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.
Because oxidative stress is a component of gastrointestinal injury, we investigated the effect of H(2)O(2) on transintestinal transport using isolated rat jejunum incubated in vitro. Millimolar concentrations of H(2)O(2) inhibited all the tested parameters without inducing any cytotoxic effect. Electrophysiological experiments indicated that H(2)O(2) decreases significantly both short circuit current and transepithelial electrical potential difference without affecting transepithelial resistance. The possibility that H(2)O(2) could influence (Na+, K+) -ATPase activity was explored using isolated basolateral membranes. Besides H(2)O(2), free radicals (O(2)(*-), HO*) were generated using different iron-dependent and independent systems; (Na+, K+) -ATPase activity was inhibited after membrane exposure to all ROS tested. The inhibition was prevented by allopurinol, superoxide dismutase or desferrioxamine. Western blot analysis showed a decreased expression of the alpha(1)-subunit of (Na+, K+) -ATPase. We conclude that H(2)O(2) may be a modulator of jejunal ion and water transport by multiple mechanisms, among which a significant inhibition of the basolateral (Na+, K+) -ATPase.  相似文献   

2.
T Offer  A Russo  A Samuni 《FASEB journal》2000,14(9):1215-1223
Native Cu,Zn-SOD and synthetic SOD mimics sometimes demonstrate an apparently anomalous bell-shaped dose-response relationship when protecting various biological systems from oxidative stress. Several mechanisms have been proposed to account for such an effect, including: overproduction of H(2)O(2), peroxidative activity of SOD, and opposing roles played by O(2)(*-) in both initiation and termination of radical chain reactions. In the present study, ferrocyanide and thiols, which are susceptible to one-electron and two-electron oxidation, respectively, were subjected to a flux of superoxide in the presence and absence of SOD or SOD mimics. The results show that 1) either O(2)(*-)/HO(2)(*) or H(2)O(2) alone partially inactivates papain, whereas when combined they act synergistically; 2) nitroxide SOD mimics, but not SOD, exhibit a bell-shaped dose-response relationship in protecting papain from inactivation; 3) SOD, which at low dose inhibits superoxide-induced oxidation of ferrocyanide, loses its antioxidative effect as its concentration increases. These findings offer an additional explanation for the pro-oxidative activity of SOD and SOD mimics without invoking any dual activity of O(2)(*-) or a combined effect of SOD and H(2)O(2). The most significant outcome of an increase in SOD level is a decrease of [O(2)(*-)](steady state), rather than any notable elevation of [H(2)O(2)](steady state). As a result, the reaction kinetics of the high oxidation state of each catalyst is altered. In the presence of ultra-low [O(2)(*-)](steady state), the oxidized form of SOD [Cu(II),Zn-SOD] or SOD mimic (oxo-ammonium cation) does not react with O(2)(*-) but rather oxidizes the target molecule that it was supposed to have protected. Consequently, these catalysts exert an anti- or pro-oxidative effect depending on their concentration.  相似文献   

3.
Oxygen activation during oxidation of the lignin-derived hydroquinones 2-methoxy-1,4-benzohydroquinone (MBQH(2)) and 2, 6-dimethoxy-1,4-benzohydroquinone (DBQH(2)) by laccase from Pleurotus eryngii was examined. Laccase oxidized DBQH(2) more efficiently than it oxidized MBQH(2); both the affinity and maximal velocity of oxidation were higher for DBQH(2) than for MBQH(2). Autoxidation of the semiquinones produced by laccase led to the activation of oxygen, producing superoxide anion radicals (Q(*-) + O(2) <--> Q + O(2)(*-)). As this reaction is reversible, its existence was first noted in studies of the effect of systems consuming and producing O(2)(*-) on quinone formation rates. Then, the production of H(2)O(2) in laccase reactions, as a consequence of O(2)(*-) dismutation, confirmed that semiquinones autoxidized. The highest H(2)O(2) levels were obtained with DBQH(2), indicating that DBQ(*-) autoxidized to a greater extent than did MBQ(*-). Besides undergoing autoxidation, semiquinones were found to be transformed into quinones via dismutation and laccase oxidation. Two ways of favoring semiquinone autoxidation over dismutation and laccase oxidation were increasing the rate of O(2)(*-) consumption with superoxide dismutase (SOD) and recycling of quinones with diaphorase (a reductase catalyzing the divalent reduction of quinones). These two strategies made the laccase reaction conditions more natural, since O(2)(*-), besides undergoing dismutation, reacts with Mn(2+), Fe(3+), and aromatic radicals. In addition, quinones are continuously reduced by the mycelium of white-rot fungi. The presence of SOD in laccase reactions increased the extent of autoxidation of 100 microM concentrations of MBQ(*-) and DBQ(*-) from 4.5 to 30.6% and from 19.6 to 40.0%, respectively. With diaphorase, the extent of MBQ(*-) autoxidation rose to 13.8% and that of DBQ(*-) increased to 39.9%.  相似文献   

4.
It has been reported that a large amount of reactive oxygen species (ROS) is produced during seed imbibition and this ROS is related to seed vigor. To make this physiological mechanism clear, we have used 2-methyl-6-(4-methoxyphenyl)-3,7-dihydroimidazo(1,2-alpha)pyrazin-3-one (MCLA) as a sensitive and physiologically compatible probe for the determination of superoxide anion (O(2)(*-)) production in vivo. Our results showed that dry rice (Oryzae sativa L.) seed embryo cells possessed the capacity to generate O(2)(*-). Conversely, the O(2)(*-) production of seed embryo cells was inhibited by quinacrine (QA) and diphenylene iodonium (DPI), two specific inhibitors of NADPH oxidase, and O(2)(*-) induced MCLA-mediated chemiluminescence was also blocked by superoxide dismutase (SOD). Additionally, O(2)(*-) -production ability increased dramatically in a NADPH-dependent way in the plasma membrane protein abstract from rice seed embryo cells, whereas SOD and the inhibitors mentioned above suppressed O(2)(*-) production. These preliminary results suggested that rice seeds contained intrinsic NADPH oxidase activity. To validate this conclusion, dichlorofluorescein (DCF) fluorescence staining was used (observed under a laser scanning microscope, LSM) to reflect the in situ assessment of O(2)(*-) -generation. The position of O(2)(*-) production located at the plasma membrane. Additionally the ability to synthesize O(2)(*-) was activated directly by calcium ions. These observations are in accord with the character of NADPH oxidase catalyzed O(2)(*-) -generation. All these results indicated that NADPH oxidase contribute to O(2)(*-) production and release to the outside. We concluded that NADPH oxidase plays an intrinsic role as an NADPH sensor, so, measuring the O(2)(*-) one can monitor the NADPH concentration, which is an index of seed vigor. Therefore the O(2)(*-) generation during early imbibition can serve as a rapid measurement of seed vigor.  相似文献   

5.
The interaction of Cu,ZnSOD with H2O2 generates an oxidant at the active site that can then cause either the inactivation of this enzyme or the oxidation of a variety of exogenous substrates. We show that the rate of inactivation, imposed by 10-mM H2O2 at 25 degrees C and pH 7.2, is not influenced by 10-mM HCO3-; whereas the oxidation of 2,2'-azino-bis-[3-ethylbenzothiazoline sulfonate] (ABTS=) is virtually completely dependent upon HCO3-. The reduction of the active site Cu(II) by H2O2, which precedes inactivation of the enzyme, occurred at the same rate in phosphate buffer with or without bicarbonate added. These results indicate that HCO3- does not play a role in facilitating the interaction of H2O2 with the active site copper, but they can be accommodated by the proposal that HCO3- is oxidized to HCO3*, which then diffuses from that site and causes the oxidation of substrates, such as ABTS=, that are too large to traverse the solvent access channel to the Cu(II).  相似文献   

6.
Arsenic trioxide (ATO) can regulate many biological functions such as apoptosis and differentiation in various cells. We investigated an involvement of ROS such as H(2)O(2) and O(2)(*-), and GSH in ATO-treated Calu-6 cell death. The levels of intracellular H(2)O(2) were decreased in ATO-treated Calu-6 cells at 72 h. However, the levels of O(2)(*-) were significantly increased. ATO reduced the intracellular GSH content. Many of the cells having depleted GSH contents were dead, as evidenced by the propidium iodine staining. The activity of CuZn-SOD was strongly down-regulated by ATO at 72 h while the activity of Mn-SOD was weakly up-regulated. The activity of catalase was decreased by ATO. ROS scavengers, Tiron and Trimetazidine did not reduce levels of apoptosis and intracellular O(2)(*-) in ATO-treated Calu-6 cells. Tempol showing a decrease in intracellular O(2)(*-) levels reduced the loss of mitochondrial transmembrane potential (DeltaPsi(m)). Treatment with NAC showing the recovery of GSH depletion and the decreased effect on O(2)(*-) levels in ATO-treated cells significantly inhibited apoptosis. In addition, BSO significantly increased the depletion of GSH content and apoptosis in ATO-treated cells. Treatment with SOD and catalase significantly reduced the levels of O(2)(*-) levels in ATO-treated cells, but did not inhibit apoptosis along with non-effect on the recovery of GSH depletion. Taken together, our results suggest that ATO induces apoptosis in Calu-6 cells via the depletion of the intracellular GSH contents rather than the changes of ROS levels.  相似文献   

7.
Here we show that both Antimycin A, a respiratory chain inhibitor inducing apoptosis, and endotoxic shock, a syndrome accompanied by both necrosis and apoptosis, cause not only an increase but also the leakage of superoxide radicals (O(2)(*-)) from rat heart mitochondria (RHM), while O(2)(*-) generated in intact RHM do not escape from mitochondria. This was shown by a set of O(2)(*-)-sensitive spin probes with varying hydrophobicity. The levels of O(2)(*-) detected in intact RHM gradually increase as the hydrophobicity of spin probes increases and were not sensitive to superoxide dismutase (SOD) added to the incubation medium. Both Antimycin A and endotoxic shock elevated O(2)(*-) levels. Elevated O(2)(*-) levels became sensitive to SOD but in a different manner. The determination of O(2)(*-) with water-soluble PPH was fully sensitive to SOD, while the determination of O(2)(*-) with the more hydrophobic CMH and CPH was only partially sensitive to SOD, suggesting the release of a portion of O(2)(*-) into the surrounding medium.  相似文献   

8.
Summary We have considered the regulatory interrelationship of the plasma membrane oxidoreductase (PMOR) system and the mitochondrial respiratory capacity of human Namalwa (lymphoblastoid) cells. To this end, we made use of mitchondrially respiratory competent (+) cells and 0 cells, which lack mitochondrial DNA (mtDNA) and consequently mitochondrial respiratory activity. NADH-fer-ricyanide reductase activity of the PMOR system is increased 3-fold in 0 Namalwa cells compared to + cells. It is also shown for the first time that addition of coenzyme Q10 and coenzyme Q10-ana-logues, which can rescue 0 Namalwa cells in the absence of pyravate, gives rise to a further 2–3-fold increase in plasma membrane NADH-ferricyanide reductase activity. These systems were examined to determine if there exists a correlation between the regulation of the PMOR system and extracellular Superoxide radical formation as measured with the fluorescence probe L-012. No correlation was found between NADH-ferricyanide reductase activity and extracellular Superoxide radical production. PMOR function in cellular proliferation appears therefore not to involve extracellular Superoxide radical production.Abbreviations CoQ10 coenzyme Q10 - EtBr ethidium bromide - HCO-60 polyoxyethylated hydrogenated castor oil - HEPES 4-(2-hydroxyethyl)-1-piperazineethanesulfonic acid - mtDNA mitochondrial DNA - L-012 8-amino-5-chloro-7-phenylpyrido(3,4-d)pyridazine-1,4(2H,3H)dione - SOD Superoxide dismutase  相似文献   

9.
Early determinants of H2O2-induced endothelial dysfunction   总被引:4,自引:0,他引:4  
Reactive oxygen species (ROS) can stimulate nitric oxide (NO(*)) production from the endothelium by transient activation of endothelial nitric oxide synthase (eNOS). With continued or repeated exposure, NO(*) production is reduced, however. We investigated the early determinants of this decrease in NO(*) production. Following an initial H(2)O(2) exposure, endothelial cells responded by increasing NO(*) production measured electrochemically. NO(*) concentrations peaked by 10 min with a slow reduction over 30 min. The decrease in NO(*) at 30 min was associated with a 2.7-fold increase in O(2)(*-) production (p < 0.05) and a 14-fold reduction of the eNOS cofactor, tetrahydrobiopterin (BH(4), p < 0.05). Used as a probe for endothelial dysfunction, the integrated NO(*) production over 30 min upon repeated H(2)O(2) exposure was attenuated by 2.1-fold (p = 0.03). Endothelial dysfunction could be prevented by BH(4) cofactor supplementation, by scavenging O(2)(*-) or peroxynitrite (ONOO(-)), or by inhibiting the NADPH oxidase. Hydroxyl radical (()OH) scavenging did not have an effect. In summary, early H(2)O(2)-induced endothelial dysfunction was associated with a decreased BH(4) level and increased O(2)(*-) production. Dysfunction required O(2)(*-), ONOO(-), or a functional NADPH oxidase. Repeated activation of the NADPH oxidase by ROS may act as a feed forward system to promote endothelial dysfunction.  相似文献   

10.
L-Amino acid oxidase is a dimeric glycosylated flavoenzyme, a major constituent of the venom-from the snake Calloselasma rhodostoma. The enzyme exhibits apoptosis inducing effects as well as antibacterial and anti-HIV activities. The structure of l-amino acid oxidase with its substrate (L-phenylalanine) has been refined to a resolution of 1.8 A. The complex structure reveals the substrate bound to the reduced flavin (FADred). Alternative conformations for the key residues His223 and Arg322 are evident, suggesting a dynamic active site. Furthermore, conformational changes are apparent for the isoalloxazine ring; the three-ring system exhibits more bending around the N5-N10 axis compared to the oxidized flavin. The implications of the observed dynamics on the mechanism of catalysis are discussed. Inspection of buried surfaces in the enzyme reveals a Y-shaped channel system extending from the external surface of the protein to the active site. One portion of this channel may serve as the entry path for O2 during the oxidative half-reaction. The second region, separated from the proposed O2 channel by the N terminus (residues 8-16) of the protein, may play a role in H2O2 release. Interestingly, the latter portion of the channel would direct the H2O2 product to the exterior surface of the protein, near the glycan moiety, thought to anchor the enzyme to the host cell. This channel location may explain the ability of the enzyme to localize H2O2 to the targeted cell and thus induce the apoptotic effect.  相似文献   

11.
Previously it has been demonstrated that the human epidermis synthesises and degrades acetylcholine and expresses both muscarinic and nicotinic receptors. These cholinergic systems have been implicated in the development of the epidermal calcium gradient and differentiation in normal healthy skin. In vitiligo severe oxidative stress occurs in the epidermis of these patients with accumulation of H2O2 in the 10(-3)M range together with a decrease in catalase expression/activity due to deactivation of the enzyme active site. It was also shown that the entire recycling of the essential cofactor (6R)-l-erythro-5,6,7,8-tetrahydrobiopterin via pterin-4a-carbinolamine dehydratase (PCD) and dihydropteridine reductase (DHPR) is affected by H2O2 oxidation of Trp/Met residues in the enzyme structure leading to deactivation of these proteins. Using fluorescence immunohistochemistry we now show that epidermal H2O2 in vitiligo patients yields also almost absent epidermal acetylcholinesterase (AchE). A kinetic analysis using pure recombinant human AchE revealed that low concentrations of H2O2 (10(-6)M) activate this enzyme by increasing the Vmax>2-fold, meanwhile high concentrations of H2O2 (10(-3)M) inhibit the enzyme with a significant decrease in Vmax. This result was confirmed by fluorescence excitation spectroscopy following the Trp fluorescence at lambdamax 280nm. Molecular modelling based on the established 3D structure of human AchE supported that H2O2-mediated oxidation of Trp(432), Trp(435), and Met(436) moves and disorients the active site His(440) of the enzyme, leading to deactivation of the protein. To our knowledge these results identified for the first time H2O2 regulation of AchE. Moreover, it was shown that H2O2-mediated oxidation of AchE contributes significantly to the well-established oxidative stress in vitiligo.  相似文献   

12.
Zhang F  Vasella A 《Carbohydrate research》2007,342(17):2546-2556
Partially deuteriated 1,5,6,6-(2)H(4)-d-glucose and 1(I),1(II),5(I),5(II),6(I),6(I),6(II),6(II)-(2)H(8)-d-cellobiose were synthesized in high yields and on a large scale from d-glucose. (2)H enrichment at C-5 and C-6 of each glucopyranosyl unit in excess of 85% and 90%, respectively, was realized by (1)H-(2)H exchange in (2)H(2)O containing deuteriated Raney Ni. Nucleophilic addition of LiAlD(4) to 5,6,6-(2)H(3)-2,3,4,6-tetra-O-benzyl-d-gluconolactone led to a 98% (2)H enrichment at C-1. Deuteriated cellobiose is of interest as building block for the synthesis of a model compound of cellulose I.  相似文献   

13.
A number of studies indicate that reactive oxygen species (ROS) are involved in neurodegenerative diseases, such as Alzheimer's disease (AD) and Parkinson's disease (PD). The neuroprotective effects of salvianolic acid B (SalB) from Radix Salviae miltiorrhizae (RSM) against hydrogen peroxide (H2O2)-induced rat pheochromocytoma line PC12 injury were evaluated in the present study. Vitamin E, a potent antioxidant, was employed as a positive control agent. Following exposure of cells to H2O2 (150 microM), a marked decrease in cell survival and activities of superoxide dismutase (SOD), catalase (CAT) and glutathione peroxidase (GSH-Px), as well as increased levels of malondialdehyde (MDA) production and lactate dehydrogenase (LDH) release were observed. In parallel, H2O2 caused significant elevation in intracellular Ca2+ level and caspase-3 activity, and induced apoptotic death as determined by flow cytometric assay. However, pretreatment of the cells with SalB (0.1-10 microM) prior to H2O2 exposure blocked these H2O2-induced cellular events noticeably. Moreover, SalB exhibited significantly higher potency as compared to Vitamin E. The present findings indicated that SalB exerts neuroprotective effects against H2O2 toxicity, which might be of importance and contribute to its clinical efficacy for the treatment of neurodegenerative diseases.  相似文献   

14.
The heme protein indoleamine 2,3-dioxygenase (IDO) initiates oxidative metabolism of tryptophan along the kynurenine pathway, and this requires reductive activation of Fe(3+)-IDO. The current dogma is that superoxide anion radical (O(2)(*-)) is responsible for this activation, based largely on previous work employing purified rabbit IDO and rabbit enterocytes. We have re-investigated this role of O(2)(*-) using purified recombinant human IDO (rhIDO), rabbit enterocytes that constitutively express IDO, human endothelial cells, and monocyte-derived macrophages treated with interferon-gamma to induce IDO expression, and two cell lines transfected with the human IDO gene. Both potassium superoxide and O(2)(*-) generated by xanthine oxidase modestly activated rhIDO, in reactions that were prevented completely by superoxide dismutase (SOD). In contrast, SOD mimetics had no effect on IDO activity in enterocytes and interferon-gamma-treated human cells, despite significantly decreasing cellular O(2)(*-) Similarly, cellular IDO activity was unaffected by increasing SOD activity via co-expression of Cu,Zn-SOD or by increasing cellular O(2)(*-) via treatment of cells with menadione. Other reductants, such as tetrahydrobiopterin, ascorbate, and cytochrome P450 reductase, were ineffective in activating cellular IDO. However, recombinant human cytochrome b(5) plus cytochrome P450 reductase and NADPH reduced Fe(3+)-IDO to Fe(2+)-IDO and activated rhIDO in a reconstituted system, a reaction inhibited marginally by SOD. Additionally, short interfering RNA-mediated knockdown of microsomal cytochrome b(5) significantly decreased IDO activity in IDO-transfected cells. Together, our data show that cytochrome b(5) rather than O(2)(*-) plays a major role in the activation of IDO in human cells.  相似文献   

15.
The present study tested the hypothesis that membrane-bound NAD(P)H oxidase in coronary arterial myocytes (CAMs) is capable of producing superoxide (O(2)(*-)) toward extracellular space to exert an autocrine- or paracrine-like action in these cells. Using a high-speed wavelength-switching fluorescent microscopic imaging technique, we simultaneously monitored the binding of dihydroethidium-oxidizing product to exogenous salmon testes DNA trapped outside CAMs and to nuclear DNA as indicators of extra- and intracellular O(2)(*-) production. It was found that a muscarinic agonist oxotremorine (OXO; 80 microM) increased O(2)(*-) levels more rapidly outside than inside CAMs. In the presence of superoxide dismutase (500 U/ml) plus catalase (400 U/ml) and NAD(P)H oxidase inhibitor diphenylene iodonium (50 microM) or apocynin (100 microM), these increases in extra- and intracellular O(2)(*-) levels were substantially abolished or attenuated. The O(2)(*-) increase outside CAMs was also confirmed by detecting oxidation of nitro blue tetrazolium and confocal microscopic localization of Matrigel-trapped OxyBURST H(2)HFF Green BSA staining around these cells. By electron spin resonance spectrometry, the extracellular accumulation of O(2)(*-) was demonstrated as a superoxide dismutase-sensitive component outside CAMs. Furthermore, RNA interference of NAD(P)H oxidase subunits Nox1 or p47 markedly blocked OXO-induced increases in both extra- and intracellular O(2)(*-) levels, whereas small inhibitory RNA of Nox4 only attenuated intracellular O(2)(*-) accumulation. These results suggest that Nox1 represents a major NAD(P)H oxidase isoform responsible for extracellular O(2)(*-) production. This rapid extracellular production of O(2)(*-) seems to be unique to OXO-induced M(1)-receptor activation, since ANG II-induced intra- and extracellular O(2)(*-) increases in parallel. It is concluded that the outward production of O(2)(*-) via NAD(P)H oxidase in CAMs may represent an important producing pattern for its autocrine or paracrine actions.  相似文献   

16.
Substantial differences in the in vivo effect of paraquat (Pq) and DTT on basic peroxidase (GBPx) activity and on H2O2 levels were found in grapevine leaves cv. Sultana. GBPx activity decreased and H2O2 levels increased in illuminated Pq treated leaf-discs. Inactivation of GBPx and accumulation of H2O2 depended on the duration and intensity of the illumination to which discs were exposed. Since GBPx was inactivated directly by H2O2 and not by Pq in leaf extracts, and since GBPx are cytosolic isoenzymes and H2O2 is a stable molecule that can easily permeate chloroplast membranes, we concluded that Pq inactivation of GBPx in vivo is mediated by H2O2. In contrast to the effect induced by Pq, DTT directly inactivated GBPx in leaf extracts. In leaf-discs, however, it reduced GBPx activity in the absence of light, although the levels of H2O2 increased only after exposure of the discs to high irradiance, suggesting that under excess of light, a significant fraction of the photosynthetically produced electrons are dissipated through the water-water cycle and H2O2 accumulates as a consequence of GBPx inactivation.  相似文献   

17.
To investigate the role of catalase and superoxide dismutase (SOD) in the acetic acid (AA) induced yeast programmed cell death (AA-PCD), we compared Saccharomyces cerevisiae cells (C-Y) and cells individually over-expressing catalase T (CTT1-Y) and Cu,Zn-SOD (SOD1-Y) with respect to cell survival, hydrogen peroxide (H2O2) levels and enzyme activity as measured up to 200 min after AA treatment. AA-PCD does not occur in CTT1-Y, where H2O2 levels were lower than in C-Y and the over-expressed catalase activity decreased with time. In SOD1-Y, AA-PCD was exacerbated; high H2O2 levels were found, SOD activity increased early, remaining constant en route to AA-PCD, but catalase activity was strongly reduced.  相似文献   

18.
The principal source of hydrogen peroxide in mitochondria is thought to be from the dismutation of superoxide via the enzyme manganese superoxide dismutase (MnSOD). However, the nature of the effect of SOD on the cellular production of H(2)O(2) is not widely appreciated. The current paradigm is that the presence of SOD results in a lower level of H(2)O(2) because it would prevent the non-enzymatic reactions of superoxide that form H(2)O(2). The goal of this work was to: a) demonstrate that SOD can increase the flux of H(2)O(2), and b) use kinetic modelling to determine what kinetic and thermodynamic conditions result in SOD increasing the flux of H(2)O(2). We examined two biological sources of superoxide production (xanthine oxidase and coenzyme Q semiquinone, CoQ(*-) that have different thermodynamic and kinetic properties. We found that SOD could change the rate of formation of H(2)O(2) in cases where equilibrium-specific reactions form superoxide with an equilibrium constant (K) less than 1. An example is the formation of superoxide in the electron transport chain (ETC) of the mitochondria by the reaction of ubisemiquinone radical with dioxygen. We measured the rate of release of H(2)O(2) into culture medium from cells with differing levels of MnSOD. We found that the higher the level of SOD, the greater the rate of accumulation of H(2)O(2). Results with kinetic modelling were consistent with this observation; the steady-state level of H(2)O(2) increases if K<1, for example CoQ(*-)+O(2)-->CoQ+O(2)(*-). However, when K>1, e.g. xanthine oxidase forming O(2)(*-), SOD does not affect the steady state-level of H(2)O(2). Thus, the current paradigm that SOD will lower the flux of H(2)O(2) does not hold for the ETC. These observations indicate that MnSOD contributes to the flux of H(2)O(2) in cells and thereby is involved in establishing the cellular redox environment and thus the biological state of the cell.  相似文献   

19.
We have isolated a cDNA clone from rat brain using a human platelet 2-adrenergic receptor genomic clone as a probe. Comparison of the deduced amino acid sequence (450 residues) corresponding to the rat brain cDNA with that of the human platelet and human kidney 2-adrenergic receptors showed 84% and 44% sequence similarity, respectively. The major sequence difference between the rat brain and human platelet proteins, was a stretch of 48 amino acids within the third cytosolic loop in which the similarity was only 42%. Analysis of the 48 amino acid-region indicated that the two receptors significantly differ in terms of their primary amino acid sequence and the predicted secondary and tertiary structural features. There was no sequence similarity between the human platelet and rat brain clone over the 177 bases of 3-noncoding sequence and a less than 50% similarity over a stretch of 210 nucleotides in the 5-untranslated region. Southern-blot analysis with a human platelet 2-adrenergic receptor probe revealed the existence of a single 5.2 kb restriction fragment (KpnI/SacI) in both human and rat genomic DNA; the rat brain 2-receptor probe, however, hybridized to a single 1.9 kb band in rat DNA. Northern-blot analysis of rat brain poly(A+) RNA with the rat brain cDNA probe under stringent hybridization conditions revealed a single 4.5 kb mRNA; none was detected by the human platelet receptor probe. The rat brain 4.5 kb mRNA was not detected in any (other than brain) tested rat tissues utilizing either rat brain or human platelet DNA probes. The rat brain cDNA was expressed in a mammalian cell line (COS-2A) and found to bind the 2-adrenergic antagonist [3H]yohimbine; based on the binding-affinity for prazosin, the presently cloned receptor was pharmacologically closer to the 2A subclass. We conclude that the rat brain cDNA encodes a new 2-adrenergic receptor subtype that may be brain-specific.Abbreviations G protein guanine nucleotide-binding proteins - cA2-47 2-adrenergic receptor cDNA from rat brain - SSC (1X SSC contains 0.15 M NaCl, 15 mM Na3citrate, pH 7.0)  相似文献   

20.
To define the mechanism of arsenite-induced tumor promotion, we examined the role of reactive oxygen species (ROS) in the signaling pathways of cells exposed to arsenite. Arsenite treatment resulted in the persistent activation of p70(s6k) and extracellular signal-regulated kinase 1/2 (ERK1/2) which was accompanied by an increase in intracellular ROS production. The predominant produced appeared to be H(2)O(2), because the arsenite-induced increase in dichlorofluorescein (DCF) fluorescence was completely abolished by pretreatment with catalase but not with heat-inactivated catalase. Elimination of H(2)O(2) by catalase or N-acetyl-L-cysteine inhibited the arsenite-induced activation of p70(s6k) and ERK1/2, indicating the possible role of H(2)O(2) in the arsenite activation of the p70(s6k) and the ERK1/2 signaling pathways. A specific inhibitor of p70(s6k), rapamycin, and calcium chelators significantly blocked the activation of p70(s6k) induced by arsenite. While the phosphatidylinositol 3-kinase (PI3K) inhibitors wortmannin and LY294002 completely abrogated arsenite activation of p70(s6k), ERK1/2 activation by arsenite was not affected by these inhibitors, indicating that H(2)O(2) might act as an upstream molecule of PI3K as well as ERK1/2. Consistent with these results, none of the inhibitors impaired H(2)O(2) production by arsenite. DNA binding activity of AP-1, downstream of ERK1/2, was also inhibited by catalase, N-acetyl-L-cysteine, and the MEK inhibitor PD98059, which significantly blocked arsenite activation of ERK1/2. Taken together, these studies provide insight into mechanisms of arsenite-induced tumor promotion and suggest that H(2)O(2) plays a critical role in tumor promotion by arsenite through activation of the ERK1/2 and p70(s6k) signaling pathways.  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