首页 | 官方网站   微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 46 毫秒
1.
The assay of Complex I activity requires the use of artificial acceptors, such as short-chain coenzyme Q homologs and analogs, because the physiological quinones, such as CoQ10, are too insoluble in water to be added as substrates to the assay media. The medical interest raised in the last years on the pathological changes of Complex I activity has focussed on the requirement of easy reliable assays for its analysis. We have undertaken a systematic examination of the assay conditions of Complex I in mitochondrial membranes, using a series of quinones as electron acceptors, particularly the coenzyme Q homologs CoQ0, CoQ1 and CoQ2, and the analogs duroquinone and decylubiquinone. Our findings have pointed out that the most suitable electron acceptor for the NADH:CoQ reductase assay is the homolog CoQ1. The analog DB, commercially available, although yielding a high activity, nevertheless causes some problems for the standardization of the assay conditions.  相似文献   

2.
The human mitochondrial ND1/3460 mutation changes Ala52 to Thr in the ND1 subunit of Complex I, and causes Leber's hereditary optic neuropathy (LHON) [Huoponen et al. (1991) Am. J. Hum. Genet. 48, 1147]. We have used a bacterial counterpart of Complex I, NDH-1 from Paracoccus denitrificans, for studying the effect of mutations in the ND1 subunit on the enzymatic activity. The LHON mutation as well as several other mutations in strictly conserved amino acids in its vicinity were introduced into the NQO8 subunit of NDH-1, a bacterial homologue of ND1. The enzymatic activity of the mutants in the presence of hexammineruthenium (rotenone-insensitive) and ubiquinone-1 (rotenone-sensitive) were assayed. In addition, the kinetics of the interaction of selected mutant enzymes with ubiquinone-1, ubiquinone-2, and decylubiquinone was studied. The results suggest that the mutated residues play an important role in ubiquinone reduction by Complex I.  相似文献   

3.
The reduction kinetics of coenzyme Q (CoQ, ubiquinone) by NADH:ubiquinone oxidoreductase (complex I, EC 1.6.99.3) was investigated in bovine heart mitochondrial membranes using water-soluble homologs and analogs of the endogenous ubiquinone acceptor CoQ10 [the lower homologs from CoQ0 to CoQ3, the 6-pentyl (PB) and 6-decyl (DB) analogs, and duroquinone]. By far the best substrates in bovine heart submitochondrial particles are CoQ1 and PB. The kinetics of NADH-CoQ reductase was investigated in detail using CoQ1 and PB as acceptors. The kinetic pattern follows a ping-pong mechanism; the Km for CoQ1 is in the range of 20 microM but is reversibly increased to 60 microM by extraction of the endogenous CoQ10. The increased Km in CoQ10-depleted membranes indicates that endogenous ubiquinone not only does not exert significant product inhibition but rather is required for the appropriate structure of the acceptor site. The much lower Vmax with CoQ2 but not with DB as acceptor, associated with an almost identical Km, suggests that the sites for endogenous ubiquinone bind 6-isoprenyl- and 6-alkylubiquinones with similar affinity, but the mode of electron transfer is less efficient with CoQ2. The Kmin (kcat/Km) for CoQ1 is 4 orders of magnitude lower than the bimolecular collisional constant calculated from fluorescence quenching of membrane probes; moreover, the activation energy calculated from Arrhenius plots of kmin is much higher than that of the collisional quenching constants. These observations strongly suggest that the interaction of the exogenous quinones with the enzyme is not diffusion-controlled. Contrary to other systems, in bovine submitochondrial particles, CoQ1 usually appears to be able to support a rate approaching that of endogenous CoQ10, as shown by application of the "pool equation" [Kr?ger, A., & Klingenberg, M. (1973) Eur. J. Biochem. 39, 313-323] relating the rate of ubiquinone reduction to the rate of ubiquinol oxidation and the overall rate through the ubiquinone pool.  相似文献   

4.
According to the 'mitochondrial theory of aging' it is expected that the activity of NADH Coenzyme Q reductase (Complex I) would be most severely affected among mitochondrial enzymes, since mitochondrial DNA encodes for 7 subunits of this enzyme. Being these subunits the site of binding of the acceptor substrate (Coenzyme Q) and of most inhibitors of the enzyme, it is also expected that subtle kinetic changes of quinone affinity and enzyme inhibition could develop in aging before an overall loss of activity would be observed. The overall activity of Complex I was decreased in several tissues from aged rats, nevertheless it was found that direct assay of Complex I using artificial quinone acceptors may underevaluate the enzyme activity. The most acceptable results could be obtained by applying the 'pool equation' to calculate Complex I activity from aerobic NADH oxidation; using this method it was found that the decrease in Complex I activity in mitochondria from old animals was greater than the activity calculated by direct assay of NADH Coenzyme Q reductase. A decrease of NADH oxidation and its rotenone sensitivity was observed in nonsynaptic mitochondria, but not in synaptic 'light' and 'heavy' mitochondria of brain cortex from aged rats. In a study of Complex I activity in human platelet membranes we found that the enzyme activity was unchanged but the titre for half-inhibition by rotenone was significantly increased in aged individuals and proposed this change as a suitable biomarker of aging and age-related diseases.  相似文献   

5.
The M(r) 30,000 polypeptide of the hydrophobic protein fraction of the energy-transducing NADH-ubiquinone oxidoreductase (complex I) of bovine heart mitochondria was identified as the ND2 gene product based on a comparison of amino acid analysis and partial N-terminal sequencing results with the known DNA sequence of ND2 (Anderson, S. et al. (1982) J. Mol. Biol. 156, 683-717). A simple purification procedure was devised for this ND2 gene product. The procedure, which is described, involves treatment of bovine complex I with a chloroform-methanol (2:1 [v/v]) solution. The antiserum raised against this purified bovine ND2 gene product cross-reacted with the approximately M(r) 39,000 polypeptide extracted from the Paracoccus denitrificans membranes with chloroform-methanol (2:1 [v/v]).  相似文献   

6.
2-tert-Butyl-1,4-benzoquinone (BQ), a metabolite of the rodent carcinogen 3-tert-butyl-4-hydroxyanisole (3-BHA), has been shown previously to have initiating activity for cell transformation. In this paper, we examined the initiating activity of quinones in a two-stage transformation assay using BALB/3T3 cells. Cells were treated first with a quinone and then with the tumor promoter 12-O-tetra-decanoylphorbol-13-acetate (TPA). The quinones tested were 1,4-benzoquinone (pQ), phenyl-1,4-benzoquinone (PhQ), menadione and 2,5-di-tert-butyl-1,4-benzoquinone (DBQ) in addition to BQ. pQ is a metabolite of benzene and phenacetin, and PhQ is a metabolite of o-phenylphenol (OPP) and sodium o-phenylphenate (OPP-Na). All of the tested quinones induced transformation in the presence of TPA but not in its absence. The extent of transformation caused by quinones followed by TPA was weak but statistically significant. Thus these quinones were shown to act as initiators in the transformation of BALB/3T3 cells. This result suggests that BQ, pQ and PhQ may be involved in carcinogenesis by 3-BHA, benzene, phenacetin, OPP and OPP-Na in vivo. Menadione has been reported to cause cytotoxic effects and mutations through active oxygen generation from semiquinone radicals. DBQ has two bulky substitutes which interfere with covalent bonds with DNA. Menadione and DBQ exhibited initiating activity in the present study. This result suggests that active oxygens generated from semiquinone radicals may play a role in the initiation of cell transformation.  相似文献   

7.
The plastid genomes of several plants contain ndh genes-homologues of genes encoding subunits of the proton-pumping NADH:ubiquinone oxidoreductase, or complex I, involved in respiration in mitochondria and eubacteria. From sequence similarities with these genes, the ndh gene products have been suggested to form a large protein complex (Ndh complex); however, the structure and function of this complex remains to be established. Herein we report the isolation of the Ndh complex from the chloroplasts of the higher plant Pisum sativum. The purification procedure involved selective solubilization of the thylakoid membrane with dodecyl maltoside, followed by two anion-exchange chromatography steps and one size-exclusion chromatography step. The isolated Ndh complex has an apparent total molecular mass of approximately 550 kDa and according to SDS/PAGE consists of at least 16 subunits including NdhA, NdhI, NdhJ, NdhK, and NdhH, which were identified by N-terminal sequencing and immunoblotting. The Ndh complex showed an NADH- and deamino-NADH-specific dehydrogenase activity, characteristic of complex I, when either ferricyanide or the quinones menadione and duroquinone were used as electron acceptors. This study describes the isolation of the chloroplast analogue of the respiratory complex I and provides direct evidence for the function of the plastid Ndh complex as an NADH:plastoquinone oxidoreductase. Our results are compatible with a dual role for the Ndh complex in the chlororespiratory and cyclic photophosphorylation pathways.  相似文献   

8.
X Yang  L Yu  D He  CA Yu 《Canadian Metallurgical Quarterly》1998,273(48):31916-31923
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.  相似文献   

9.
A combination of structural, thermodynamic, and transient kinetic data on wild-type and mutant Anabaena vegetative cell ferredoxins has been used to investigate the nature of the protein-protein interactions leading to electron transfer from reduced ferredoxin to oxidized ferredoxin:NADP+ reductase (FNR). We have determined the reduction potentials of wild-type vegetative ferredoxin, heterocyst ferredoxin, and 12 site-specific mutants at seven surface residues of vegetative ferredoxin, as well as the one- and two-electron reduction potentials of FNR, both alone and in complexes with wild-type and three mutant ferredoxins. X-ray crystallographic structure determinations have been carried out for six of the ferredoxin mutants. None of the mutants showed significant structural changes in the immediate vicinity of the [2Fe-2S] cluster, despite large decreases in electron-transfer reactivity (for E94K and S47A) and sizable increases in reduction potential (80 mV for E94K and 47 mV for S47A). Furthermore, the relatively small changes in Calpha backbone atom positions which were observed in these mutants do not correlate with the kinetic and thermodynamic properties. In sharp contrast to the S47A mutant, S47T retains electron-transfer activity, and its reduction potential is 100 mV more negative than that of the S47A mutant, implicating the importance of the hydrogen bond which exists between the side chain hydroxyl group of S47 and the side chain carboxyl oxygen of E94. Other ferredoxin mutations that alter both reduction potential and electron-transfer reactivity are E94Q, F65A, and F65I, whereas D62K, D68K, Q70K, E94D, and F65Y have reduction potentials and electron-transfer reactivity that are similar to those of wild-type ferredoxin. In electrostatic complexes with recombinant FNR, three of the kinetically impaired ferredoxin mutants, as did wild-type ferredoxin, induced large (approximately 40 mV) positive shifts in the reduction potential of the flavoprotein, thereby making electron transfer thermodynamically feasible. On the basis of these observations, we conclude that nonconservative mutations of three critical residues (S47, F65, and E94) on the surface of ferredoxin have large parallel effects on both the reduction potential and the electron-transfer reactivity of the [2Fe-2S] cluster and that the reduction potential changes are not the principal factor governing electron-transfer reactivity. Rather, the kinetic properties are most likely controlled by the specific orientations of the proteins within the transient electron-transfer complex.  相似文献   

10.
To determine the environment of different positions in the paclitaxel molecule when bound to tubulin, we have synthesized six fluorescent analogues in which a (dimethylamino)benzoyl group has been introduced into the 7- and 10-positions, and the benzoyl groups at the 2- and N- as well as the 3'-phenyl ring have been modified with dimethylamino functions. In a tubulin assembly assay, the N-m- and N-p-(dimethylamino)benzoyl derivatives had activities comparable to the activity of paclitaxel. The 2-, 3'-, and 10-analogues had slightly reduced activity, and the 7-derivative was about 5% as active as paclitaxel. On the basis of the results of studies of the effect of solvents on the fluorescence emission spectra, it is proposed that the unbound analogues form hydrogen bonds with protic solvents. But the 7- and 10-substituted analogues appear to be more affected by protic solvents than the other analogues. Previously, we studied the binding of the N-meta derivative to tubulin and microtubules [Sengupta, S., et al. (1995) Biochemistry 34, 11889-11894]. In this study, we extended the studies to include the 2-, 7-, and 10-derivatives. Similar to the N-substituted analogue, binding of the 2-derivative to tubulin was accompanied by a large blue shift, whereas a very small shift occurred when the 7- and 10-substituted derivatives bound. The 2- and N-substituted analogues bind to microtubules with an increase in fluorescence intensity over that which was observed with tubulin, whereas binding of the 7- and 10-substituted analogues was accompanied by a large quenching in fluorescence. This quenching may be due to the presence of charged residues in the protein near the 7- and 10-(dimethylamino)benzoyl groups or to pi stacking of the groups with an aromatic side chain. The presence of paclitaxel with microtubules prevented the fluorescence increase of the 2- and N-derivatives and quenching of the 7- and 10-derivatives. The difference in behavior of the fluorescent analogues upon binding to polymerized tubulin, coupled with the solvent studies on the free drugs, suggests that the 2- and N-benzoyl groups of paclitaxel bind in a hydrophobic pocket of tubulin but could participate in hydrogen bonding, and the 7- and 10-positions are in a more hydrophilic environment.  相似文献   

11.
Based on the DNA sequence of its structural genes, clustered in the hnd operon, the NADP-reducing hydrogenase of Desulfovibrio fructosovorans is thought to be a heterotetrameric complex in which HndA and HndC constitute the NADP-reducing unit and HndD constitutes the hydrogenase unit, respectively. The weak representativity of the enzyme among cell proteins has prevented its purification. This paper discusses the purification and characterization of the HndA subunit of this unique tetrameric iron hydrogenase overproduced in Escherichia coli. The purified subunit contains 1.7 mol of non-heme iron and 1.7 mol of acid-labile sulfide/mol. EPR analysis of the reduced form of HndA indicates that it contains a single binuclear [2Fe-2S] cluster. This cluster exhibits a spectrum of rhombic symmetry with values of gx, gy, and gz equal to 1.915, 1.950, and 2. 000, respectively, and a midpoint redox potential of -395 mV. The UV-visible and EPR spectra of the [2Fe-2S] cluster indicate that HndA belongs to the [2Fe-2S] family typified by the Clostridium pasteurianum [2Fe-2S] ferredoxin. The C-terminal sequence of HndA shows 27% identity with the C-terminal sequence of the 25-kDa subunit of NADH: quinone oxidoreductase from Paracoccus denitrificans, 33% identity with the C-terminal sequence of the 24-kDa subunit from Bos taurus complex I, and 32% identity with the entire sequence of C. pasteurianum [2Fe-2S] ferredoxin. The four cysteine residues involved in HndA cluster binding have been tentatively identified on the basis of sequence identity considerations. Evidence of a HndA organization based on two independent structural domains is discussed.  相似文献   

12.
The redox capacities of p-benzoquinone (I) and its analogs p-benzoquinone imine (VI) and p-benzoquinone diimine (XI) as the simplest model systems for the biochemically important quinone site of the pharmacophores of the anthracyclines has been investigated by AM1 semi-empirical and ab initio methods. The reductive activation of the parent (Q) model systems to their various redox states (quinone radical anion (Q.-), semiquinone (QH.), semiquinone anion (QH-) and hydroquinone (QH2)), the internal geometrical reorganization and the redox capacities of the redox states have been examined by using energy-partitioning analysis, reaction enthalpies/energies for electron and proton attachments, adiabatic ionization potentials (IPad) and electron affinities (EAad), adiabatic electronegativities (Xad), dipole moments, electrostatic potentials and spin-density surfaces. EAad data and results of energy-partitioning analysis suggest that the one-electron Q to Q.- reducibility of VI is diminished when compared to that of I. The data also predict that reduction to QH., QH- and QH2 is more favorable in VI (cf. I). Deprotonation enthalpy/energy calculations predict that the oxidizability of the reduced forms of VI is diminished when compared to I. Overall, the calculations suggest that the redox cycling of VI should be diminished if deprotonation is the first step of the autoxidation of the reduced forms. The results suggest that the electron affinity of Q and deprotonation of the reduced forms (e.g., QH.) may play important roles in the redox cycling of the anthracyclines. It is further suggested that these same factors are probably responsible for the reduced toxicity of 5-iminodaunomycin, which consists of VI as part of its pharmacophore. A comparison of the AM1 results with ab initio results suggests that the AM1 method is capable of predicting trends in redox capacity, nucleophilicity, electrophilicity and electron affinity in the systems investigated.  相似文献   

13.
Light-induced FTIR difference spectra of the photoreduction of the primary quinone acceptor QA have been obtained for Rhodobacter sphaeroides RCs reconstituted with a series of isotopically labeled quinones in order to separate the contributions of the quinone from those of the protein. The isotopic shifts observed in the QA-/QA spectra of RCs reconstituted with ubiquinones (Q1, Q6) or vitamin K1 18O-labeled on their carbonyl oxygens and with fully 13C-labeled Q8 lead to a clear identification of the quinone bands from both the neutral and anion forms. Double-difference spectra from pairs of QA-/QA spectra obtained from 18O/16O Q6, 18O/16O Q1, 13C/12C Q8, 13C18O/12C16O Q8, and 18O/16O vitamin K1 allow the C = O modes of QA in vivo to be identified unambiguously for the first time. For all the investigated unlabeled quinones, two carbonyl bands are demasked, at 1660 and 1628 cm-1 for neutral ubiquinones and at 1651 and 1640 cm-1 for vitamin K1, while C = C bands are found at 1608 and 1588 cm-1 for vitamin K1 and at 1601 cm-1 for ubiquinones. Compared with the spectra of the isolated quinones, the generally smaller width observed for the C = O and C = C bands in vivo suggests precise interactions between the quinone and the contours of the protein at a single, well-defined QA site. The different frequency downshifts of the two C = O bands upon binding to the QA site underscore the inequivalence of the two carbonyls in providing asymmetrical bonding interactions with the protein. The comparison of the isotopic shifts observed for the various quinone C = O and C = C bands in vitro and in vivo demonstrates that the admixture of C = O and C = C characters in these modes is strongly affected by the binding of QA to its anchoring site. In particular, the bands at 1628 and 1601 cm-1 of Q6 in vivo exhibit highly mixed C = O and C = C characters. In contrast, the methoxy groups of the ubiquinones do not appear to suffer large strain upon binding. The closeness of the QA-/QA spectra for Q1 and Q6 indicates that a possible role of the chain in providing the proper positioning of the quinone ring in the site for both the oxidized and reduced states of QA cannot extend significantly beyond the first isoprene unit.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

14.
A series of substituted 2-(4-methoxyphenyl)-1H-benzimidazoles were synthesized and evaluated as inhibitors of topoisomerase I. The presence of a 5-formyl-, 5-(aminocarbonyl)-, or 5-nitro group (i.e., substituents capable of acting as hydrogen bond acceptors) correlated with the potential of select derivatives to inhibit topoisomerase I. In contrast to bi- and terbenzimidazoles, the substituted benzimidazoles that were active as topoisomerase I poisons exhibited minimum or no DNA binding affinity. 5-Nitro-2-(4-methoxyphenyl)-1H-benzimidazole exhibited the highest activity and was significantly more active than the 4-nitro positional isomer. The 5- and 6-nitro derivatives of 2-(4-methoxyphenyl) benzoxazole, 2-(4-methoxyphenyl)benzothiazole, and 2-(4-methoxyphenyl)indole were synthesized and their relative activity as topoisomerase I inhibitors determined. None of these heterocyclic analogues were effective in significantly inhibiting cleavable-complex formation in the presence of DNA and topoisomerase I, suggesting a high degree of structural specificity associated with the interaction of these substituted benzimidazoles with the enzyme or the enzyme-DNA complex. In evaluating their cytotoxicity, these new topoisomerase I poisons also exhibited no significant cross-resistance against cell lines that express camptothecin-resistant topoisomerase I.  相似文献   

15.
The metals of the cytochrome c oxidase structures of the bovine heart mitochondrion (PDB code 1occ) and of the soil bacterium Paracoccus denitrificans (1arl) include a dicopper center (CuA), magnesium, two proximal hemes, a copper (CuB) atom, and a calcium. The mitochondrial structure also possesses a bound distant zinc ion. The extended environments of the metal sites are analyzed emphasizing residues of the second shell in terms of polarity, hydrophobicity, secondary structure, solvent accessibility, and H-bonding networks. A significant difference in the CuA metal environments concerns D-51 I in 1occ, absent from 1arl. The D-51 I appears to play an important role in the proton pumping pathway. Our analysis uncovers several statistically significant residue clusters, including a cysteine-histidine-tyrosine cluster overlapping the CuA-Mg complex; a histidine-acidic cluster enveloping the environment of Mg, the two hemes, and CuB; and on the protein surface a mixed charge cluster, which may help stabilize the quaternary structure and/or mediate docking to cytochrome c. These clusters may constitute possible pathways for electron transfer, for O2 diffusion, and for H2O movement. Many hydrogen bonding relations along the interface of subunits I and II demarcate this surface as a potential participant in proton pumping.  相似文献   

16.
1-Phenylbenzimidazoles are shown to be a new class of ATP-site inhibitors of the platelet-derived growth factor receptor (PDGFR). Structure-activity relationships (SARs) are narrow, with closely related heterocycles being inactive. A systematic study of substituted 1-phenylbenzimidazoles showed clear SARs. Substituents at the 4'- and 3'-positions of the phenyl ring are tolerated but do not significantly improve activity, while substituents at the 2'-position abolish it. Substituents in the 2-, 4-, and 7-positions of the benzimidazole ring (with the exception of 4-OH) also abolish activity. Most substituents at the 5- and 6-positions maintain or increase activity, with the 5-OH, 5-OMe, 5-COMe, and 5-CO2Me analogues being >10-fold more potent than the parent 1-phenylbenzimidazole. The 5-OMe analogue was both the most potent inhibitor, and showed the highest selectivity (50-fold) between PDGFR and FGFR isolated enzymes, and also a moderately effective inhibitor (IC50 = 1.9 microM) of PDGF-stimulated PDGFR autophosphorylation in rat aorta smooth muscle cells.  相似文献   

17.
3-?4-[2-(Benzoxazol-2-ylmethylamino)ethoxy]phenyl?-(2S)-((2- benzoylph enyl)amino)propionic acid (1) and (2S)-((2-benzoylphenyl)amino)-3-?4-[2-(5-methyl-2-phenyloxazol-4-y l)e thoxy]phenyl?propionic acid (2) are peroxisome proliferator-activated receptor gamma (PPARgamma) agonists and have antidiabetic activity in rodent models of type 2 diabetes. As part of an effort to develop the SAR of the N-2-benzoylphenyl moiety of 1 and 2, a series of novel carboxylic acid analogues, 23-66, modified only in the N-2-benzoylphenyl moiety were synthesized from L-tyrosine and evaluated as PPARgamma agonists. In general, only modest changes in the N-2-benzoylphenyl moiety of 1 and 2 are tolerated. More specifically, the best changes involve bioisosteric replacement of one of the two phenyl rings of this moiety. Addition of substituents to this moiety generally produced compounds that are less active in the cell-based functional assays of PPARgamma activity although binding affinity to PPARgamma may be maintained. A particularly promising set of analogues is the anthranilic acid esters 63-66 in which the phenyl ring in the 2-benzoyl group of 1 and 2 has been replaced by an alkoxy group. In particular, (S)-2-(1-carboxy-2-?4-[2-(5-methyl-2-phenyloxazol-4-yl)ethoxy]phen yl? ethylamino)benzoic acid methyl ester (63) has a pKi of 8.43 in the binding assay using human PPARgamma ligand binding domain and a pEC50 of 9.21 in the in vitro murine lipogenesis functional assay of PPARgamma activity. Finally, 63 was found to normalize glycemia when dosed at 3 mg/kg bid po in the Zucker diabetic fatty rat model of type 2 diabetes.  相似文献   

18.
The binding of TNP-ATP (2' or 3'-O-(2,4,6-trinitrophenyl)-ATP) to cytochrome c oxidase (COX) from bovine heart and liver and to the two-subunit COX of Paracoccus denitrificans was measured by its change of fluorescence. Three binding sites, two with high (dissociation constant Kd = 0.2 microM) and one with lower affinity (Kd = 0.9 microM), were found at COX from bovine heart and liver, while the Paracoccus enzyme showed only one binding site (Kd = 3.6 microM). The binding of [35S]ATP alpha S was measured by equilibrium dialysis and revealed seven binding sites at the heart enzyme (Kd = 7.5 microM) and six at the liver enzyme (Kd = 12 microM). The Paracoccus enzyme had only one binding site (Kd = 16 microM). The effect of variable intraliposomal ATP/ADP ratios, but at constant total concentration of [ATP + ADP] = 5 mM, on the H+/e- stoichiometry of reconstituted COX from bovine heart and liver were studied. Above 98% ATP the H+/e- stoichiometry of the heart enzyme decreased to about half of the value measured at 100% ATP. In contrast, the H+/e-stoichiometry of the liver enzyme was not influenced by the ATP/ADP ratio. It is suggested that high intramitochondrial ATP/ADP ratios, corresponding to low cellular work load, will decrease the efficiency of energy transduction and result in elevated thermogenesis for the maintenance of body temperature.  相似文献   

19.
The radical pair P700.+Q.- (P700 = primary electron donor, Q = quinone acceptor) in native photosystem I and in preparations in which the native acceptor (vitamin K1) is replaced by different quinones is investigated by pulsed EPR spectroscopy. In a two-pulse experiment, the light-induced radical pair causes an out-of-phase electron spin echo, showing an envelope modulation. From the modulation frequency, the dipolar coupling, and therefore the distance between the two cofactors, can be derived. The observation of nearly identical distances of about 25.4 A between P700.+ and Q.- in all preparations investigated here leads to the conclusion that the reconstituted quinones are bound to the native A1 binding pocket. Since the orientation of the reconstituted naphthoquinone relative to the axis joining P700.+ and Q*- differs drastically from that of the native vitamin K1, it cannot be bonded to the protein in the same way as the native acceptor. This implies that the function of A1 as an electron acceptor does not depend on the orientation or hydrogen bonding of the quinone.  相似文献   

20.
1,2-Naphthoquinones, such as beta-lapachone, 4-alkoxy-1,2-naphthoquinones, and tetrahydrofuran-1,2-naphthoquinones, react rapidly with 2-mercaptoethanol in benzene to give 1,4-, 1,2-, 1,3- and 1,6-Michael-type adducts that are formed by the addition of the thiol group to the quinone ring. Menadione (2-methyl-1,4-naphthoquinone) reacts with the thiol reagent very slowly under the same reaction conditions. Although the formation of the adducts can be followed by 1H-NMR, attempts to isolate the adducts failed due to their retroconversion to the starting products. On addition of a Lewis acid, however, the adducts undergo cyclization reactions that give stable derivatives that can be isolated and characterized. Determination of the structures of the derivatives allowed for the identification of the adducts from which they originated. Thus, beta-lapachone and 2,3-dinordunnione underwent 1,4- and 1,2-Michael type additions to the quinone ring, while 4-pentyloxy-1,2-naphthoquinone underwent two simultaneous Michael additions to the quinone ring of the naphthoquinone. Menadione underwent a single 1,3-addition. The alkylation rates of the thiol group of 2-mercaptoethanol by the naphthoquinones parallel the naphthoquinones efficiencies in inducing DNA cleavage through DNA-bound topoisomerase II. These results support our hypothesis that the cytotoxic effect of the naphthoquinones derive, at least in part, from their alkylation of exposed thiol residues on the topoisomerase II-DNA complex.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司    京ICP备09084417号-23

京公网安备 11010802026262号