Langerhans; Quinones; Rats; Rats, Inbred Strains; Ubiquinone | 1991 |
Reduction of the Q-pool by duroquinol via the two quinone-binding sites of the QH2: cytochrome c oxidoreductase. A model for the equilibrium between cytochrome b-562 and the Q-pool.
The steady-state reduction of exogenous ubiquinone-2 by duroquinol as catalysed by the ubiquinol: cytochrome c oxidoreductase was studied in bovine heart mitoplasts. The reduction of ubiquinone-2 by duroquinol proceeds both in the absence of inhibitors of the enzyme, in the presence of outside inhibitors, e.g., myxothiazol, and in the presence of inside inhibitors, e.g., antimycin, but not in the presence of both inside and outside inhibitors. It is concluded that both the Qin-binding domain and the Qout-binding domain may independently catalyse this reaction. The rate of the reduction of ubiquinone-2 by duroquinol via the Qin-binding domain is dependent on the type of outside inhibitor used. The maximal rate obtained for the reduction of ubiquinone-2 by DQH2 via the Qout-binding domain, measured in the presence of antimycin, is similar to that catalysed by the Qin-binding domain of the non-inhibited enzyme and depends on the redox state of the high-potential electron carriers of the respiratory chain. The reduction of ubiquinone-2 by DQH2 via the Qin-binding domain can be described by a mechanism in which duroquinol reduces the enzyme, upon which the reduced enzyme is rapidly oxidized by ubiquinone-2 yielding ubiquinol-2. By determination of the initial rate under various conditions and simulation of the time course of reduction of ubiquinone-2 using the integrated form of the steady-state rate equation the values of the various kinetic constants were calculated. During the course of reduction of ubiquinone-2 by duroquinol in the presence of outside inhibitors only cytochrome b-562 becomes reduced. At all stages during the reaction, cytochrome b-562 is in equilibrium with the redox potential of the ubiquinone-2/ubiquinol-2 couple but not with that of the duroquinone/duroquinol couple. At low pH values, cytochrome b-562 is reduced in a single phase; at high pH separate reduction phases are observed. In the absence of inhibitors three reduction phases of cytochrome b-562 are discernible at low pH values and two at high pH values. In the presence of antimyin cytochrome b becomes reduced in two phases. Cytochrome b-562 is reduced in the first phase and cytochrome b-566 in the second phase after substantial reduction of ubiquinone-2 to ubiquinol-2 has occurred. In ubiquinone-10 depleted preparations, titration of cytochrome b-562, in the presence of myxothiazol, with the duroquinone/duroquinol redox couple yields a value of napp = 2, both at low and high pH.( Topics: Animals; Antimycin A; Benzoquinones; Binding Sites; Cattle; Cytochrome b Group; Electron Transport; Electron Transport Complex III; Hydrogen-Ion Concentration; Hydroquinones; Methacrylates; Myocardium; Oxidation-Reduction; Thiazoles; Ubiquinone | 1991 |
Probing the secondary quinone (QB) environment in photosynthetic bacterial reaction centers by light-induced FTIR difference spectroscopy.
The photoreduction of the secondary electron acceptor, QB, has been characterized by light-induced Fourier transform infrared difference spectroscopy of Rb. sphaeroides and Rp. viridis reaction centers. The reaction centers were supplemented with ubiquinone (UQ10 or UQ0). The QB- state was generated either by continuous illumination at very low intensity or by single flash in the presence of redox compounds which rapidly reduce the photooxidized primary electron donor P+. This approach yields spectra free from P and P+ contributions making possible the study of the microenvironment of QB and QB-. Assignments are proposed for the C...O vibration of QB- and tentatively for the C = O and C = C vibrations of QB. Topics: Benzoquinones; Fourier Analysis; Oxidation-Reduction; Photic Stimulation; Photosynthetic Reaction Center Complex Proteins; Rhodobacter sphaeroides; Spectrophotometry, Infrared; Ubiquinone | 1991 |
Effect of substituents of the benzoquinone ring on electron-transfer activities of ubiquinone derivatives.
The effect of substituents on the 1,4-benzoquinone ring of ubiquinone on its electron-transfer activity in the bovine heart mitochondrial succinate-cytochrome c reductase region is studied by using synthetic ubiquinone derivatives that have a decyl (or geranyl) side-chain at the 6-position and various arrangements of methyl, methoxy and hydrogen in the 2, 3 and 5 positions of the benzoquinone ring. The reduction of quinone derivatives by succinate is measured with succinate-ubiquinone reductase and with succinate-cytochrome c reductase. Oxidation of quinol derivatives is measured with ubiquinol-cytochrome c reductase. The electron-transfer efficacy of quinone derivatives is compared to that of 2,3-dimethoxy-5-methyl-6-decyl-1,4-benzoquinone. When quinone derivatives are used as the electron acceptor for succinate-ubiquinone reductase, the methyl group at the 5-position is less important than are the methoxy groups at the 2- and 3-positions. Replacing the 5-methyl group with hydrogen causes a slight increase in activity. However, replacing one or both of 2- and 3-methoxy groups with a methyl completely abolishes electron-acceptor activity. Replacing the 3-methoxy group with hydrogen results in a complete loss of electron-acceptor activity, while replacing the 2-methoxy with hydrogen results in an activity decrease by 70%, suggesting that the methoxy group at the 3-position is more specific than that at the 2-position. The structural requirements for quinol derivatives to be oxidized by ubiquinol-cytochrome c reductase are less strict. All 1,4-benzoquinol derivatives examined show partial activity when used as electron donors for ubiquinol-cytochrome c reductase. Derivatives that possess one unsubstituted position at 2, 3 or 5, with a decyl group at the 6-position, show substrate inhibition at high concentrations. Such substrate inhibition is not observed when fully substituted derivatives are used. The structural requirements for quinone derivatives to be reduced by succinate-cytochrome c reductase are less specific than those for succinate-ubiquinone reductase. Replacing one or both of the 2- and 3-methoxy groups with a methyl and keeping the 5-position unsubstituted (plastoquinone derivatives) yields derivatives with no acceptor activity for succinate-Q reductase. However, these derivatives are reducible by succinate in the presence of succinate-cytochrome c reductase. This reduction is antimycin-sensitive and requires endogenous ubiquinone, suggesting t Topics: Benzoquinones; Electron Transport; Electron Transport Complex II; Electron Transport Complex III; Kinetics; Micelles; Mitochondria; Molecular Structure; Multienzyme Complexes; Oxidation-Reduction; Oxidoreductases; Plastoquinone; Quinones; Succinate Cytochrome c Oxidoreductase; Succinate Dehydrogenase; Ubiquinone | 1990 |
Stimulation of nerve growth factor synthesis/secretion by 1,4-benzoquinone and its derivatives in cultured mouse astroglial cells.
Previously we reported that astroglial cells cultured from mouse brain synthesize and secrete nerve growth factor (NGF) and that, in quiescent cells, catecholamines markedly increase the NGF content in the conditioned medium (CM). We wished to further assess the structural properties required for exhibition of such effect of compounds containing a ring structure analogous to that of catechol on astroglial NGF synthesis. During our study, we found that hydroquinone, which was confirmed not to stimulate NGF synthesis in mouse fibroblast cells in another of our investigations, is a potent stimulator of NGF synthesis in astroglial cells and that 1,4-benzoquinone, an oxidized form of hydroquinone, is a more effective stimulator than hydroquinone itself. In addition, the results of experiments with 1,2-benzoquinone derivatives indicated that the presence of a long aliphatic side chain in the molecule eliminates the stimulatory effect of 1,4-benzoquinone on NGF synthesis in astroglial cells. Topics: Animals; Astrocytes; Benzoquinones; Brain; Catechols; Cells, Cultured; Hydroquinones; Mice; Mice, Inbred ICR; Molecular Structure; Nerve Growth Factors; Quinones; Resorcinols; Structure-Activity Relationship; Ubiquinone | 1990 |
The Rhodospirillum rubrum cytochrome bc1 complex: peptide composition, prosthetic group content and quinone binding.
A cytochrome bc1 complex, essentially free of bacteriochlorophyll, has been purified from the photosynthetic purple non-sulfur bacterium Rhodospirillum rubrum. The complex catalyzes electron flow from quinol to cytochrome c (turnover number = 75 s-1) that is inhibited by low concentrations of antimycin A and myxothiazol. The complex contains only three peptide subunits: cytochrome b (Mr = 35,000); cytochrome c1 (Mr = 31,000) and the Rieske iron-sulfur protein (Mr = 22,400). Em values (pH 7.4) were measured for cytochrome c1 (+320 mV) and the two hemes of cytochrome b (-33 and -90 mV). Electron flow from quinol to cytochrome c is inhibited when the complex is pre-illuminated in the presence of a ubiquinone photoaffinity analog (azido-Q). During illumination, the azido-Q becomes covalently attached to the cytochrome b peptide and, to a lesser extent, to cytochrome c1. Topics: Affinity Labels; Azides; Benzoquinones; Cytochrome b Group; Cytochrome c Group; Electron Transport; Electron Transport Complex III; Electrophoresis, Polyacrylamide Gel; Iron-Sulfur Proteins; Molecular Weight; Peptides; Photochemistry; Quinones; Rhodospirillum rubrum; Spectrophotometry; Ubiquinone | 1989 |
Discrete catalytic sites for quinone in the ubiquinol-cytochrome c2 oxidoreductase of Rhodopseudomonas capsulata. Evidence from a mutant defective in ubiquinol oxidation.
A non-photosynthetic mutant (Ps-) of Rhodopseudomonas capsulata, designated R126, was analyzed for a defect in the cyclic electron transfer system. Compared to a Ps+ strain MR126, the mutant was shown to have a full complement of electron transfer components (reaction centers, ubiquinone-10, cytochromes b, c1, and c2, the Rieske 2-iron, 2-sulfur (Rieske FeS) center, and the antimycin-sensitive semiquinone). Functionally, mutant R126 failed to catalyze complete cytochrome c1 + c2 re-reduction or cytochrome b reduction following a short (10 microseconds) flash of actinic light. Evidence (from flash-induced carotenoid band shift) was characteristic of inhibition of electron transfer proximal to cytochrome c1 of the ubiquinol-cytochrome c2 oxidoreductase. Three lines of evidence indicate that the lesion of R126 disrupts electron transfer from quinol to Rieske FeS: 1) the degree of cytochrome c1 + c2 re-reduction following a flash is indicative of electron transfer from Rieske FeS to cytochrome c1 + c2 without redox equilibration with an additional electron from a quinol; 2) inhibitors that act at the Qz site and raise the Rieske FeS midpoint redox potential (Em), namely 5-undecyl-6-hydroxy-4,7-dioxobenzothiazole or 3-alkyl-2-hydroxy-1,4-napthoquinone, have no effect on cytochrome c1 + c2 oxidation in R126; 3) the Rieske FeS center, although it exhibits normal redox behavior, is unable to report the redox state of the quinone pool, as metered by its EPR line shape properties. Flash-induced proton binding in R126 is indicative of normal functional primary (QA) and secondary (QB) electron acceptor activity of the photosynthetic reaction center. The Qc functional site of cytochrome bc1 is intact in R126 as measured by the existence of antimycin-sensitive, flash-induced cytochrome b reduction. Topics: Antimycin A; Benzoquinones; Cytochrome c Group; Electron Spin Resonance Spectroscopy; Electron Transport; Electron Transport Complex III; Methacrylates; Multienzyme Complexes; Mutation; Oxidation-Reduction; Photolysis; Quinone Reductases; Quinones; Rhodopseudomonas; Thiazoles; Ubiquinone | 1986 |
Regulation of electron transfer by the quinone pool.
Strong evidence for a random collisional mechanism for ubiquinone-mediated electron transfer is provided by the characteristic kinetic properties of respiratory chains originally explored by Kröger, A., and Klingenberg, M. (1973), Eur. J. Biochem. 34, 313-323. A kinetic model which leads to this so-called "simple Q-pool behavior" has been described and we use this in reviewing evidence that electron transfer is diffusion-controlled as well as diffusion-coupled. We also consider mechanisms by which the kinetics of electron transfer might deviate from simple Q-pool behavior and how these might be implicated in the regulation of electron transport. Topics: Benzoquinones; Diffusion; Electron Transport; Electron Transport Complex III; Kinetics; Mathematics; Membrane Lipids; Membrane Proteins; NAD(P)H Dehydrogenase (Quinone); Quinone Reductases; Quinones; Ubiquinone | 1986 |
Radical-pair energetics and decay mechanisms in reaction centers containing anthraquinones, naphthoquinones or benzoquinones in place of ubiquinone.
In reaction centers from Rhodobacter sphaeroides (formerly called Rhodopseudomonas sphaeroides), light causes an electron-transfer reaction that forms the radical pair state (P+I-, or PF) from the initial excited singlet state (P) of a bacteriochlorophyll dimer (P). Subsequent electron transfer to a quinone (Q) produces the state P+Q-. Back electron transfer can regenerate P from P+Q-, giving rise to 'delayed' fluorescence that decays with approximately the same lifetime as P+Q-. The free-energy difference between P+Q- and P can be determined from the initial amplitude of the delayed fluorescence. In the present work, we extracted the native quinone (ubiquinone) from Rps. sphaeroides reaction centers, and replaced it by various anthraquinones, naphthoquinones, and benzoquinones. We found a rough correlation between the halfwave reduction potential (E1/2) of the quinone used for reconstitution (as measured polarographically in dimethylformamide) and the apparent free energy of the state P+Q- relatively to P. As the E1/2 of the quinone becomes more negative, the standard free-energy gap between P+Q- and P decreases. However, the correlation is quantitatively weak. Apparently, the effective midpoint potentials (Em) of the quinones in situ depend subtly on interactions with the protein environment in the reaction center. Using the value of the Em for ubiquinone determined in native reaction centers as a reference, and the standard free energies determined for P+Q- in reaction centers reconstituted with other quinones, the effective Em values of 12 different quinones in situ are estimated. In native reaction centers, or in reaction centers reconstituted with quinones that give a standard free-energy gap of more than about 0.8 eV between P+Q- and P*, charge recombination from P+Q- to the ground state (PQ) occurs almost exclusively by a temperature-insensitive mechanism, presumably electron tunneling. When reaction centers are reconstituted with quinones that give a free-energy gap between P+Q- and P* of less than 0.8 with quinones that give a free-energy gap between P+Q- and P* of less than 0.8 eV, part or all of the decay proceeds through a thermally accessible intermediate. There is a linear relationship between the log of the rate constant for the decay of P+Q- via the intermediate state and the standard free energy of P+Q-. The higher the free energy, the faster the decay. The kinetic and thermodynamic properties of the intermediate appear not to depend str Topics: Anthraquinones; Bacterial Proteins; Benzoquinones; Electron Transport; Energy Transfer; Kinetics; Light-Harvesting Protein Complexes; Mathematics; Models, Chemical; Naphthoquinones; Photosynthetic Reaction Center Complex Proteins; Quinones; Rhodopseudomonas; Ubiquinone | 1986 |
Influence of in vitro ubiquinone antagonists on doxorubicin toxicity in vivo.
Doxorubicin is an anthracycline antibiotic with a very wide spectrum of anticancer activity. It has a great potential for clinical cardiotoxicity, however. One mechanism suggested for the cardiotoxicity is inhibition of ubiquinone-dependent enzymes. It was our purpose to study this possible mechanism using ubiquinone antagonists as probes. The effect on doxorubicin toxicity of three in vitro ubiquinone antagonists was tested in mice. Two of the antagonists, 2-hydroxy-3-n-dodecylmercapto-1,4-naphthoquinone and 2,3-dimethoxy-5-beta-naphthylmercapto-1,4-benzoquinone, enhanced doxorubicin toxicity in vivo as measured by survival. The latter was significantly toxic to mice, by itself. This effect was completely blocked by ubiquinone pretreatment, but only reduced by tocopherol pretreatment. Neither ubiquinone nor tocopherol was able to decrease the toxic interaction between doxorubicin and either of the ubiquinone antagonists. Cardiac and hepatic glutathione reductase and glutathione peroxidase activities were measured in studies using the 2,3-dimethoxy-5-beta-naphthylmercapto-1,4-benzoquinone. This compound appeared to cause a slight reduction in the activity of hepatic glutathione reductase. It appears that these antagonists are not useful to probe the relationship between doxorubicin cardiotoxicity and ubiquinone enzyme inhibition. Topics: Animals; Benzoquinones; Doxorubicin; Female; Glutathione Peroxidase; Glutathione Reductase; Male; Mice; Mice, Inbred BALB C; Mice, Inbred Strains; Naphthoquinones; Quinones; Ubiquinone; Vitamin E | 1986 |
An NADH:quinone oxidoreductase of the halotolerant bacterium Ba1 is specifically dependent on sodium ions.
The rate of NADH oxidation by inverted membrane vesicles prepared from the halotolerant bacterium Ba1 of the Dead Sea is increased specifically by sodium ions, as observed earlier in whole cells. The site of this sodium effect is identified as the NADH: quinone oxidoreductase, similarly to the other such system known, Vibrio alginolyticus (H. Tokuda and T. Unemoto (1984) J. Biol. Chem. 259, 7785-7790). Sodium accelerates quinone reduction severalfold, but oxidation of the quinol, with oxygen as terminal electron acceptor, is unaffected. The sodium-dependent pathway of quinone reduction exhibits higher apparent affinity to extraneous quinone (Q-2) than the sodium-insensitive pathway, and is specifically inhibited by 2-heptyl-4-hydroxyquinoline N-oxide. ESR spectra of the membranes contain a feature at g = 1.98 which is tentatively identified as one originating from semiquinone. This feature is increased by NADH and decreased by addition of Na+, suggesting that, as proposed from different kinds of evidence for the V. alginolyticus system, sodium affects the semiquinone reduction step. As in the other system, the site of sodium stimulation in Ba1 probably corresponds to the site of sodium translocation, which was shown earlier (S. Ken-Dror, R. Shnaiderman, and Y. Avi-Dor (1984) Arch. Biochem. Biophys. 229, 640-649) to be linked directly to a redox reaction in the respiratory chain. Topics: Bacteria; Benzoquinones; Cytochrome b Group; Electron Transport; Hydrogen-Ion Concentration; Hydroxyquinolines; NAD; Oxidation-Reduction; Quinone Reductases; Quinones; Sodium; Succinates; Succinic Acid; Ubiquinone; Vibrio | 1986 |
[Intermediators of the Krebs cycle and electron acceptor systems as new cardiac agents].
Topics: Animals; Benzoquinones; Cardiovascular Agents; Citric Acid Cycle; Coronary Circulation; Coronary Disease; Electron Transport; Hypoxia; Malates; Mice; Myocardium; Quinones; Rats; Succinates; Ubiquinone | 1982 |
BIOSYNTHESIS OF THE BENZOQUINONE RING OF UBIQUINONE IN TETRAHYMENA PYRIFORMIS.
Topics: Benzoquinones; Carbon Isotopes; Glucose; Metabolism; Phenylalanine; Quinones; Research; Tetrahymena; Tetrahymena pyriformis; Ubiquinone | 1965 |
THE UBIQUINONE-LIKE ACTIVITY OF NATURALLY OCCURRING BENZOQUINONE DERIVATIVES.
Topics: Animals; Benzoquinones; Electron Transport Complex II; Manometry; Mitochondria; Myocardium; Pharmacology; Quinones; Research; Succinate Dehydrogenase; Swine; Ubiquinone | 1964 |
THE BIOSYNTHESIS OF THE BENZOQUINONE RING OF UBIQUINONE FROM P-HYDROXYBENZALDEHYDE AND P-HYDROXYBENZOIC ACID IN RAT KIDNEY, AZOTOBACTER VINELANDII, AND BAKER'S YEAST.
Topics: Aldehydes; Azotobacter; Azotobacter vinelandii; Benzaldehydes; Benzoates; Benzoquinones; Carbon Isotopes; Chromatography; Hydroxybenzoates; Kidney; Metabolism; Pharmacology; Rats; Research; Saccharomyces; Saccharomyces cerevisiae; Tyrosine; Ubiquinone | 1964 |
THE CONVERSION OF P-HYDROXYBENZALDEHYDE TO THE BENZOQUINONE RING OF UBIQUINONE IN RHODOSPIRILLUM RUBRUM.
Topics: Aldehydes; Benzaldehydes; Benzoquinones; Metabolism; Research; Rhodospirillum; Rhodospirillum rubrum; Ubiquinone | 1963 |
Kofler's quinone and ubiquinone assayed for vitamin K activity by the curative blood clotting method.
Topics: Antifibrinolytic Agents; Benzoquinones; Blood Coagulation; Humans; Naphthoquinones; Quinones; Ubiquinone; Vitamin K | 1962 |