首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
In order to clarify the poorly understood mechanisms of two-electron reduction of quinones by flavoenzymes, we examined the quinone reductase reactions of a member of a structurally distinct old yellow enzyme family, Enterobacter cloacae PB2 pentaerythritol tetranitrate reductase (PETNR). PETNR catalyzes two-electron reduction of quinones according to a 'ping-pong' scheme. A multiparameter analysis shows that the reactivity of quinones increases with an increase in their single-electron reduction potential and pK(a) of their semiquinones (a three-step (e(-),H(+),e(-)) hydride transfer scheme), or with an increase in their hydride-transfer potential (E(7)(H(-))) (a single-step (H(-)) hydride transfer scheme), and decreases with a decrease in their van der Waals volume. However, the pH-dependence of PETNR reactivity is more consistent with a single-step hydride transfer. A comparison of X-ray data of PETNR, mammalian NAD(P)H : quinone oxidoreductase (NQO1), and Enterobacter cloacae nitroreductase, which reduce quinones in a two-electron way, and their reactivity revealed that PETNR is much less reactive, and much less sensitive to the quinone substrate steric effects than NQO1. This may be attributed to the lack of pi-pi stacking between quinone and the displaced aromatic amino acid in the active center, e.g., with Phe-178' in NQO1.  相似文献   

2.
Enterobacter cloacae NAD(P)H:nitroreductase (NR; EC 1.6.99.7) catalyzes two-electron reduction of a series of quinoidal compounds according to a "ping-pong" scheme, with marked substrate inhibition by quinones. The steady-state catalytic constants (k(cat)) range from 0.1 to 1600s(-1), and bimolecular rate constants (k(cat)/K(m)) range from 10(3) to 10(8)M(-1)s(-1). Quinones, nitroaromatic compounds and competitive to NADH inhibitor dicumarol, quench the flavin mononucleotide (FMN) fluorescence of nitroreductase. The reactivity of NR with single-electron acceptors is consistent with an "outer-sphere" electron transfer model, taking into account high potential of FMN semiquinone/FMNH(-) couple and good solvent accessibility of FMN. However, the single-electron acceptor 1,1(')-dibenzyl-4,4(')-bipyridinium was far less reactive than quinones possessing similar single-electron reduction potentials (E(1)(7)). For all quinoidal compounds except 2-hydroxy-1,4-naphthoquinones, there existed parabolic correlations between the log of rate constants of quinone reduction and their E(1)(7) or hydride-transfer potential (E(7)(Q/QH(-))). Based on pH dependence of rate constants, a single-step hydride transfer seems to be a more feasible quinone reduction mechanism. The reactivities of 2-hydroxy-1,4-naphthoquinones were much higher than expected from their reduction potential. Most probably, their enhanced reactivity was determined by their binding at or close to the binding site of NADH and dicumarol, whereas other quinones used the alternative, currently unidentified binding site.  相似文献   

3.
It is supposed that the main cytotoxicity mechanism of antitumour aziridinyl-substituted benzoquinones is their two-electron reduction to alkylating products by NAD(P)H:quinone oxidoreductase (NQO1, DT-diaphorase, EC 1.6.99.2). However, other possible cytotoxicity mechanisms, e.g., oxidative stress, are studied insufficiently. In the single-electron reduction of quinones including a novel compound RH1 (2,5-diaziridinyl- 3-(hydroxymethyl)-6-methyl-1,4-benzoquinone), by NADPH:cytochrome P-450 reductase (EC 1.6.2.4, P-450R), their reactivity increased with an increase in the redox potential of quinone/semiquinone couple (E(1)7), reaching a limiting value at E(1)7> or =-0.1V. The reactivity of quinones towards NQO1 did not depend on their E(1)7. The cytotoxicity of aziridinyl-unsubstituted quinones in bovine leukemia virus-transformed lamb kidney fibroblasts (line FLK) mimics their reactivity in P-450R-catalyzed reactions, exhibiting a parabolic dependence on their E(1)7. The toxicity of aziridinyl-benzoquinones, although being higher, also followed this trend and did not depend on their reactivity towards NQO1. The action of aziridinylbenzoquinones in FLK cells was accompanied by an increase in lipid peroxidation, their toxicity decreased by desferrioxamine and the antioxidant N,N'-diphenyl-p-phenylene diamine, and potentiated by 1,3-bis-(2-chloroethyl)-1-nitrosourea. The inhibitor of NQO1, dicumarol, protected against the toxicity of aziridinyl-benzoquinones except of 2,5-bis-(2'-hydroxyethylamino)-3,6-diaziridinyl-1,4-benzoquinone (BZQ), which was almost inactive as NQO1 substrate. The same events except the absence of pronounced effect of dicumarol were characteristic in the cytotoxicity of aziridinyl-unsubstituted quinones. These findings indicate that in addition to the activation by NQO1, the oxidative stress presumably initiated by single-electron transferring enzymes may be an important factor in the cytotoxicity of aziridinylbenzoquinones. The information obtained may contribute to the understanding of the molecular mechanisms of aziridinylquinone cytotoxicity and may be useful in the design of future bioreductive drugs.  相似文献   

4.
The physiological function of NAD(P)H:quinone oxidoreductase (NQO1, DT-diaphorase) is to detoxify potentially reactive quinones by direct transfer of two electrons. A similar detoxification role has not been established for its homologue NRH:quinone oxidoreductase 2 (NQO2). Estrogen quinones, including estradiol(E(2))-3,4-Q, generated by estrogen metabolism, are thought to be responsible for estrogen-initiated carcinogenesis. In this investigation, we have shown for the first time that NQO2 catalyzes the reduction of electrophilic estrogen quinones and thereby may act as a detoxification enzyme. ESI and MALDI mass spectrometric binding studies involving E(2)-3,4-Q with NQO2 clearly support the formation of an enzyme-substrate physical complex. The problem of spontaneous reduction of substrate by cofactor, benzyldihydronicotinamide riboside (BNAH), was successfully overcome by taking advantage of the ping-pong mechanism of NQO2 catalysis. The involvement of the enzyme in the reduction of E(2)-3,4-Q was further supported by addition of the inhibitor quercetin to the assay mixture. NQO2 is a newly discovered binding site (MT3) of melatonin. However, addition of melatonin to the assay mixture did not affect the catalytic activity of NQO2. Preliminary kinetic studies show that NQO2 is faster in reducing estrogen quinones than its homologue NQO1. Both UV and liquid chromatography-tandem mass spectrometry assays unequivocally corroborate the reduction of estrogen ortho-quinones by NQO2, indicating that it could be a novel target for prevention of breast cancer initiation.  相似文献   

5.
Prooxidant nitroaromatic and quinoidal compounds possess antimalarial activity, which might be attributed either to their formation of reactive oxygen species or to their inhibition of antioxidant enzyme glutathione reductase (GR, EC 1.6.4.2). We have examined the activity in vitro against Plasmodium falciparum of 24 prooxidant compounds of different structure (nitrobenzenes, nitrofurans, quinones, 1,1'-dibenzyl-4,4'-bipyridinium, and methylene blue), which possess a broad range of single-electron reduction potentials (E(1)(7)) and erythrocyte glutathione reductase inhibition constants (K(i(GR))). For a series of homologous derivatives of 2-(5'-nitrofurylvinyl)quinoline-4-carbonic acid, the relationship between compound K(i(GR)) and concentration causing 50% parasite growth inhibition (IC(50)) was absent. For all the compounds examined in this study, the dependence of IC(50) on their K(i(GR)) was insignificant. In contrast, IC(50) decreased with an increase in E(1)(7) and positive electrostatic charge of aromatic part of molecule (Z): log IC(50) (microM) = -(0.9846 +/- 0.3525) - (7.2850 +/- 1.2340) E(1)(7) (V) - (1.1034 +/- 0.1832) Z (r(2) = 0.8015). The redox cycling activity of nitroaromatic and quinoidal compounds in ferredoxin:NADP(+) reductase-catalyzed reaction and the rate of oxyhemoglobin oxidation in lysed erythrocytes increased with an increase in their E(1)(7) value. Our findings imply that the antiplasmodial activity of nitroaromatic and quinoidal compounds is mainly influenced by their ability to form reactive oxygen species, and much less significantly by the GR inhibition.  相似文献   

6.
Madeo J  Gunner MR 《Biochemistry》2005,44(33):10994-11004
Bacterial reaction centers (RCs) catalyze a series of electron-transfer reactions reducing a neutral quinone to a bound, anionic semiquinone. The dissociation constants and association rates of 13 tailless neutral and anionic benzo- and naphthoquinones for the Q(A) site were measured and compared. The K(d) values for these quinones range from 0.08 to 90 microM. For the eight neutral quinones, including duroquinone (DQ) and 2,3-dimethoxy-5-methyl-1,4-benzoquinone (UQ(0)), the quinone concentration and solvent viscosity dependence of the association rate indicate a second-order rate-determining step. The association rate constants (k(on)) range from 10(5) to 10(7) M(-)(1) s(-)(1). Association and dissociation rate constants were determined at pH values above the hydroxyl pK(a) for five hydroxyl naphthoquinones. These negatively charged compounds are competitive inhibitors for the Q(A) site. While the neutral quinones reach equilibrium in milliseconds, anionic hydroxyl quinones with similar K(d) values take minutes to bind or dissociate. These slow rates are independent of ionic strength, solvent viscosity, and quinone concentration, indicating a first-order rate-limiting step. The anionic semiquinone, formed by forward electron transfer at the Q(A) site, also dissociates slowly. It is not possible to measure the association rate of the unstable semiquinone. However, as the protein creates kinetic barriers for binding and releasing anionic hydroxyl quinones without greatly increasing the affinity relative to neutral quinones, it is suggested that the Q(A) site may do the same for anionic semiquinone. Thus, the slow semiquinone dissociation may not indicate significant thermodynamic stabilization of the reduced species in the Q(A) site.  相似文献   

7.
We aimed to elucidate the role of electronic and structural parameters of nitroaromatic compounds in their two-electron reduction by NAD(P)H:quinone oxidoreductase (NQO1, DT-diaphorase, EC 1.6.99.2). The multiparameter regression analysis shows that the reactivity of nitroaromatic compounds (n=38) increases with an increase in their single-electron reduction potential and the torsion angle between nitrogroup(s) and the aromatic ring. The binding efficiency of nitroaromatics in the active center of NQO1 exerted a less evident role in their reactivity. The reduction of nitroaromatics is characterized by more positive entropies of activation than the reduction of quinones. This points to a less efficient electronic coupling of nitroaromatics with the reduced isoalloxazine ring of FAD, and may explain their lower reactivity as compared to quinones. Another important but poorly understood factor enhancing the reactivity of nitroaromatics is their ability to bind at the dicumarol/quinone binding site in the active center of NQO1.  相似文献   

8.
The toxicity of 2,4,6-trinitrotoluene (TNT), a widespread environmental contaminant, is exerted through its enzymatic redox cycling and/or covalent binding of its reduction products to proteins and DNA. In this study, we examined the possibility of another cytotoxicity mechanism of the amino- and hydroxylamino metabolites of TNT, their flavoenzyme-catalyzed redox cycling. The above compounds acted as redox-cycling substrates for single-electron transferring NADPH:cytochrome P-450 reductase (P-450R) and ferredoxin:NADP(+) reductase (FNR), as well as substrates for the two-electron transferring flavoenzymes rat liver NAD(P)H:quinone oxidoreductase (NQO1) and Enterobacter cloacae NAD(P)H:nitroreductase (NR). Their reactivity in P-450R-, FNR-, and NR-catalyzed reactions increased with an increase in their single-electron reduction potential (E(1)(7)) or the decrease in the enthalpy of free radical formation. The cytotoxicity of the amino- and hydroxylamino metabolites of TNT towards bovine leukemia virus-transformed lamb kidney fibroblasts (line FLK) was partly prevented by the antioxidant N,N'-diphenyl-p-phenylene diamine and desferrioxamine, and potentiated by 1,3-bis-(2-chloroethyl)-1-nitrosourea, thus pointing to the involvement of oxidative stress. In general, their cytotoxicity increased with an increase in their electron accepting properties, or their reactivity towards the single-electron transferring FNR and P-450R. Thus, our data imply that the flavoenzyme-catalyzed redox cycling of amino and hydroxylamino metabolites of TNT may be an important factor in their cytotoxicity.  相似文献   

9.
M S Graige  M L Paddock  G Feher  M Y Okamura 《Biochemistry》1999,38(35):11465-11473
A proton-activated electron transfer (PAET) mechanism, involving a protonated semiquinone intermediate state, had been proposed for the electron-transfer reaction k(2)AB [Q(A)(-)(*)Q(B)(-)(*) + H(+) <--> Q(A)(-)(*)(Q(B)H)(*) --> Q(A)(Q(B)H)(-)] in reaction centers (RCs) from Rhodobacter sphaeroides [Graige, M. S., Paddock, M. L., Bruce, M. L., Feher, G., and Okamura, M. Y. (1996) J. Am. Chem. Soc. 118, 9005-9016]. Confirmation of this mechanism by observing the protonated semiquinone (Q(B)H)(*) had not been possible, presumably because of its low pK(a). By replacing the native Q(10) in the Q(B) site with rhodoquinone (RQ), which has a higher pK(a), we were able to observe the (Q(B)H)(*) state. The pH dependence of the semiquinone optical spectrum gave a pK(a) = 7.3 +/- 0.2. At pH < pK(a), the observed rate for the reaction was constant and attributed to the intrinsic electron-transfer rate from Q(A)(-)(*) to the protonated semiquinone (i.e., k(2)AB = k(ET)(RQ) = 2 x 10(4) s(-)(1)). The rate decreased at pH > pK(a) as predicted by the PAET mechanism in which fast reversible proton transfer precedes rate-limiting electron transfer. Consequently, near pH 7, the proton-transfer rate k(H) > 10(4) s(-)(1). Applying the two step mechanism to RCs containing native Q(10) and taking into account the change in redox potential, we find reasonable values for the fraction of (Q(B)H)(*) congruent with 0.1% (consistent with a pK(a)(Q(10)) of approximately 4.5) and k(ET)(Q(10)) congruent with 10(6) s(-)(1). These results confirm the PAET mechanism in RCs with RQ and give strong support that this mechanism is active in RCs with Q(10) as well.  相似文献   

10.
NAD(P)H:quinone oxidoreductase (NQO1; EC 1.6.99.2) catalyzes a two-electron transfer involved in the protection of cells from reactive oxygen species. These reactive oxygen species are often generated by the one-electron reduction of quinones or quinone analogs. We report here on the previously unreported Fe(III) reduction activity of human NQO1. Under steady state conditions with Fe(III) citrate, the apparent Michaelis-Menten constant (Km(app)) was approximately 0.3 nM and the apparent maximum velocity (Vmax(app)) was 16 U mg(-1). Substrate inhibition was observed above 5 nM. NADH was the electron donor, Km(app)= 340 microM and Vmax(app) = 46 Umg(-1). FAD was also a cofactor with a Km(app) of 3.1 microM and Vmax(app) of 89 U mg(-1). The turnover number for NADH oxidation was 25 s(-1). Possible physiological roles of the Fe(III) reduction by this enzyme are discussed.  相似文献   

11.
The bacterial reaction center (RC) converts light into chemical energy through the reduction of an internal quinone molecule Q(B) to Q(B)H(2). In the native RC, proton transfer is coupled to electron transfer and is not rate-controlling. Consequently, proton transfer is not directly observable, and its rate was unknown. In this work, we present a method for making proton transfer rate-controlling, which enabled us to determine its rate. The imidazole groups of the His-H126 and His-H128 proton donors, located at the entrance of the transfer pathways, were removed by site-directed mutagenesis (His --> Ala). This resulted in a reduction in the observed proton-coupled electron transfer rate [(Q(A)(-)(*)Q(B))Glu(-) + H(+) --> (Q(A)Q(B)(-)(*))GluH], which became rate-controlled by proton uptake to Glu-L212 [Adelroth, P., et al. (2001) Biochemistry 40, 14538-14546]. The proton uptake rate was enhanced (rescued) in a controlled fashion by the addition of imidazole or other amine-containing acids. From the dependence of the observed rate on acid concentration, an apparent second-order rate constant k((2)) for the "rescue" of the rate was determined. k((2)) is a function of the proton transfer rate and the binding of the acid. The dependence of k((2)) on the acid pK(a) (i.e., the proton driving force) was measured over 9 pK(a) units, resulting in a Br?nsted plot that was characteristic of general acid catalysis. The results were fitted to a model that includes the binding (facilitated by electrostatic attraction) of the cationic acid to the RC surface, proton transfer to an intermediate proton acceptor group, and subsequent proton transfer to Glu-L212. A proton transfer rate constant of approximately 10(5) s(-)(1) was determined for transfer from the bound imidazole group to Glu-L212 (over a distance of approximately 20 A). The same method was used to determine a proton transfer rate constant of 2 x 10(4) s(-)(1) for transfer to Q(B)(-)(*). The relatively fast proton transfer rates are explained by the presence of an intermediate acceptor group that breaks the process into sequential proton transfer steps over shorter distances. This study illustrates an approach that could be generally applied to obtain information about the individual rates and energies for proton transfer processes, as well as the pK(a)s of transfer components, in a variety of proton translocating systems.  相似文献   

12.
A procedure for the preparation of the fully reduced Cu(I) form of galactose oxidase, GOase(red), involving reduction of GOase(semi) (or GOase(ox)) with non-coordinating [Ru(NH(3))(6)](2+) (51 mV vs. nhe) is described. Air-free conditions and a two-fold excess of [Ru(NH(3))(6)](2+) give a stable product with no further UV-Vis changes over >1.5 h. Rate constants for the reduction of GOase(semi) (k(f)=860 M(-1) s(-1)) give a first-order [H(+)]-dependence (pK(1a)=7.9), but the reverse process involving [Ru(NH(3))(6)](3+) oxidation of GOase(red) (k(b)=18.6 M(-1) s(-1)) is independent of pH (5.5 to 9.5). The reduction potential E(2)(o)' (vs. nhe) for the GOase(semi)/GOase(red) (i.e. Cu(II)/Cu(I)) couple is 149 mV at pH 7.5, which varies from 160 mV (pH 5.5) to 120 mV (pH 10.5), suggesting pK(1a) (GOase(semi)) and pK(2a) (GOase(red)) acid dissociation constants both involving Tyr-495. It is concluded that pK(2a) is for acid dissociation of uncoordinated H(+)Tyr-495. Consistent with this interpretation rate constants/M(-1) s(-1) for the GOase(semi) Tyr495 Phe variant, k(f)=1.59x10(3) and k(b)=16.1, respectively, are independent of pH and give a reduction potential of 169 mV. Comparisons are made of reduction potentials (E(1)(o)'/mV pH 7.5) for the GOase(ox)/GOase(semi) (i.e. Tyr(.)/Tyr) couple, and are for the Cys228Gly variant (630), for enzyme with N(3)(-) for H(2)O at the substrate binding exogenous site (393), and for apo-protein (570). These compare with previously reported values for the variants Trp290His (730) and Tyr495Phe (450), and together serve to quantify different contributions to the unusually small E(1)(o)' of 400 mV for the Tyr(.)/Tyr couple. At pH 7.5 the reduction potential for the two-equivalent GOase(ox)/GOase(red) couple is calculated to be 275 mV. The rate constant for the reaction of GOase(red) with GOase(ox) is 4.4x10(3) M(-1) s(-1) at pH 7.5.  相似文献   

13.
Ishikita H  Knapp EW 《FEBS letters》2006,580(18):4567-4570
To elucidate the role of the non-heme iron complex (Fe-complex) in the electron transfer (ET) events of bacterial photosynthetic reaction centers (bRC), we calculated redox potentials of primary/secondary quinones Q(A/B) (E(m)(Q(A/B))) in the Fe-depleted bRC. Removing the Fe-complex, the calculated E(m)(Q(A/B)) are downshifted by approximately 220 mV/ approximately 80 mV explaining both the 15-fold decrease in ET rate from bacteriopheophytin (H(A)(-)) to Q(A) and triplet state occurrence in Fe-depleted bRC. The larger downshift in E(m)(Q(A)) relative to E(m)(Q(B)) increases the driving-energy for ET from Q(A) to Q(B) by 140 meV, in agreement with approximately 100 meV increase derived from kinetic studies.  相似文献   

14.
One- and two-electron reduction of quinones by glutathione reductase   总被引:1,自引:0,他引:1  
Yeast glutathione reductase (E.C. 1.6.4.2) catalyzes the oxidation of NADPH by p-quinones and ferricyanide with a maximal turnover number (TNmax) of 4-5 s-1.NADP+ stimulates the reaction and the TNmax/Km value of acceptors is reached at NADP+/NADPH greater than or equal to 100. TNmax is increased up to 30-33 s-1. The stimulatory effect of NADP+ may be associated with its complexation with the NADPH-binding site in the reduced enzyme (Kd = 40-60 microM). It is suggested that NADP+ shifts the electron density towards FAD in the two-electron-reduced enzyme and, evidently, changes its one-electron-reduction potentials, while quinones oxidize an equilibrium form of glutathione reductase containing reduced FAD. In the absence of NADP+ the reduction of quinones by glutathione reductase proceeds mainly in a two-electron manner. At NADP+/NADPH = 100 a one-electron reduction makes up 44% of the total process. At pH 6.0-7.0 the reduced forms of naphthoquinones undergo cyclic redox conversions. A hyperbolic dependence exists of the log TN/Km of quinones on their one-electron-reduction potentials.  相似文献   

15.
Ishikita H  Knapp EW 《Biochemistry》2005,44(45):14772-14783
In photosystem II (PSII), the redox properties of the non-heme iron complex (Fe complex) are sensitive to the redox state of quinones (Q(A/)(B)), which may relate to the electron/proton transfer. We calculated the redox potentials for one-electron oxidation of the Fe complex in PSII [E(m)(Fe)] based on the reference value E(m)(Fe) = +400 mV at pH 7 in the Q(A)(0)Q(B)(0) state, considering the protein environment in atomic detail and the associated changes in protonation pattern. Our model yields the pH dependence of E(m)(Fe) with -60 mV/pH as observed in experimental redox titration. We observed significant deprotonation at D1-Glu244 in the hydrophilic loop region upon Fe complex oxidation. The calculated pK(a) value for D1-Glu244 depends on the Fe complex redox state, yielding a pK(a) of 7.5 and 5.5 for Fe(2+) and Fe(3+), respectively. To account for the pH dependence of E(m)(Fe), a model involving not only D1-Glu244 but also the other titratable residues (five Glu in the D-de loops and six basic residues near the Fe complex) seems to be needed, implying the existence of a network of residues serving as an internal proton reservoir. Reduction of Q(A/B) yields +302 mV and +268 mV for E(m)(Fe) in the Q(A)(-)Q(B)(0) and Q(A)(0)Q(B)(-) states, respectively. Upon formation of the Q(A)(0)Q(B)(-) state, D1-His252 becomes protonated. Forming Fe(3+)Q(B)H(2) by a proton-coupled electron transfer process from the initial state Fe(2+)Q(B)(-) results in deprotonation of D1-His252. The two EPR signals observed at g = 1.82 and g = 1.9 in the Fe(2+)Q(A)(-) state of PSII may be attributed to D1-His252 with variable and fixed protonation, respectively.  相似文献   

16.
Enterobacter cloacae NAD(P)H:nitroreductase (NR; EC 1.6.99.7) catalyzes the reduction of a series of nitroaromatic compounds with steady-state bimolecular rate constants (kcat/Km) ranging from 10(4) to 10(7) M(-1) s(-1). In agreement with a previously proposed scheme of two-step four-electron reduction of nitroaromatics by NR (Koder, R. L., and Miller, A.-F. (1998) Biochim. Biophys. Acta 1387, 395-405), 2 mol NADH per mole mononitrocompound were oxidized. An oxidation of excess NADH by polinitrobenzenes, including explosives 2,4,6-trinitrotoluene (TNT) and 2,4,6-trinitrophenyl-N-methylnitramine (tetryl), has been observed as a slower secondary process, accompanied by O2 consumption. This type of "redox cycling" was not related to reactions of nitroaromatic anion-radicals, but was caused by the autoxidation of relatively stable reaction products. The initial reduction of tetryl and other polinitrophenyl-N-nitramines by E. cloacae NR was analogous to a two-step four-electron reduction mechanism of TNT and other nitroaromatics. The logs kcat/Km of all the compounds examined exhibited parabolic dependence on their enthalpies of single-electron or two-electron (hydride) reduction, obtained by quantum mechanical calculations. This type of quantitative structure-activity relationship shows that the reactivity of nitroaromatics towards E. cloacae nitroreductase depends mainly on their hydride accepting properties, but not on their particular structure, and does not exclude the possibility of multistep hydride transfer.  相似文献   

17.
The quinones duroquinone (DQ) and coenzyme Q(1) (CoQ(1)) and quinone reductase inhibitors have been used to identify reductases involved in quinone reduction on passage through the pulmonary circulation. In perfused rat lung, NAD(P)H:quinone oxidoreductase 1 (NQO1) was identified as the predominant DQ reductase and NQO1 and mitochondrial complex I as the CoQ(1) reductases. Since inhibitors have nonspecific effects, the goal was to use Nqo1-null (NQO1(-)/(-)) mice to evaluate DQ as an NQO1 probe in the lung. Lung homogenate cytosol NQO1 activities were 97 ± 11, 54 ± 6, and 5 ± 1 (SE) nmol dichlorophenolindophenol reduced·min(-1)·mg protein(-1) for NQO1(+/+), NQO1(+/-), and NQO1(-/-) lungs, respectively. Intact lung quinone reduction was evaluated by infusion of DQ (50 μM) or CoQ(1) (60 μM) into the pulmonary arterial inflow of the isolated perfused lung and measurement of pulmonary venous effluent hydroquinone (DQH(2) or CoQ(1)H(2)). DQH(2) efflux rates for NQO1(+/+), NQO1(+/-), and NQO1(-/-) lungs were 0.65 ± 0.08, 0.45 ± 0.04, and 0.13 ± 0.05 (SE) μmol·min(-1)·g dry lung(-1), respectively. DQ reduction in NQO1(+/+) lungs was inhibited by 90 ± 4% with dicumarol; there was no inhibition in NQO1(-/-) lungs. There was no significant difference in CoQ(1)H(2) efflux rates for NQO1(+/+) and NQO1(-/-) lungs. Differences in DQ reduction were not due to differences in lung dry weights, wet-to-dry weight ratios, perfusion pressures, perfused surface areas, or total DQ recoveries. The data provide genetic evidence implicating DQ as a specific NQO1 probe in the perfused rodent lung.  相似文献   

18.
A series of quinolinequinones bearing various substituents has been synthesized, and the effects of substituents on the metabolism of the quinones by recombinant human NAD(P)H:quinone oxidoreductase (hNQO1) was studied. A range of quinolinequinones were selected for study, and were specifically designed to probe the effects of aryl substituents at C-2. A range of 28 quinolinequinones 2-29 was prepared using three general strategies: the palladium(0) catalyzed coupling of 2-chloroquinolines, the classical Friedl?nder synthesis and the double-Vilsmeier reaction of acetanilides. One example of an isoquinolinequinone 30 was also prepared, and the reduction potentials of the quinones were measured by cyclic voltammetry. For simple substituents R(2) at the quinoline 2-position, the rates of quinone metabolism by hNQO1 decrease for R(2)=Cl>H approximately Me>Ph. For aromatic substituents, the rate of reduction decreases dramatically for R(2)=Ph>1-naphthyl>2-naphthyl>4-biphenyl. Compounds containing a pyridine substituent are the best substrates, and the rates decrease as R(2)=4-pyridyl>3-pyridyl>2-pyridyl>4-methyl-2-pyridyl>5-methyl-2-pyridyl. The toxicity toward human colon carcinoma cells with either no detectable activity (H596 or BE-WT) or high NQO1 activity (H460 or BE-NQ) was also studied in representative quinones. Quinones that are good substrates for hNQO1 are more toxic to the NQO1 containing or expressing cell lines (H460 and BE-NQ) than the NQO1 deficient cell lines (H596 and BE-WT).  相似文献   

19.

Background

Multimeric naphthoquinones are redox-active compounds that exhibit antineoplastic, antiprotozoal, and antiviral activities. Due to their multimodal effect on perturbation of cellular oxidative state, these compounds hold great potential as therapeutic agents against highly proliferative neoplastic cells. In our previous work, we developed a series of novel dimeric naphthoquinones and showed that they were selectively cytotoxic to human acute myeloid leukemia (AML), breast and prostate cancer cell lines. We subsequently identified the oxidoreductase NAD(P)H dehydrogenase, quinone 1 (NQO1) as the major target of dimeric naphthoquinones and proposed a mechanism of action that entailed induction of a futile redox cycling.

Results

Here, for the first time, we describe a direct physical interaction between the bromohydroxy dimeric naphthoquinone E6a and NQO1. Moreover, our studies reveal an extensive binding interface between E6a and the isoalloxazine ring of the flavin adenine dinucleotide (FAD) cofactor of NQO1 in addition to interactions with protein side chains in the active site. We also present biochemical evidence that dimeric naphthoquinones affect the redox state of the FAD cofactor of NQO1. Comparison of the mode of binding of E6a with those of other chemotherapeutics reveals unique characteristics of the interaction that can be leveraged in future drug optimization efforts.

Conclusion

The first structure of a dimeric naphthoquinone-NQO1 complex was reported, which can be used for design and synthesis of more potent next generation dimeric naphthoquinones to target NQO1 with higher affinity and specificity.
  相似文献   

20.
Li J  Takahashi E  Gunner MR 《Biochemistry》2000,39(25):7445-7454
The electron transfer from the reduced primary quinone (Q(A)(-)) to the secondary quinone (Q(B)) can occur in two phases with a well-characterized 100 micros component (tau(2)) and a faster process occurring in less than 10 micros (tau(1)). The fast reaction is clearly seen when the native ubiquinone-10 at Q(A) is replaced with naphthoquinones. The dependence of tau(1) on the free-energy difference between the P(+)Q(A)(-)Q(B) and P(+)Q(A)Q(B)(-) states (-) and on the pH was measured using naphthoquinones with different electrochemical midpoint potentials as Q(A) in Rhodobacter sphaeroides reaction centers (RCs) and in RCs where - is changed by mutation of M265 in the Q(A) site from Ile to Thr (M265IT). Q(B) was ubiquinone (UQ(B)) in all cases. Electron transfer was measured by using the absorption differences of the naphthosemiquinone at Q(A) and the ubisemiquinone at Q(B) between 390 and 500 nm. As - was changed from -90 to -250 meV tau(1) decreased from 29 to 0.2 micros. The free-energy dependence of tau(1) provides a reorganization energy of 850 +/- 100 meV for the electron transfer from Q(A)(-) to Q(B). The slower reaction at tau(2) is free-energy independent, so processes other than electron transfer determine the observed rate. The fraction of the reaction at tau(1) increases with increasing driving force and is 100% of the reaction when - is approximately 100 meV more favorable than in the native RCs with ubiquinone as Q(A). The fast phase, tau(1), is pH independent from pH 6 to 11 while tau(2) slows above pH 9. As the Q(A) isoprene tail length is increased from 2 to 10 isoprene units the fraction at tau(1) decreases. However, tau(1), tau(2), and the fraction of the reaction in each phase are independent of the tail length of UQ(B).  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