首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
Thirteen diversified antimetabolites of coenzyme Q10 which have antitumor activity in vivo were tested for inhibition of uptake of tritiated thymidine and uridine into DNA and RNA, respectively, of L1210 cells grown in tissue culture. Eight of these antimetabolites have inhibitory activities of the same order of magnitude as the used anticancer drugs, rubidazone and ellipticine. 5-ω-Phenylpropylmercapto-2,3-dimethoxy-1,4-benzoquinone was particularly potent to inhibit nucleic acid synthesis; ED50 for DNA = 2.1 μM and ED50 for RNA = 4.0 μM.  相似文献   

2.
The splitting of the carbonyl infrared bands of 2-methoxy-1,4-benzoquinone in solution can be related to a mesomeric resonance phenomenon leading to a conformation of the O-CH3 bond coplanar to the quinone ring. The delocalization of the electron density induces a frequency downshift of the C4=O carbonyl compared to 1,4-benzoquinone. This in turns decouples the two carbonyls leading to an upshift of the C1=O vibration. Using selective 13C-labeling of Q0 (2,3-dimethoxy-5-methyl-1,4-benzoquinone), we show that the effect of mesomeric resonance is an essential determinant of the carbonyl frequencies of ubiquinone in solution.  相似文献   

3.

Background

Coenzyme Q10 (CoQ10) and its analogs are used therapeutically by virtue of their functions as electron carriers, antioxidant compounds, or both. However, published studies suggest that different ubiquinone analogs may produce divergent effects on oxidative phosphorylation and oxidative stress.

Methodology/Principal Findings

To test these concepts, we have evaluated the effects of CoQ10, coenzyme Q2 (CoQ2), idebenone, and vitamin C on bioenergetics and oxidative stress in human skin fibroblasts with primary CoQ10 deficiency. A final concentration of 5 µM of each compound was chosen to approximate the plasma concentration of CoQ10 of patients treated with oral ubiquinone. CoQ10 supplementation for one week but not for 24 hours doubled ATP levels and ATP/ADP ratio in CoQ10 deficient fibroblasts therein normalizing the bioenergetics status of the cells. Other compounds did not affect cellular bioenergetics. In COQ2 mutant fibroblasts, increased superoxide anion production and oxidative stress-induced cell death were normalized by all supplements.

Conclusions/Significance

These results indicate that: 1) pharmacokinetics of CoQ10 in reaching the mitochondrial respiratory chain is delayed; 2) short-tail ubiquinone analogs cannot replace CoQ10 in the mitochondrial respiratory chain under conditions of CoQ10 deficiency; and 3) oxidative stress and cell death can be counteracted by administration of lipophilic or hydrophilic antioxidants. The results of our in vitro experiments suggest that primary CoQ10 deficiencies should be treated with CoQ10 supplementation but not with short-tail ubiquinone analogs, such as idebenone or CoQ2. Complementary administration of antioxidants with high bioavailability should be considered if oxidative stress is present.  相似文献   

4.
2,6-dibromothymoquinone (DBMIB) and other coenzyme Q analogs partially inhibit electron transport and the membrane-bound Mg++ stimulated ATPase of E. coli membranes. The inhibitions by DBMIB are fully reversed by coenzyme Q6, and other analogs show partial reversal by coenzyme Q6. Electron transport reactions inhibited are NADH and lactate oxidase, NADH menadione reductase, lactate phenazinemethosulfate reductase and duroquinol oxidase. The concentrations of DBMIB required are similar for electron transport and ATPase inhibition and inhibitions are all increased by uncouplers. Electron transport and ATPase are not inhibited in a DBMIB insensitive mutant. Soluble ATPase extracted from the membranes does not show DBMIB inhibition under either high or low Mg++ conditions. Lipophilic chelators show additional inhibition over DBMIB. It appears that coenzyme Q functions at three sites in E. coli electron transport where ATPase activity is controlled. Coenzyme Q deficient mutants also show decreased electron transport and ATPase activity which is restored by coenzyme Q.  相似文献   

5.
Superior antitumor activity of 1-β-D-arabinofuranosylcytosine (ara-C) conjugates of prednisolone and prednisone against L1210 leukemic mice, based on ara-C content, has encouraged us to synthesize 5′-(cortisone-21-phosphoryl)-1-β-D-arabinofuranosylcytosine (I) and 5′-(cortisone-21-phosphoryl)-1-β-d-arabinofuranosylcytosine (II) by condensation of N4,2′,3′-triacetyl-1-β-d-arabinofuranosylcytosine 5′-monophosphate with cortisol and cortisone in the presence of N,N′-dicyclohexylcarbodiimide at room temperature followed by removing the acetyl groups in 2 N methanolic ammonia in 20% yield. The conjugates I and II inhibited the invitro growth of L1210 by 50% (ED50) at 0.25 μM and 0.07 μM, respectively, while ara-C showed ED50 0.1 μM. However, the conjugates I and II exhibited 287% and 238% of TC at 50 mg/kg/day × 5 doses against L1210 leukemic mice, respectively, while ara-C at 25 mg and 50 mg/kg/day × 5 gave the respective 127% and 110% of TC.  相似文献   

6.
The effects of 33 quinone derivatives on mitochondrial electron transfer in yeast were examined. Twenty-two of the compounds were also tested for their effects on the growth of yeast cells. Four strong inhibitors of electron transfer were identified: 5-n-undecyl-6-hydroxy-4, 7-dioxobenzothiazole, 7-ω-cyclohexyloctyl-6-hydroxy-5,8-quinolinequinone, 7-n-hexadecyl-mercapto-6-hydroxy-5, 8-quinolinequinone, and 3-n-dodecylmercapto-2-hydroxy-1, 4-naphthoquinone. They inhibit the growth of yeast with ethanol as an energy source, but not when glucose is the energy source. The NADH oxidase activity of isolated mitochondria is 50% inhibited by these quinone derivatives at about 10?8m, or 0.5 μmol/g mitochondrial protein; 1000-fold higher concentrations do not affect electron transfer from NADH or succinate to coenzyme Q2. The effects of the inhibitors on cytochrome spectra indicate that they block electron transfer between cytochromes b and c1. A possible antagonism between these compounds and coenzyme Q at a site between cytochromes b and C1 is discussed in terms of Mitchell's “protonmotive Q cycle” hypothesis (Mitchell, P. (1976) J. Theor. Biol. 62, 327–367). 6-β-naphthylmercapto-5-chloro-2,3-dimethoxy-1,4-benzoquinone inhibits electron transfer between succinate and coenzyme Q2 or phenazine methosulfate, suggesting a site in the succinate-coenzyme Q reductase complex with a different quinone specificity from that of the site in the cytochrome bc1 complex. Seven of the quinone derivatives inhibit growth on both glucose and ethanol media, indicating that their effect is not the result of inhibition of respiration.  相似文献   

7.
The quinones 1,4-naphthoquinone (NQ), methyl-1,4-naphthoquinone (MNQ), trimethyl-1,4-benzoquinone (TMQ) and 2,3-dimethoxy-5-methyl-1,4-benzoquinone (UQ-0) enhance the rate of nitric oxide (NO) reduction by ascorbate in nitrogen-saturated phosphate buffer (pH 7.4). The observed rate constants for this reaction were determined to be 16±2,215±6,290±14 and 462±18?M-1?s-1, for MNQ, TMQ, NQ and UQ-0, respectively. These rate constants increase with an increase in quinone one-electron redox potential at neutral pH, E71. Since NO production is enhanced under hypoxia and under certain pathological conditions, the observations obtained in this work are very relevant to such conditions.  相似文献   

8.
2-Deoxy-2,3-dehydro-N-acetylneuraminic acid and its methyl ester are competitive inhibitors of Arthrobacter sialophilus neuraminidase with Ki = 1.4 × 10?6M and 4.8 × 10?5M, respectively. The Km for the substrate, N-acetylneuraminlactose, is 1.0 × 10?3M. These data, taken together with the conformation of these compounds, indicate that these compounds are transition-state analogs of the enzyme. These results also suggest that the substrate upon binding to neuraminidase is distorted to a conformation approaching that of a half-chair.  相似文献   

9.
Studies were made on the boundary conditions for thermotropic ovalbumin gelation at pH within the range 2.5 to 10.0. The pH dependence of the gelation threshold, C0, and denaturation temperature, Td, were obtained. The dependence C0(pH) has a sharp minimum close to the isoelectric point (pl). Over pH range 2.5 to 4.0 the dependence Td(pH) is linear; although above pI it shows unusual behaviour. Td increases smoothly, becoming a constant value (Td=80°C) at pH 7. Analysis of the temperature dependence of Leu's line integral intensity in the p.m.r. spectrum of ovalbumin shows that the temperature threshold of thermotropic gelation closely approximates to Td. A diagram for the state of an ovalbumin -water system was constructed in temperature-concentration-pH coordinates. The dependences of the initial shear modulus for thermotropic ovalbumin gels on the concentration (0.06≤C≤0.25g/cm3 were obtained at pH 4.0, 7.0, 8.5, 10.0. They are equivalent to the concentration dependence of the equilibrium elastic modulus Ee(C). The dependences obtained may be reduced to the theoretical master dependence of Hermans, Ee(rmC?), where C?=C/C0 is the reduced concentration. Hermans' theory, based o the model for random cross-linking of linear identical macromolecules without cyclization, adequately describes the equilibrium elastic properties of thermotropic ovalbumin gels.  相似文献   

10.
Prostaglandin congeners wherein the 15-hydroxy group is moved to the C16, C17, or C20 position or is replaced by a hydroxymethyl group were prepared via the 1,4-addition of a lithium trialkyl-trans-alkenyl alanate to an appropriate cyclopentenone. Several of the 16-hydroxy derivatives showed significant activity as constrictors of the isolated gerbil colon and in bronchodilator and anti-secretory assays.  相似文献   

11.
A number of previous studies of the involvement of 2-methyl-6-phytyl-1,4-benzoquinol in the biosynthesis of α-tocopherol have failed to take account of the fact that this quinol and its quinone have very similar chromatographic properties to those of 2-methyl-3-phytyl-1,4-benzoquinol and 2-methyl-3-phytyl-1,4-benzoquinone respectively. It has now been shown that the two quinones can be separated from each other either by multidevelopment TLC or by HPLC and that the claims made earlier with regard to the biosynthesis and metabolism of 2-methyl-6-phytyl-1,4-benzoquinol in chloroplasts are correct. In particular, it has been established that this quinol is the only methyl phytylbenzoquinol formed from homogentisate and phytyl pyrophosphate in chloroplast preparations. It has also been shown for the first time that lettuce chloroplasts are able to synthesize 3H-labelled α- and γ-tocopherols from [methylene-3H] homogentisate.  相似文献   

12.
Complexes of the formula cis-[Pt(HN+N)(L)Cl2], where (HN+N) are the protonated diamines including 3-aminoquinuclidine, N-aminopiperidine, piperazine, N-methylpiperazine, 1,1,4-trimethylpiperazine, and N-methyl-1,4-diazabicyclo [2,2,2] octane (N-methyl-dabco) and L = SCN?, NO2?, Br?, and F?, were synthesized from the protonated diamine complexes, [Pt(HN+N)Cl3]. The antitumor activities of the complexes were evaluated in vitro against L1210 murine leukemia cells, and ID50 values for the L-substituted complexes were compared to values of the parent complexes. In each case it was found that replacement of a chloride ion by SCN?, NO2?, Br?, or F?, either reduced or completely eliminated antitumor activity. This effect is explained in terms of the trans-directing ability of the ligand, L, compared to chloride. The NO2-substituted complex of 3- aminoquinuclidine was tested in vivo and found to exhibit little or no antitumor activity.  相似文献   

13.
A functionally active, spin labeled ubiquinone derivative, 2,3-dimethoxy -5-methyl-6-{10-(2,2,5,5-tetramethyl-3-pyrrolin-1-oxyl-3-carboxy)-decyl}-1,4-benzoquinone, has been synthesized for the study of ubiquinone binding in ubiquinol-cytochrome c reductase. When this spin labeled ubiquinone derivative interacted with ubiquinone- and phospholipid-depleted reductase, the spin label was totally immobilized. However, when phospholipids were replenished, the spin label showed mobility behaviour similar to that observed in a hydrophobic environment, indicating that the alkyl side chain of ubiquinone is extended into the hydrophobic region of intact reductase and has some degree of mobility.  相似文献   

14.
Oxidative cleavage of aromatic compounds is often part of a degradative process and is widely observed in nature. The immediate catabolic products can sometimes cyclize or rearrange to new secondary metabolites. The enzymatic contraction of a dehydroisocoumarin to yield cyclopentenoid metabolites in Cryptosporiopsis sp. is reported. The label distribution of (+) cryptosporiopsin, a chlorinated cyclopentenone, was determined by analysis of the [13C]nmr of [1-13C] and [2-13C]acetate enriched-cryptosporiopsin. The putative aromatic precursor of cyclopentenoid metabolites, 2,3-dihydro-6,8-dihydroxy-2-methylisocoumarin (6), was isolated from Aspergillus terreus. This metabolite (6) was prepared doubly labeled (T14C). The aromatic origin of the Cryptosporiopsis chlorinated cyclopentenoid metabolites was rigorously proven from feeding experiments with doubly labeled compound 6. A related but nonchlorinated metabolite, terrein, was isolated from A. terreus and was also shown to be derived from [T14C]-2,3-dihydro-6,8-dihydroxy-2-methylisocoumarin.  相似文献   

15.
Both pairs of dl-ll-desoxy- and dl-13-cis-erythro-15, 16-dihydroxyprostaglandins have been synthesized via 1,4-conjugate additions of an appropriately functionalized cis-vinyl cuprate to the requisite cyclopentenone. These prostaglandin analogs are considerably less potent than PGE2 as gastric secretion inhibitors or as bronchodilators.  相似文献   

16.
The effects of various quinone herbicides and fungicides on the photosynthetic 14CO2 fixation and the incorporation of 14C among the products of photosynthesis in Chlorella pyrenoidosa was investigated. Addition of 30 μm 2,3-dichloro-1,4-naphthoquinone (dichlone), 2-amino-3-chloro-1,4-naphthoquinone (06K-quinone), or 2,3,5,6-tetrachloro-1,4-benzoquinone (chloranil) inhibited CO2 fixation, whereas 1,4-benzoquinone had no effect. Treatment with 3 μm or higher concentrations of dichlone, 06K-quinone or 1,4-benzoquinone also produced marked changes in the pattern of 14C distribution. A noticeable effect was an increase in the proportion of 14C in sucrose and glycine accompanied by a reduction in 14C lipids and glutamic acid. These changes appear to occur as a result of shifts in the flow of carbon along various biosynthetic pathways of photosynthetic CO2 fixation. It is suggested that inactivation of coenzyme A and shortage of reduced triphosphopyridine nucleotide in the quinone-treated cells inhibited the synthesis of lipids and glutamic acid, thereby diverting more carbon into sucrose and glycine.  相似文献   

17.
To identify the structural features required for regulation of the mitochondrial permeability transition pore (PTP) by ubiquinone analogs (Fontaine, E., Ichas, F., and Bernardi, P. (1998) J. Biol. Chem. 40, 25734-25740), we have carried out an analysis with quinone structural variants. We show that three functional classes can be defined: (i) PTP inhibitors (ubiquinone 0, decylubiquinone, ubiquinone 10, 2,3-dimethyl-6-decyl-1,4-benzoquinone, and 2,3,5-trimethyl-6-geranyl-1,4-benzoquinone); (ii) PTP inducers (2,3-dimethoxy-5-methyl-6-(10-hydroxydecyl)-1,4-benzoquinone and 2,5-dihydroxy-6-undecyl-1,4-benzoquinone); and (iii) PTP-inactive quinones that counteract the effects of both inhibitors and inducers (ubiquinone 5 and 2,3,5-trimethyl-6-(3-hydroxyisoamyl)-1,4-benzoquinone) . The structure-function correlation indicates that minor modifications in the isoprenoid side chain can turn an inhibitor into an activator, and that the methoxy groups are not essential for the effects of quinones on the PTP. Since the ubiquinone analogs used in this study have a similar midpoint potential and decrease mitochondrial production of reactive oxygen species to the same extent, these results support the hypothesis that quinones modulate the PTP through a common binding site rather than through oxidation-reduction reactions. Occupancy of this site can modulate the PTP open-closed transitions, possibly through secondary changes of the PTP Ca(2+) binding affinity.  相似文献   

18.
The photoaffinity analogues of ubiquinone 2,3-dimethoxy-5-methyl-6-[2-[1-oxo-3-(4-azido-2-nitroanilino) propoxy]-3-methylbutyl]-1,4-benzoquinone (2'-ANAP-Q-1) and 2,3-dimethoxy-5-methyl-6-[3-[1-oxo-3-(4-azido-2-nitroanilino) propoxy]-3-methylbutyl]-1,4-benzoquinone (3'-ANAP-Q-1) have been synthesized. The required intermediate alcohols 2,3-dimethoxy-5-methyl-6-(2-hydroxy-3-methylbutyl)-1,4-benzoquinone and 2,3-dimethoxy-5-methyl-6-(3-hydroxy-3-methylbutyl)-1,4-benzoquinone were prepared in good yield from ubiquinone 1 by hydration of the side-chain double bond via hydroboration or acid catalysis, respectively. These alcohols were then coupled with 3-(4-azido-2-nitroanilino)propanoic acid, with p-toluenesulfonyl chloride in dry pyridine, to give 2'- and 3'-ANAP-Q-1. The synthetic methods presented should be of general utility in the preparation of derivatives of ubiquinone in which a reactive or reporter group is relatively close to the ubiquinone ring. By use of membrane vesicles prepared from a ubi-men-strain of Escherichia coli described previously [Wallace, B., & Young, I. G. (1977) Biochim. Biophys. Acta 461, 84-100], it has been shown that 2'- and 3'-ANAP-Q-1 substitute for ubiquinone 8 in the NADH, succinate, and D-lactate oxidase systems. Thus, these compounds may be of value in labeling respiratory chain proteins that interact with ubiquinone.  相似文献   

19.
A number of β-carboline analogs have been obtained or synthesized, and their in vitro receptor affinities and in vivo antagonist activities determined. The choice of analogs was made in order to explore the importance of the N9-H, the aromatic nitrogen and the C3-ester moiety for high-receptor affinity and antagonist activity of this class of benzodiazepine antagonist. Among the analogs investigated, we describe the properties of 3-cyano-β-carboline (lh), the first potent β-carboline antagonist without a carbonyl at the C3-position.The results obtained indicate: (1) Specific interactions of the C3-substituent with key cationic receptor sites rather than electron-withdrawing properties are important for high-receptor affinity and antagonist activity. (2) Specific in-plane interactions of the atomatic nitrogen with a cationic receptor site, rather than stacking with neutral aromatic residues of the receptor are also important for high affinity and antagonist activity. (3) While the presence of an N9H enhances receptor affinity, interaction with an anionic receptor site does not appear essential for antagonist activity.  相似文献   

20.
Several azido-ubiquinones have been synthesized for the study of protein-ubiquinone interaction in succinate-cytochrome c reductase. In the absence of light, azido-ubiquinones are partially effective in restoring enzymatic activity to ubiquinone- and phospholipid-depleted reductase and the binding of azido-ubiquinones can be partially reversed by 5-(10-bromodecyl)-ubiquinone. When 2-azido-3-methoxy-5-geranyl-6-methyl-1,4-benzoquinone reactivated reductase is illuminated with long wavelength UV light, a complete and irreversible inhibition is observed. This specific photo-inactivation, exerted only by 2-azido-3-methoxy-5-geranyl-6-methyl-1,4-benzoquinone, and not by other azido-ubiquinone derivatives, is evidence for the existence of a specific benzoquinone ring binding site in the enzyme.  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