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.
Image cytometry was applied to study the intracellular localization of autofluorescence and the influence of an oxidative stress on this emission. K562 erythroleukemia cancer cells were analyzed with a microspectrofluorometer, coupled with a Argon laser (Ar+) (363 nm). From each cell, 15 x 15 emission spectra were recorded in the 400-600 nm spectral range to generate a spectral image of autofluorescence. The intracellular locations of the autofluorescence emission and of the specific mitochondrial probe rhodamine 123 (R123) were matched. Under a 363 nm excitation, all spectra from K562 cells show equivalent profiles with a 455 nm maximum emission, near of reduced nicotinamide adenine dinucleotide-(Phosphate) solution (NAD(P)H) (465 nm maximum emission). The spatial distribution of autofluorescence is homogeneous and different from the one of R123. Hydrogen peroxide (H2O2) (200 microM) and menadione (Men) (5 microM) induce a weak spectral change and a decrease in autofluorescence intensity, down to 40% of the initial emission. Doxorubicin (Dox) induces a dose-dependent decrease in autofluorescence emission and a release of intracellular free radicals. When cells were pre-treated 1 h with 1 mM glutathione (GSH), Dox induces a lower free radicals release, no significant variation of autofluorescence intensity and a lower growth inhibitory effect. Images cytometry of autofluorescence suggest that the intracellular NAD(P)H would not be restricted to mitochondrial compartments. The release of free radicals was associated with a decrease in autofluorescence intensity, mainly attributed to NAD(P)H oxidation both inside and outside mitochondria.  相似文献   

2.
Two-photon excitation fluorescence microscopy (TPEFM) permits the investigation of the topology of intercellular events within living animals. TPEFM was used to monitor the distribution of mitochondrial reduced nicotinamide adenine dinucleotide (NAD(P)H) in murine skeletal muscle in vivo. NAD(P)H fluorescence emission was monitored (~460 nm) using 710–720 nm excitation. High-resolution TPEFM images were collected up to a depth of 150 μm from the surface of the tibialis anterior muscle. The NAD(P)H fluorescence images revealed subcellular structures consistent with subsarcolemmal, perivascular, intersarcomeric, and paranuclear mitochondria. In vivo fiber typing between IIB and IIA/D fibers was possible using the distribution and content of mitochondria from the NAD(P)H fluorescence signal. The intersarcomeric mitochondria concentrated at the Z-line in the IIB fiber types resulting in a periodic pattern with a spacing of one sarcomere (2.34 ± 0.17 μm). The primary inner filter effects were nearly equivalent to water, however, the secondary inner filter effects were highly significant and dynamically affected the observed emission frequency and amplitude of the NAD(P)H fluorescence signal. These data demonstrate the feasibility, and highlight the complexity, of using NAD(P)H TPEFM in skeletal muscle to characterize the topology and metabolic function of mitochondria within the living mouse.  相似文献   

3.
The blue autofluorescence (351 nm excitation, 450 nm emission) of single skeletal muscle fibers from Xenopus was characterized to be originating from mitochondrial NAD(P)H on the basis of morphological and functional correlations. This fluorescence signal was used to estimate the oxygen availability to isolated single Xenopus muscle fibers during work level transitions by confocal microscopy. Fibers were stimulated to generate two contractile periods that were only different in the PO2 of the solution perfusing the single fibers (PO2 of 30 or 0-2 Torr; pH = 7.2). During contractions, mean cellular NAD(P)H increased significantly from rest in the low PO2 condition with the core (inner 10%) increasing to a greater extent than the periphery (outer 10%). After the cessation of work, NAD(P)H decreased in a manner consistent with oxygen tensions sufficient to oxidize the surplus NAD(P)H. In contrast, NAD(P)H decreased significantly with work in 30 Torr PO2. However, the rate of NAD(P)H oxidation was slower and significantly increased with the cessation of work in the core of the fiber compared with the peripheral region, consistent with a remaining limitation in oxygen availability. These results suggest that the blue autofluorescence signal in Xenopus skeletal muscle fibers is from mitochondrial NAD(P)H and that the rate of NAD(P)H oxidation within the cell is influenced by extracellular PO2 even at high extracellular PO2 during the contraction cycle. These results also demonstrate that although oxygen availability influences the rate of NAD(P)H oxidation, it does not prevent NAD(P)H from being oxidized through the process of oxidative phosphorylation at the onset of contractions.  相似文献   

4.
Microspectrofluorometry of cell coenzymes (NAD(P)H, flavins) in conjunction with sequential microinjections into the same cell of metabolites and modifiers, reveals aspects of the regulatory mechanisms of transient redox changes of mitochondrial and extramitochondrial nicotinamide adenine dinucleotides. The injection of ADP in the course of an NAD(P)H transient produced by glycolytic (e.g. glucose 6-phosphate, G6P) or mitochondrial (e.g. malate) substrate leads to sharp reoxidation (state III, Chance and Williams, 1955), followed by a spontaneous state III to IV transition, and an ultimate return to original redox steady state. The response to ADP alone is biphasic, i.e. a small oxidation-reduction transient followed by a larger reverse transient. Similarities between responses to injected ATP and ADP suggest possible intracellular interconversions. Sequential injections of glycolytic and Krebs cycle substrates into the same cell, produce a two-step NAD(P) response, possibly revealing the intracellular compartmentation of this coenzyme. A two-step NAD(P)H response to sequentially injected fructose 1,6-diphosphate and G6P indicates the dynamic or even structural compartmentation of glycolytic phosphate esters in separate intracellular pools. The intracellular regulation and compartmentation of bioenergetic pathways and cell-to-cell metabolic inhomogeneities provide the basis on which the quantitative biochemistry of the intact living cell may be reconciled with these in situ findings.  相似文献   

5.
Summary The blue fluorescence emitted by microbial cells irradiated with UV light at 360 nm is usually supposed to provide a good estimate of the cell NAD(P)H content. Here we present an example of a microbial fermentation in which culture fluorescence, both in the cells and in the medium, was almost exclusively due to the presence of a fluorophore that displayed an emission spectrum very similar to that of NAD(P)H but that we show by biochemical studies to be a different compound. Our results demonstrate that studies on the redox state of cells should be based on on-line fluorescence data only after appropriate control experiments to establish a definitive correlation between fluorescence and NAD(P)H levels. Offprint requests to: J. E. Bailey  相似文献   

6.
Two-photon (2P) ratiometric redox fluorometry and microscopy of pyridine nucleotide (NAD(P)H) and flavoprotein (FP) fluorescence, at 800-nm excitation, has been demonstrated as a function of mitochondrial metabolic states in isolated adult dog cardiomyocytes. We have measured the 2P-excitation spectra of NAD(P)H, flavin adenine dinucleotide (FAD), and lipoamide dehydrogenase (LipDH) over the wavelength range of 720-1000 nm. The 2P-excitation action cross sections (sigma2P) increase rapidly at wavelengths below 800 nm, and the maximum sigma2P of LipDH is approximately 5 and 12 times larger than those of FAD and NAD(P)H, respectively. Only FAD and LipDH can be efficiently excited at wavelengths above 800 nm with a broad 2P-excitation band around 900 nm. Two autofluorescence spectral regions (i.e., approximately 410-490 nm and approximately 510-650 nm) of isolated cardiomyocytes were imaged using 2P-laser scanning microscopy. At 750-nm excitation, fluorescence of both regions is dominated by NAD(P)H emission, as indicated by fluorescence intensity changes induced by mitochondrial inhibitor NaCN and mitochondria uncoupler carbonyl cyanide p-(trifluoromethoxy) phenyl hydrazone (FCCP). In contrast, 2P-FP fluorescence dominates at 900-nm excitation, which is in agreement with the sigma2P measurements. Finally, 2P-autofluorescence emission spectra of single cardiac cells have been obtained, with results suggesting potential for substantial improvement of the proposed 2P-ratiometric technique.  相似文献   

7.
Mitochondrial production of reactive oxygen species (ROS) at Complex I of the electron transport chain is implicated in the etiology of neural cell death in acute and chronic neurodegenerative disorders. However, little is known regarding the regulation of mitochondrial ROS production by NADH-linked respiratory substrates under physiologically realistic conditions in the absence of respiratory chain inhibitors. This study used Amplex Red fluorescence measurements of H2O2 to test the hypothesis that ROS production by isolated brain mitochondria is regulated by membrane potential (DeltaPsi) and NAD(P)H redox state. DeltaPsi was monitored by following the medium concentration of the lipophilic cation tetraphenylphosphonium with a selective electrode. NAD(P)H autofluorescence was used to monitor NAD(P)H redox state. While the rate of H2O2 production was closely related to DeltaPsi and the level of NAD(P)H reduction at high values of DeltaPsi, 30% of the maximal rate of H2O2 formation was still observed in the presence of uncoupler (p-trifluoromethoxycarbonylcyanide phenylhydrazone) concentrations that provided for maximum depolarization of DeltaPsi and oxidation of NAD(P)H. Our findings indicate that ROS production by mitochondria oxidizing physiological NADH-dependent substrates is regulated by DeltaPsi and by the NAD(P)H redox state over ranges consistent with those that exist at different levels of cellular energy demand.  相似文献   

8.
The soluble hydrogenase (hydrogen: NAD+ oxidoreductase, EC 1.12.1.2) from Alcaligenes eutrophus H 16 was purified 68-fold with a yield of 20% and a final specific activity (NAD reduction) of about 54 mumol H2 oxidized/min per mg protein. The enzyme was shown to be homogenous by polyacrylamide gel electrophoresis. Its molecular weight and isoelectric point were determined to be 205 000 and 4.85 respectively. The oxidized hydrogenase, as purified under aerobic conditions, was of high stability but not reactive. Reductive activation of the enzyme by H2, in the presence of catalytic amounts of NADH, or by reducing agents caused the hydrogenase to become unstable. The purified enzyme, in its active state, was able to reduce NAD, FMN, FAD, menaquinone, ubiquinone, cytochrome c, methylene blue, methyl viologen, benzyl viologen, phenazine methosulfate, janus green, 2,6-dichlorophenoloindophenol, ferricyanide and even oxygen. In addition to hydrogenase activitiy, the enzyme exhibited also diaphorase and NAD(P)H oxidase activity. The reversibility of hydrogenase function (i.e. H2 evolution from NADH, methyl viologen and benzyl viologen) was demonstrated. With respect to H2 as substrate, hydrogenase showed negative cooperativity; the Hill coefficient was n = 0.4. The apparent Km value for H2 was found to be 0.037 mM. The absorption spectrum of hydrogenase was typical for non-heme iron proteins, showing maxima (shoulders) at 380 and 420 nm. A flavin component could be extracted from native hydrogenase characterized by its absorption bands at 375 and 447 nm and a strong fluorescense at 526 nm.  相似文献   

9.
Paclitaxel, a widely used antimicrotubular agent, predominantly eliminates rapidly proliferating cancer cells, while slowly proliferating and quiescent cells can survive the treatment, which is one of the main reasons for tumor recurrence and non-responsiveness to the drug. To improve the efficacy of chemotherapy, biomarkers need to be developed to enable monitoring of tumor responses. In this study we considered the auto-fluorescent metabolic cofactors NAD(P)H and FAD as possible indicators of cancer cell response to therapy with paclitaxel. It was found that, among the tested parameters (the fluorescence intensity-based redox ratio FAD/NAD(P)H, and the fluorescence lifetimes of NAD(P)H and FAD), the fluorescence lifetime of NAD(P)H is the most sensitive in tracking the drug response, and is capable of indicating heterogeneous cellular responses both in cell monolayers and in multicellular tumor spheroids. We observed that metabolic reorganization to a more oxidative state preceded the morphological manifestation of cell death and developed faster in cells that were more responsive to the drug. Our results suggest that noninvasive, label-free monitoring of the drug-induced metabolic changes by noting the NAD(P)H fluorescence lifetime is a valuable approach to characterize the responses of cancer cells to anti-cancer treatments and, therefore, to predict the effectiveness of chemotherapy.  相似文献   

10.
The low potential c-type cytochrome from the phototrophic purple sulphur bacterium Thiocapsa roseopersicina, strain BBS was isolated in electrophoretically homogeneous state. The bulk of the cytochrome (approximately 90%) after disruption of the cells remained in the membrane fraction. The absorption spectrum of the cytochrome was characterized by the maxima at 420, 523 and 552 nm in the reduced state and at 408 nm in the oxidized one. The cytochrome interacted with CO in the reduced state. The molecular weight of the cytochrome is 50 000. The cytochrome contains great amounts of phenylalanine, leucine, valine, aspartic and glutamic acids and can be reduced by dithionite but not by cysteine, sulfide or ascorbate. Besides, the cytochrome can also be reduced by NAD(P)H in the presence of NAD(P)-reductases of T. roseopersicina, when ferredoxin of Spirulina platensis or benzyl viologen are added to the reaction mixture. The cytochrome can act as an electron donor (acceptor) for T. roseopersicina hydrogenase.  相似文献   

11.
Rapid microspectrofluorometry has been used to evaluate 1-pyrene-butyric acid as an oxygen probe in single living EL2 ascites tissue culture cells. Despite instrumental conditions preventing detection of the pyrene butyric acid maxima at 380 and 400 nm, the probe having penetrated the cell can be easily identified (maximum around 440 nm in unconnected spectra) from the fluorescence emission spectrum, as compared with NAD(P)H emission in controls (maximum around 460 nm). Fluorescence changes during gradually increasing anaerobiosis under nitrogen flow, are compatible with a linear relationship between the reciprocal of the fluorescence intensity and the intracellular oxygen concentration (increase in 430, 434, 442/461 nm ratios at anaerobiosis). The cells having absorbed the probe continue to catabolize glycolytic substrate, but some inhibition is noticeable (e.g. from the amplitude of the NAD(P)H fluorescence increase spectrum due to intracellular addition of glucose-6-P). In principle rapid microspectrofluorometry allows a multiprobe (e.g. 1-pyrene-butyric acid for oxygen, vs NAD(P)H for metabolism) exploration of the living cell.  相似文献   

12.
A multiwavelength fluorescence probe is proposed for in situ monitoring of Eschscholtzia californica and Catharanthus roseus plant cell cultures. The potential of the probe as a tool for real-time estimation of biomass and production in secondary metabolites has been studied. The probe excitation range is 270-550 nm and the emission range is 310-590 nm, with a step of 20 nm for both excitation and emission filters. Many endogenous fluorophores such as NAD(P)H, riboflavins (riboflavin and derivatives such as FMN, FAD), tryptamine and tryptophan, and fluorescent secondary metabolites were analyzed simultaneously. NAD(P)H fluorescence signal (350/450 nm) showed to be an adequate signal for estimating cells activity. Riboflavins fluorescence signal (450/530 nm) followed C. roseus cell concentration both for the growth phase and after elicitation with jasmonic acid. Fluorescence from the alkaloids interfered with NAD(P)H signal during the production phase. For C. roseus, tryptophan, tryptamine, ajmalicine and serpentine were monitored by the probe. For E. californica, fluorescence from alkaloids overlapped with riboflavins preventing from using the probe to follow cell growth but global alkaloids production could be followed using the probe.  相似文献   

13.
Lambert AJ  Buckingham JA  Brand MD 《FEBS letters》2008,582(12):1711-1714
The relationship between the rate of superoxide production by complex I and NAD(P)H redox state was investigated in rat skeletal muscle mitochondria. A high rate of superoxide production was observed during succinate oxidation; the rate during pyruvate oxidation was over fourfold lower. However, the NAD(P)H pool was significantly less reduced during succinate oxidation than during pyruvate oxidation. We conclude that there is no unique relationship between superoxide production by complex I and the reduction state of the NAD(P)H pool. Our data suggest that less than 10% of the superoxide originates from the flavin site during reverse electron transport from succinate.  相似文献   

14.
A fiber-optic probe was interfaced to an analytical spectrofluorophotometeru and used to measure NAD(P)H fluorescence of hybridoma cells in a bioreactor. NAD(P)H fluorescence was found to qualitatively represent metabolic state during various induced metabolic transitions. NAD(P)H fluorescence increased immediately following aerobic-anaerobic transitions, and decreased immediately upon transition back to aerobic metabolism. Pulsing of glucose to glucose-depleted cultures caused NAD(P)H fluorescence to first increase immediately after the pulse, and then decrease gradually before reaching a constant level. Pulsing of glutamine to glutamine-depleted cultures resulted in a gradual increase in NAD(P)H fluorescence which lagged a simultaneous increase in oxygen uptake. ATP production and oxygen uptake also varied with metabolic state. The decrease in oxidative phosphorylation following transition from aerobic to anaerobic metabolism was found to be only partially compensated by the concomitant increase in substrate-level phosphorylation, as shown by decreases of 35-52% in calculated total specific ATP production rates. The specific oxygen uptake rate decreased by 6-38% following glucose pulses of between 0.2 and 0.5 g/L, respectively, and by 50% following glutamine depletion. Subsequent pulsing of glutamine after depletion caused oxygen uptake to increase by 50%.  相似文献   

15.
Incubation of rat liver mitochondria with benzoquinone derivatives in the presence of succinate plus rotenone has been shown to cause NAD(P)H oxidation followed by Ca2+ release. Further investigation revealed: (1)p-Benzoquinone-induced Ca2+ release was not initiated by a collapse of the mitochondrial membrane potential. However, Ca2+ release and subsequent Ca2+ cycling caused limited increased membrane permeability. (2) p-Benzoquinone-induced NAD(P)H oxidation and Ca2+ release were prevented by isocitrate, 3-hydroxybutyrate, and glutamate but not by pyruvate or 2-oxoglutarate. (3) Inhibition of pyruvate and 2-oxoglutarate dehydrogenases by p-benzoquinone was attributed to arylation of the SH groups of the cofactors, CoA and lipoic acid. Isocitrate dehydrogenase was also inhibited by p-benzoquinone, but the cofactors NAD(P)H and Mn2+ protected the enzyme. Glutamate dehydrogenase was not inhibited by p-benzoquinone. (4) Arylation of mitochondrial protein thiols by p-benzoquinone was associated with an inhibition of state 3 respiration, which was attributed to the inactivation of the phosphate translocase. In contrast, state 4 respiration, and the F1.F0-ATPase and ATP/ADP translocase activities were not inhibited. It was concluded that inhibition of mitochondrial NAD(P)H dehydrogenases by arylation of critical thiol groups will decrease the NAD(P)+-reducing capacity, and possibly lower the NAD(P)H/NAD(P)+ redox status in favor of Ca2+ release.  相似文献   

16.
Cellular autofluorescence was characterized in normal human esophageal cells and in malignant esophageal epithelial cells. The study was performed under excitation at 351 nm where the cell fluorescence is mainly due to the reduced pyridine nucleotides (NAD(P)H) with a very small contribution from the oxidized flavins (FMN, FAD) or lipopigments. The autofluorescence emission of squamous cell carcinoma, adenocarcinoma on Barrett's mucosa and normal cells was characterized by microspectrofluorimetry on monolayers and by spectrofluorimetry on cell suspensions. The relative contribution of each fluorophore to the fluorescence emission of the different cell types was evaluated by a curve-fitting analysis. A statistically highly significant difference was observed between the average intensity of the raw spectra of the different cell types. Tumoral cells had a fluorescence intensity approximately twice as high as that of normal cells. The results of the NAD(P)H quantitation analyzed by microspectrofluorimetry on single living cells and spectrofluorimetry on cell suspensions were consistent with those obtained by biochemical cycling assays, showing that the amount of intracellular NAD(P)H is higher in tumoral cells than in normal cells. Bound NAD(P)H concentration was found to be quite stable whatever the cell type while the amount of free NAD(P)H showed a very important increase in tumoral cells.  相似文献   

17.
Aralar, the mitochondrial aspartate-glutamate carrier present in beta-cells, is a component of the malate-aspartate NADH shuttle (MAS). MAS is activated by Ca2+ in mitochondria from tissues with aralar as the only AGC isoform with an S0.5 of approximately 300 nm. We have studied the role of aralar and its Ca2+-binding EF-hand motifs in glucose-induced mitochondrial NAD(P)H generation by two-photon microscopy imaging in INS-1 beta-cells lacking aralar or expressing aralar mutants blocked for Ca2+ binding. Aralar knock-down in INS-1 beta-cell lines resulted in undetectable levels of aralar protein and complete loss of MAS activity in isolated mitochondria and in a 25% decrease in glucose-stimulated insulin secretion. MAS activity in mitochondria from INS-1 cells was activated 2-3-fold by extramitochondrial Ca2+, whereas aralar mutants were Ca2+ insensitive. In Ca2+-free medium, glucose-induced increases in mitochondrial NAD(P)H were small (1.3-fold) and unchanged regardless of the lack of aralar. In the presence of 1.5 mm Ca2+, glucose induced robust increases in mitochondrial NAD(P)H (approximately 2-fold) in cell lines with wild-type or mutant aralar. There was a approximately 20% reduction in NAD(P)H response in cells lacking aralar, illustrating the importance of MAS in glucose action. When small Ca2+ signals that resulted in extremely small mitochondrial Ca2+ transients were induced in the presence of glucose, the rise in mitochondrial NAD(P)H was maintained in cells with wild-type aralar but was reduced by approximately 50% in cells lacking or expressing mutant aralar. These results indicate that 1) glucose-induced activation of MAS requires Ca2+ potentiation; 2) Ca2+ activation of MAS represents a larger fraction of glucose-induced mitochondrial NAD(P)H production under conditions where suboptimal Ca2+ signals are associated with a glucose challenge (50 versus 20%, respectively); 3) inactivation of EF-hand motifs in aralar prevents activation of MAS by small Ca2+ signals. The results suggest a possible role for aralar and MAS in priming the beta-cell by Ca2+-mobilizing neurotransmitter or hormones.  相似文献   

18.
19.
In this study we examined the utility of NAD(P)H fluorescence for monitoring aerobic fermentations of the threonine auxotroph Corynebacterium glutamicum ATCC 14296. Instead of attempting complicated mathematical corrections for inner-filter effects, we found that it is possible to use the information contained in the on-line NAD(P)H fluorescence signal to assess culture metabolic activities during fermentation. The first derivative of the filtered fluorescence signal, which approximates the turnover rate of the NAD(P)H pool, can be used to precisely identify the temporal points of threonine and glucose exhaustion.  相似文献   

20.
Coenzyme fluorescence spectra of single living cells are due to free pyridine nucleotides (folded configuration), bound pyridine nucleotides (unfolded configuration) and a third component, possibly a mixture or flavins. Such spectra can be used to recognize possible differences in coenzyme composition between cell lines or changes of metabolic pathways due to chemicals acting at levels below or above cytotoxicity, by high resolution spectrofluorometry. A study of spectra recorded from cultured Ehrlich ascites (EL2), and Harding Passey melanoma cells (HPM-67 and HPM-73 line) grown under comparable conditions, shows that free NAD(P)H predominates in HPM-67 and EL2, while this coenzyme is bound in HPM-73. The free/bound ratio may be profoundly modifed by chemicals, e.g. in the HPM-73 increase of free and decrease of bound NAD(P)H occurred upon treatment with 10(-6) oligomycin. When atebrine at levels (10(-6) M) below cytotoxicity was added, there was a decrease of the free NAD(P)H spectrum possibly through energy transfer from NAD(P)H to atebrine. Consideration of long range energy transfer i.e., excitation of atebrine by fluorescence of NAD(P)H vs. short range transfer of excitation energy from free NAD(P)H to atebrine, favors the latter mechanism. A transient (reversible) increase in atebrine fluorescence is seen following intracellular microinjection of substrate (e.g. glucose-6-P) leading to an increase in free NAD(P)H. At cytotoxic levels of atebrine (e.g 2 x 10(-5) M) an irreversible increase of atebrine fluorescence is seen. The microspectrofluorometric technique appears therefore well suited to study physiological processes at the level of intracellular coenzymes, as well as possible processes of intermolecular energy transfer in the microenvironment.  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