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.
Antioxidants can scavenge highly reactive radicals. As a result the antioxidants are converted into oxidation products that might cause damage to vital cellular components. To prevent this damage, the human body possesses an intricate network of antioxidants that pass over the reactivity from one antioxidant to another in a controlled way. The aim of the present study was to investigate how the semi-synthetic flavonoid 7-mono-O-(β-hydroxyethyl)-rutoside (monoHER), a potential protective agent against doxorubicin-induced cardiotoxicity, fits into this antioxidant network. This position was compared with that of the well-known flavonoid quercetin. The present study shows that the oxidation products of both monoHER and quercetin are reactive towards thiol groups of both GSH and proteins. However, in human blood plasma, oxidized quercetin easily reacts with protein thiols, whereas oxidized monoHER does not react with plasma protein thiols. Our results indicate that this can be explained by the presence of ascorbate in plasma; ascorbate is able to reduce oxidized monoHER to the parent compound monoHER before oxidized monoHER can react with thiols. This is a major difference with oxidized quercetin that preferentially reacts with thiols rather than ascorbate. The difference in selectivity between monoHER and quercetin originates from an intrinsic difference in the chemical nature of their oxidation products, which was corroborated by molecular quantum chemical calculations. These findings point towards an essential difference between structurally closely related flavonoids in their interplay with the endogenous antioxidant network. The advantage of monoHER is that it can safely channel the reactivity of radicals into the antioxidant network where the reactivity is completely neutralized.  相似文献   

2.
Flavonoids protect against oxidative stress by scavenging free radicals. During this protection flavonoids are oxidized. The oxidized flavonoids formed are often reactive. Consequently, protection by flavonoids can result in the formation of toxic products. In this study the oxidation of 7-mono-O-(β-hydroxyethyl)rutoside (monoHER), which is a constituent of the registered drug Venoruton, was studied in the absence and presence of glutathione (GSH). MonoHER was oxidized by horseradish peroxidase/H2O2. Spectrophotometric and HPLC analysis showed that in the presence of GSH, a monoHER–GSH conjugate was formed, which was identified as 2′-glutathionyl monohydroxyethylrutoside by mass spectrometric analysis and 1H NMR. Preferential formation of this glutathione adduct in the B ring at C2′ was confirmed by molecular quantum chemical calculations. This conjugate was also detected in the bile fluid of a healthy volunteer after iv administration of monoHER, demonstrating its formation in vivo. These results indicate that in the process of offering protection against free radicals, monoHER is converted into an oxidation product that is reactive toward thiols. The formation of this thiol-reactive oxidation product is potentially harmful. Thus, the supposed beneficial effect of monoHER as an antioxidant may be accompanied by the formation of products with an electrophilic, toxic potential.  相似文献   

3.
GSH was readily depleted by a flavonoid, H(2)O(2), and peroxidase mixture but the products formed were dependent on the redox potential of the flavonoid. Catalytic amounts of apigenin and naringenin but not kaempferol (flavonoids that contain a phenol B ring) when oxidized by H(2)O(2) and peroxidase co-oxidized GSH to GSSG via a thiyl radical which could be trapped by 5,5-dimethyl-1-pyrroline-N-oxide (DMPO) to form a DMPO-glutathionyl radical adduct detected by ESR spectroscopy. On the other hand, quercetin and luteolin (flavonoids that contain a catechol B ring) or kaempferol depleted GSH stoichiometrically without forming a thiyl radical or GSSG. Quercetin, luteolin, and kaempferol formed mono-GSH and bis-GSH conjugates, whereas apigenin and naringenin did not form GSH conjugates. MS/MS electrospray spectroscopy showed that mono-GSH conjugates for quercetin and luteolin had peaks at m/z 608 [M + H](+) and m/z 592 [M + H](+) in the positive-ion mode, respectively. (1)H NMR spectroscopy showed that the GSH was bound to the quercetin A ring. Spectral studies indicated that at a physiological pH the luteolin-SG conjugate was formed from a product with a UV maximum absorbance at 260 nm that was reducible by potassium borohydride. The quercetin-SG conjugate or kaempferol-SG conjugate on the other hand was formed from a product with a UV maximum absorbance at 335 nm that was not reducible by potassium borohydride. These results suggest that GSH was oxidized by apigenin/naringenin phenoxyl radicals, whereas GSH conjugate formation involved the o-quinone metabolite of luteolin or the quinoid (quinone methide) product of quercetin/kaempferol.  相似文献   

4.
The oxidation of quercetin by horseradish peroxidase/H(2)O(2) was studied in the absence but especially also in the presence of glutathione (GSH). HPLC analysis of the reaction products formed in the absence of GSH revealed formation of at least 20 different products, a result in line with other studies reporting the peroxidase-mediated oxidation of flavonoids. In the presence of GSH, however, these products were no longer observed and formation of two major new products was detected. (1)H NMR identified these two products as 6-glutathionylquercetin and 8-glutathionylquercetin, representing glutathione adducts originating from glutathione conjugation at the A ring instead of at the B ring of quercetin. Glutathione addition at positions 6 and 8 of the A ring can best be explained by taking into consideration a further oxidation of the quercetin semiquinone, initially formed by the HRP-mediated one-electron oxidation, to give the o-quinone, followed by the isomerization of the o-quinone to its p-quinone methide isomer. All together, the results of the present study provide evidence for a reaction chemistry of quercetin semiquinones with horseradish peroxidase/H(2)O(2) and GSH ultimately leading to adduct formation instead of to preferential GSH-mediated chemical reduction to regenerate the parent flavonoid.  相似文献   

5.
A novel cyclolignanic quinone, 7-acetyl-3',4'-didemethoxy-3',4'-dioxopodophyllotoxin (CLQ), inhibits topoisomerase II (TOPO II) activity. The extent of this inhibition was greater than that produced by the etoposide quinone (EQ) or etoposide. Glutathione (GSH) reduces EQ and CLQ to their corresponding semiquinones under anaerobic conditions. The latter were detected by EPR spectroscopy in the presence of MgCl2 but not in its absence. Semiquinone EPR spectra change with quinone/GSH mol ratio, suggesting covalent binding of GSH to the quinones. Quinone-GSH covalent adducts were isolated and identified by ESI-MS. These orthoquinones also react with nucleophilic groups from BSA to bind covalently under anaerobic conditions. BSA thiol consumption and covalent binding by these quinones are enhanced by MgCl2. Complex formation between the parent quinones and Mg+2 was also observed. Density functional calculations predict the observed blue-shifts in the absorption spectra peaks and large decreases in the partial negative charge of electrophilic carbons at the quinone ring when the quinones are complexed to Mg+2. These observations suggest a possible role of Mg+2 chelation by these quinones in increasing TOPO II thiol and/or amino/imino reactivity with these orthoquinones.  相似文献   

6.
Quercetin is one of the most prominent dietary antioxidants. During its antioxidant activity, quercetin becomes oxidized into its o-quinone/quinone methide QQ. QQ is toxic since it instantaneously reacts with thiols of, e.g., proteins. In cells, QQ will initially form an adduct with glutathione (GSH), giving GSQ. We have found that GSQ is not stable; it dissociates continuously into GSH and QQ with a half life of 2min. Surprisingly, GSQ incubated with 2-mercapto-ethanol (MSH), a far less reactive thiol, results in the conversion of GSQ into the MSH-adduct MSQ. A similar conversion of GSQ into relatively stable protein thiol-quercetin adducts is expected. With the dithiol dihydrolipoic acid (L(SH)(2)), quercetin is formed out of GSQ. These results indicate that GSQ acts as transport and storage of QQ. In that way, the initially highly focussed toxicity of QQ is dispersed by the formation of GSQ that finally spreads QQ-induced toxicity, probably even over cells.  相似文献   

7.
Covalent protein–phenol adducts, generated by reaction of protein nucleophiles with quinones, have recently attracted increased attention because the interactions change the functionality and physicochemical properties of proteins in biological and food systems. The formation of such covalent adducts between β-lactoglobulin (β-LG) and the quinone of 4-methylcatechol, 4-methylbenzoquinone (4MBQ), and subsequent reduction by dithiothreitol (DTT), tris(2-carboxyethyl)phosphine (TCEP), or sodium sulfite was investigated by mass spectrometry. The results showed that 19.0 ± 8.8% of β-LG reacted with 4MBQ when present in equimolar ratio at 20 °C (pH 8.0) to yield the protein–phenol adduct (β-LG-Q). Following treatment with sulfite, DTT, or TCEP, 75, 68, or 36%, respectively, of the formed β-LG-Q adduct dissociated. Different reaction mechanisms were proposed for the reduction of β-LG and β-LG-Q by each of the reducing agents. These results show that on reductive sample preparation for analysis of protein samples, not only are protein polymers formed through oxidative disulfide bonds reduced into the individual protein constituents but also a large part of any protein–phenol adducts present will dissociate and, thus, give a false picture of the level of protein–protein interactions that have occurred in the sample.  相似文献   

8.
When an antioxidant scavenges a reactive species, i.e., when it exerts its antioxidant activity, the antioxidant is converted into potentially harmful oxidation products. In this way, the antioxidant quercetin might yield an ortho-quinone, denoted as QQ, which has four tautomeric forms, i.e., the ortho-quinone and three quinonmethides. We evaluated the interaction of QQ with ascorbate or glutathione (GSH). Ascorbate recycles QQ to the parent compound quercetin, while GSH forms two adducts with QQ, i.e., 6-GSQ and 8-GSQ. When both GSH and ascorbate are present, QQ is converted exclusively into GSQ. In the absence of GSH, protein thiols will be arylated by QQ. This protein arylation is not prevented by ascorbate. Thiol arylation by quinones and quinonmethides can impair several vital enzymes. This implies that the product formed when quercetin displays its antioxidant scavenging effect is toxic in the absence of GSH. Therefore, an adequate GSH level should be maintained when quercetin is supplemented.  相似文献   

9.
The estrogen metabolites catecholestrogens (or hydroxyestrogens) are involved in carcinogenesis and the development of resistance to methotrexate. This induction of drug resistance correlates with the relative efficiency of catecholestrogens in the generation of reactive oxygen species (ROS) and the induction of DNA strand breaks. Although antioxidants can neutralize ROS, the generation of these reactive species by catecholestrogens can be enhanced by electron donors like NADH. Therefore, this study was undertaken to determine the ability of different thiol agents (GSH, NAC, DTT, DHLA) to either inhibit or enhance the level of DNA damage induced by the H(2)O(2) generating system 4-hydroxyestradiol/Cu(II). Our results show that GSH, DTT, and DHLA inhibited the induction of the 4-hydroxyestradiol/Cu(II)-mediated DNA damage, with GSH showing the best potential. In contrast, the GSH precursor NAC at low concentrations was able to enhance the level of oxidative damage, as observed with NADH. NAC can reduce Cu(II) to Cu(I) producing the radical NAC&z.rad;, which can generate the superoxide anion. However, the importance of this pathway appears to be relatively minor since the addition of NAC to the 4-hydroxyestradiol/Cu(II) system generates about 15 times more DNA strand breaks than NAC and Cu(II) alone. We suggest that NAC can perpetuate the redox cycle between the quinone and the semiquinone forms of the catecholestrogens, thereby enhancing the production of ROS. In conclusion, this study demonstrates the crucial importance of the choice of antioxidant as potential therapy against the negative biological effects of estrogens.  相似文献   

10.
Enhancement of acetaldehyde-protein adduct formation by L-ascorbate   总被引:5,自引:0,他引:5  
The effect of L-ascorbate on the binding of [14C]acetaldehyde to bovine serum albumin was examined. In the absence of ascorbate, acetaldehyde reacted with albumin to form both unstable (Schiff bases) and stable adducts. Ascorbate (5 mM) caused a time-dependent increase in the formation of total acetaldehyde-albumin adducts, which were comprised mainly of stable adducts. Significant enhancement of adduct formation by ascorbate was observed at acetaldehyde concentrations as low as 5 microM. An ascorbate concentration as low as 0.5 mM was still effective in stimulating stable adduct formation. The electron acceptor, 2,6 dichlorophenolindophenol, prevented the ascorbate-induced increase in albumin-adduct formation. Ascorbate also caused enhanced acetaldehyde adduct formation with other purified proteins, including cytochrome c and histones, as well as the polyamino acid, poly-L-lysine. These results indicate that ascorbate, acting as a reducing agent, can convert unstable acetaldehyde adducts to stable adducts, and can thereby increase and stabilize the binding of acetaldehyde to proteins.  相似文献   

11.
The covalent binding of [14C]acetaldehyde to purified beef brain tubulin was characterized. As we have found for several other proteins, tubulin bound acetaldehyde to form both stable and unstable adducts. Unstable adducts (Schiff bases) were stabilized, and rendered detectable, by treating incubated reaction mixtures with the reducing agent sodium borohydride. In short-term incubations, the majority of the adducts formed were unstable, but the percentage of total adducts that were stable gradually increased with time. Stable adduct formation was greatly increased by the inclusion of sodium cyanoborohydride in reaction mixtures (reductive ethylation). When reaction mixtures were submitted to sodium dodecyl sulfate-polyacrylamide gel electrophoresis to separate the alpha- and beta-chains of the heterodimeric tubulin molecule, the alpha-chain of free tubulin, but not intact microtubules, was the preferential site of stable adduct formation under both reductive and nonreductive conditions. Denaturation studies showed that the native tubulin conformation was necessary for the alpha-chain to show enhanced reactivity toward acetaldehyde. Competition binding studies showed that alpha-tubulin could effectively compete with beta-tubulin and bovine serum albumin for a limited amount of acetaldehyde. Unstable acetaldehyde adducts with free tubulin or microtubules did not exhibit alpha-chain selectivity. Analysis of reaction mixtures indicates that lysine residues are the major group of the protein participating in adduct formation. These data indicate that the alpha-chain of free tubulin is the preferential site of stable acetaldehyde-tubulin adduct formation. Further, these data raise the possibility that alpha-tubulin may be a selective target for acetaldehyde adduct formation in cellular systems.  相似文献   

12.
Polycyclic aromatic hydrocarbon (PAH) o-quinones are products of an NADP+ dependent oxidation of non-K-region trans-dihydrodiols catalyzed by dihydrodiol dehydrogenase (EC 1.3.1.20). Since these PAH o-quinones could be detoxified by non-enzymatic or enzymatic conjugation with cellular thiols, their reactivity with 2-mercaptoethanol, cysteine and glutathione (GSH) was examined by ion-pair reverse phase high pressure liquid chromatography (RP-HPLC). Second-order rate constants for the addition of these thiols to naphthalene-1,2-dione (NPQ) in water ranging from 4.9 x 10(3) - 1.1 x 10(4) min-1 M-1 and the reactions were complete within 10 min. When these reactions were conducted at near physiological pH (50 mM potassium phosphate buffer pH 7.0), the rate constants increased by 2-orders of magnitude. When benzo[a]pyrene-7,8-dione (BPQ) was substituted in these reactions the second-order rate constants decreased by 2-3 orders of magnitude and the reactions took several hours to reach completion. The decrease in reactivity can be explained by the presence of the bay region in BPQ. Methylation influenced the reactivity of PAH o-quinones with GSH and the following order of reactivity was observed: 7,12-dimethyl-benz[a]anthracene-3,4-dione (7,12-DMBAQ) > 12-methyl-BAQ, 7-methyl-BAQ and BAQ > BPQ. Of these quinones 7,12-dimethyl-BAQ was almost equi-reactive with NPQ. This suggests that methyl substitution in the bay and peri regions enhances reactivity with GSH. Using NPQ as a model for other PAH o-quinones, N-acetyl-L-cysteine, L-cysteine and GSH conjugates of NPQ were synthesized and characterized by [1H]- and [13C]NMR. Evidence for Michael type 1,4-addition products was obtained in which the resultant adduct could exist as either a catechol or o-quinone. By contrast, L-cysteine was able to form adducts via S- or N-attack and N-attack gave a purple p-iminoquinone. There was no evidence for the formation of bis-N-acetyl-L-cysteinyl-, bis-glutathionyl adducts or phenolic coupled products. The toxicity of thiol conjugates of NPQ remains to be explored.  相似文献   

13.
The molar absorptivities of the quinones produced from different o-diphenols, triphenols, and flavonoids were calculated by generating the respective quinones through oxidation with an excess of periodate. Oxidation of these substrates by this reagent was analogous to oxidation by tyrosinase with molecular oxygen, although the procedure showed several advantages over the enzymatic method in that oxidation took place almost immediately and quinone stability was favored because no substrate remained. The o-diphenols studied were pyrocatechol, 4-methylcatechol, 4-tert-butylcatechol, 3,4-dihydroxyphenylalanine, 3,4-dihydroxyphenylethylamine, 3,4-dihydroxyphenylacetic acid, 3,4-dihydroxyphenylpropionic acid, and caffeic acid; the triphenols studied were pyrogallol, 1,2,4-benzenetriol, 6-hydroxydopa, and 6-hydroxydopamine; and the flavonoids studied were (+)catechin, (-)epicatechin, and quercetin. In addition, the stability of the quinones generated by oxidation of the compounds by [periodate]0/[substrate]0 < 1 was studied. Taking the findings into account, tyrosinase could be measured by following o-quinone formation in rapid kinetic studies using the stopped-flow method. However, measuring o-quinone formation could not be useful for steady-state studies. Therefore, several methods for following tyrosinase activity are proposed, and a kinetic characterization of the enzyme's action on these substrates is made.  相似文献   

14.
The kinetic parameters of the redox transitions subsequent to the two-electron transfer implied in the glutathione (GSH) reductive addition to 2- and 6-hydroxymethyl-1,4-naphthoquinone bioalkylating agents were examined in terms of autoxidation, GSH consumption in the arylation reaction, oxidation of the thiol to glutathione disulfide (GSSG), and free radical formation detected by the spin-trapping electron spin resonance method. The position of the hydroxymethyl substituent in either the benzenoid or the quinonoid ring differentially influenced the initial rates of hydroquinone autoxidation as well as thiol oxidation. Thus, GSSG- and hydrogen peroxide formation during the GSH reductive addition to 6-hydroxymethyl-1,4-naphthoquinone proceeded at rates substantially higher than those observed with the 2-hydroxymethyl derivative. The distribution and concentration of molecular end products, however, was the same for both quinones, regardless of the position of the hydroxymethyl substituent. The [O2]consumed/[GSSG]formed ratio was above unity in both cases, thus indicating the occurrence of autoxidation reactions other than those involved during GSSG formation. EPR studies using the spin probe 5,5'-dimethyl-1-pyrroline-N-oxide (DMPO) suggested that the oxidation of GSH coupled to the above redox transitions involved the formation of radicals of differing structure, such as hydroxyl and thiyl radicals. These were identified as the corresponding DMPO adducts. The detection of either DMPO adduct depended on the concentration of GSH in the reaction mixture: the hydroxyl radical adduct of DMPO prevailed at low GSH concentrations, whereas the thiyl radical adduct of DMPO prevailed at high GSH concentrations. The production of the former adduct was sensitive to catalase, whereas that of the latter was sensitive to superoxide dismutase as well as to catalase. The relevance of free radical formation coupled to thiol oxidation is discussed in terms of the thermodynamic and kinetic properties of the reactions involved as well as in terms of potential implications in quinone cytotoxicity.  相似文献   

15.
A novel anti-5,5-dimethyl-1-pyrroline N-oxide (DMPO) polyclonal antiserum that specifically recognizes protein radical-derived DMPO nitrone adducts has been developed. In this study, we employed this new approach, which combines the specificity of spin trapping and the sensitivity of antigen-antibody interactions, to investigate protein radical formation from lactoperoxidase (LPO). When LPO reacted with GSH in the presence of DMPO, we detected an LPO radical-derived DMPO nitrone adduct using enzyme-linked immunosorbent assay and Western blotting. The formation of this nitrone adduct depended on the concentrations of GSH, LPO, and DMPO as well as pH values, and GSH could not be replaced by H(2)O(2). The level of this nitrone adduct was decreased significantly by azide, catalase, ascorbate, iodide, thiocyanate, phenol, or nitrite. However, its formation was unaffected by chemical modification of free cysteine, tyrosine, and tryptophan residues on LPO. ESR spectra showed that a glutathiyl radical was formed from the LPO/GSH/DMPO system, but no protein radical adduct could be detected by ESR. Its formation was decreased by azide, catalase, ascorbate, iodide, or thiocyanate, whereas phenol or nitrite increased it. GSH caused marked changes in the spectrum of compound II of LPO, indicating that GSH binds to the heme of compound II, whereas phenol or nitrite prevented these changes and reduced compound II back to the native enzyme. GSH also dose-dependently inhibited the peroxidase activity of LPO as determined by measuring 2,2'-azino-bis(3-ethylbenzothiazoline-6-sulfonic acid) oxidation. Taken together, these results demonstrate that the GSH-dependent LPO radical formation is mediated by the glutathiyl radical, possibly via the reaction of the glutathiyl radical with the heme of compound II to form a heme-centered radical trapped by DMPO.  相似文献   

16.
《Free radical research》2013,47(12):1462-1472
Abstract

In our in vitro study, we analyzed the effects of incubation of J774A.1 macrophages with reduced glutathione (GSH) and quercetin on the extent of cellular cholesterol efflux by high-density lipoprotein (HDL) or apolipoprotein A1 (apoA1). This combination was the most potent one among other exogenous and endogenous antioxidant combinations, since it significantly increased the extent of HDL-mediated cholesterol efflux from macrophages by 47% versus control cells, whereas quercetin (20 μM) or GSH (200 μM) alone increased it by only 37% or 17%, respectively. Similarly, apoA1-mediated cholesterol efflux was increased by 11% or 22% in quercetin or quercetin + GSH-treated cells, respectively, versus control cells. These stimulatory effects were noted only after 20 h of cell incubation. The combination of quercetin + GSH demonstrated high scavenging capacity of free radicals versus quercetin or GSH alone. In addition, quercetin + GSH significantly decreased macrophage oxidative stress as measured by the scavenging capacity of free radicals in the cells, the formation of reactive oxygen species, and the levels of cellular glutathione and lipid peroxides. There was no significant effect of quercetin + GSH on cellular HDL binding, on ATP-binding cassette A1 (ABCA1) activity, or on ABCG1 messenger RNA (mRNA) levels.

In contrast, mRNA levels for ABCA1 and peroxisome proliferator-activated receptor alpha (PPARα) were both significantly increased by 89% and 93%, respectively, in quercetin + GSH-treated cells versus control cells. Quercetin alone increased the mRNA levels for ABCA1 or PPARα by 42% or 77%, respectively, whereas GSH alone increased it by 22% or 28%, respectively. Mass spectra analysis revealed that oxidized quercetin reacts with GSH to form a new adduct product. We thus conclude that the stimulatory effects of quercetin + GSH on apoA1- or HDL-mediated macrophage cholesterol efflux are related to the ability of GSH to preserve quercetin in its reduced form.  相似文献   

17.
A novel cyclolignanic quinone, 7-acetyl-3′,4′-didemethoxy-3′,4′-dioxopodophyllotoxin (CLQ), inhibits topoisomerase II (TOPO II) activity. The extent of this inhibition was greater than that produced by the etoposide quinone (EQ) or etoposide. Glutathione (GSH) reduces EQ and CLQ to their corresponding semiquinones under anaerobic conditions. The latter were detected by EPR spectroscopy in the presence of MgCl2 but not in its absence. Semiquinone EPR spectra change with quinone/GSH mol ratio, suggesting covalent binding of GSH to the quinones. Quinone-GSH covalent adducts were isolated and identified by ESI-MS. These orthoquinones also react with nucleophilic groups from BSA to bind covalently under anaerobic conditions. BSA thiol consumption and covalent binding by these quinones are enhanced by MgCl2. Complex formation between the parent quinones and Mg+2 was also observed. Density functional calculations predict the observed blue-shifts in the absorption spectra peaks and large decreases in the partial negative charge of electrophilic carbons at the quinone ring when the quinones are complexed to Mg+2. These observations suggest a possible role of Mg+2 chelation by these quinones in increasing TOPO II thiol and/or amino/imino reactivity with these orthoquinones.  相似文献   

18.
Anthracycline antibiotics (e.g. doxorubicin and daunorubicin) are among the most effective and widely used anticancer drugs. Unfortunately, their clinical use is limited by the dose-dependent cardiotoxicity. Flavonoids represent a potentially attractive class of compounds to mitigate the anthracycline cardiotoxicity due to their iron-chelating, antioxidant and carbonyl reductase-inhibitory effects. The relative contribution of various characteristics of the flavonoids to their cardioprotective activity is, however, not known. A series of ten flavonoids including quercetin, quercitrin, 7-monohydroxyethylrutoside (monoHER) and seven original synthetic compounds were employed to examine the relationships between their inhibitory effects on carbonyl reduction, iron-chelation and antioxidant properties with respect to their protective potential against doxorubicin-induced cardiotoxicity. Cardioprotection was investigated in the neonatal rat ventricular cardiomyocytes whereas the H9c2 cardiomyoblast cells were used for cytotoxicity testing. Iron chelation was examined via the calcein assay and antioxidant effects and site-specific scavenging were quantified by means of inhibition of lipid peroxidation and hydroxyl radical scavenging activity, respectively. Inhibition of carbonyl reductases was assessed in cytosol from human liver. None of the flavonoids tested had better cardioprotective action than the reference cardioprotector, monoHER. However, a newly synthesized quaternary ammonium analog with comparable cardioprotective effects has been identified. No direct correlation between the iron-chelating and/or antioxidant effect and cardioprotective potential has been found. A major role of carbonyl reductase inhibition seems unlikely, as the best two cardioprotectors of the series are only weak reductase inhibitors.  相似文献   

19.
The effects of the dietary antioxidant N-acetylcysteine (NAC) on alcoholic liver damage were examined in a total enteral nutrition (TEN) model of ethanol toxicity in which liver pathology occurs in the absence of endotoxemia. Ethanol treatment resulted in steatosis, inflammatory infiltrates, occasional foci of necrosis, and elevated ALT in the absence of increased expression of the endotoxin receptor CD 14, a marker of Kupffer cell activation by LPS. In addition, ethanol treatment induced CYP 2 E1 and increased TNFalpha and TGFbeta mRNA expression accompanied by suppressed hepatic IL-4 mRNA expression. Ethanol treatment also resulted in the hepatic accumulation of malondialdehyde (MDA) and hydroxynonenal (HNE) protein adducts, decreased antioxidant capacity, and increased antibody titers toward serum hydroxyethyl radical (HER), MDA, and HNE adducts. NAC treatment increased cytosolic antioxidant capacity, abolished ethanol-induced lipid peroxidation, and inhibited the formation of antibodies toward HNE and HER adducts without interfering with CYP 2 E1 induction. NAC also decreased ethanol-induced ALT release and inflammation and prevented significant loss of hepatic GSH content. However, the improvement in necrosis score and reduction of TNFalpha mRNA elevation did not reach statistical significance. Although a direct correlation was observed among hepatic MDA and HNE adduct content and TNFalpha mRNA expression, inflammation, and necrosis scores, no correlation was observed between oxidative stress markers or TNFalpha and steatosis score. These data suggest that ethanol-induced oxidative stress can contribute to inflammation and liver injury even in the absence of Kupffer cell activation by endotoxemia.  相似文献   

20.
Abstract: Oxidation of l -3,4-dihydroxyphenylalanine ( l -DOPA) and dopamine (DA) to generate semiquinones/quinones, oxygen radicals, and other reactive oxygen species may play a role in neuronal cell death in Parkinson's disease (PD). In particular, semiquinones/quinones can form conjugates with thiol compounds such as GSH and cysteine. Exposure of l -DOPA, DA, and other catecholamines to a system generating O2•− radical led to O2•−-dependent depletion of added GSH (or cysteine), accompanied by the formation of thiol-DA or -DOPA adducts as detected by HPLC. Superoxide could additionally cause destruction of these adducts. Iron or copper ions could also promote conjugate formation between GSH or cysteine and DA and l -DOPA, especially if H2O2 was present. We applied HPLC to measure glutathionyl and cysteinyl conjugates of l -DOPA, DA, and 3,4-dihydroxyphenylacetic acid (DOPAC) in postmortem brain samples from PD patients and normal control subjects. Conjugates were detected in most brain areas examined, but levels were highest in the substantia nigra and putamen. In most regions, adduct levels were lower in PD, but there were significant increases in cysteinyl adducts of l -DOPA, DA, and DOPAC in PD substantia nigra, suggesting that acceleration of l -DOPA/DA oxidation occurs in PD, although we cannot say if this is a primary feature of the disease or if it is related to therapy with l -DOPA. In vitro, conjugate formation could be inhibited by the dithiol dihydrolipoate but not by its oxidised form, lipoic acid.  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