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 46 毫秒
1.
The cis-isomer of parinaric acid, a naturally occurring C-18 polyene fatty acid, was incubated with brain subcellular fractions and the polarization of fluorescence increased in a time dependent manner. Greatest increases occurred in synaptosomal and microsomal membranes. This increase in polarization of fluorescence was found with the cis, but not the trans, isomer of parinaric acid and required Mg2+ or Ca2+ and was stimulated by coenzyme A and ATP. Synaptosomes were incubated with cis-parinaric acid and lipids were extracted and examined by high performance liquid chromatography. The highest incorporations of cis-parinaric acid were found in phosphatidylcholine (71%) and phosphatidylethanolamine (20%) while only traces were found in phosphatidylserine and phosphatidylinositol. [3H]Oleic acid was also incorporated into membrane phospholipids and unlabeled oleic acid blocked incorporation of cis-parinaric acid. It is proposed that cis-parinaric acid, like fatty acids normally found in brain, is incorporated into membrane phospholipids by an acyl-CoA acyltransferase. The presence of this enzyme in nervous tissue may make it possible to easily introduce fluorescent fatty acid probes into membrane phospholipids and to thereby facilitate study of membrane-mediated processes.  相似文献   

2.
Loss of latency due to membrane lipid peroxidation induced in vitro was studied in highly purified rat liver lysosomes. Enriched fractions of lysosomes were isolated by free flow electrophoresis. Lipid peroxidation of lysosomes, assayed as malondialdehyde formation, was catalyzed by a radical generating system consisting of dihydroxyfumaric acid and Fe3+-ADP. The peroxidation reaction occurred readily at 37 degrees C and reached a plateau at 10 min; however, the loss of lysosomal latency, determined as increased percentage free beta-N-acetylglucosaminidase activity, occurred more gradually and reached a maximum after 30 min. Scavengers of superoxide, hydrogen peroxide, singlet oxygen, and hydroxyl radicals did not inhibit the peroxidation reaction nor prevent the loss of lysosomal latency. However, preincubation of the lysosomes with alpha-tocopherol effectively blocked the induction of peroxidation and substantially reduced the loss of lysosomal latency. These results indicate that the lysosomal membrane is susceptible to free radical-induced lipid peroxidation; further, this process may be the immediate cause of the subsequent disintegration of the lysosome. The nature of the protective effect of alpha-tocopherol is unclear but may be due to its interaction with the unsaturated membrane lipids and the subsequent interruption of the chain-reaction initiated by free radicals.  相似文献   

3.
The effect of alpha-tocopherol on the lipid fluidity of porcine intestinal brush-border membranes was studied using pyrene as a fluorescent probe. Addition of alpha-tocopherol to the medium decreased fluorescence intensity and lifetime, but increased the fluorescence polarization of pyrene-labeled membranes. beta-, gamma-, and delta-Tocopherols gave no appreciable effect on the fluorescence intensity and polarization of the complex. The apparent dissociation constant (3.1 +/- 0.12 microM) of the interaction of alpha-tocopherol with the membranes, estimated from the change in the fluorescence intensity with varying concentrations of alpha-tocopherol, was in good agreement with the concentration required to cause the half-maximal inhibition of lipid peroxidation of the membranes performed by incubation with 100 microM ascorbic acid and 10 microM Fe2+. Decrease of the slope in the thermal Perrin plot of the polarization of pyrene-labeled membranes by alpha-tocopherol suggests that the movement of pyrene molecules in the membranes is restricted by binding of the tocopherol. This interpretation was confirmed by an increased harmonic mean of the rotational relaxation time of the dye molecules in the membranes from 10.9 +/- 0.16 to 18.5 +/- 0.51 microseconds after addition of 25 microM alpha-tocopherol to the medium. The perturbation of lipid phase in the membranes induced by alpha-tocopherol was also suggested from a decreased quenching rate constant of pyrene fluorescence in the membranes for Tl+. Based on these results, the effect of alpha-tocopherol on the lipid fluidity of the membranes is discussed.  相似文献   

4.
Ethanol-Induced Cell Death by Lipid Peroxidation in PC12 Cells   总被引:8,自引:0,他引:8  
Free radical generation is hypothesized to be the cause of alcohol-induced tissue injury. Using fluorescent cis-parinaric acid and TBARS, lipid peroxidation was shown to be increased in the presence of trace amounts of free ferrous ion in PC12 cells. This increase in lipid peroxidation was enhanced by ethanol in a dose dependent manner and also correlated with loss of cell viability, as measured by increased release of lactate dehydrogenase (LDH). Resveratrol, a potent antioxidant, had a protective effect against lipid peroxidation and cell death. These findings strongly suggest that ethanol-induced tissue injury and cell death is a free radical mediated process, and may be important in alcohol-related premature aging and other degenerative diseases.  相似文献   

5.
The lipid-phase structures of brush border membrane vesicles (BBMV) and basolateral membrane vesicles (BLMV) isolated from rabbit renal cortex were compared by steady-state and phase-modulation measurements of diphenylhexatriene (DPH) and trans- and cis-parinaric acid (tPnA and cPnA) fluorescence. A temperature-scanning system was used which gave reproducible temperature profiles of steady-state and dynamic fluorescence parameters with a resolution of 0.1 degrees C. Steady-state anisotropy of DPH showed a triphasic dependence on temperature with slope discontinuities at 22 +/- 4 and 47 +/- 3 degrees C (BBMV) and at 23 +/- 2 and 48 +/- 1 degrees C (BLMV). At all temperatures, DPH anisotropy in BBMV was greater than that in BLMV. Ground-state heterogeneity analysis of tPnA and cPnA fluorescence lifetime data demonstrated the presence of long (greater than 12 ns) and short (less than 5 ns) lifetime components, interpreted in terms of solid-phase and fluid-phase lipid domains. The fraction of solid-phase phospholipid decreased from 0.9 to 0.1 for BBMV and from 0.7 to 0.3 in BLMV with increasing temperature (10-50 degrees C). In both membranes, tryptophan-PnA fluorescence energy-transfer measurements showed that membrane proteins were surrounded by a fluidlike phospholipid phase. These results demonstrate the inadequacy of steady-state DPH anisotropy data in defining the structural characteristics of complex biological membranes. Results obtained with the phase-sensitive parinaric acid probes demonstrate major differences in the phase structure of the two opposing cell membranes in both the bulk lipid and the lipid microenvironment around membrane proteins.  相似文献   

6.
Rat liver microsomal membranes were exposed to either beta-nicotinamide adenine dinucleotide phosphate (NADPH), adenosine 5'-diphosphate (ADP), and Fe+3 or to azocompounds, and the antioxidant activities of beta-carotene and alpha-tocopherol were studied. Lipid peroxidation was monitored either by malondialdehyde (MDA) formation in the thiobarbituric acid assay at 535 nm or by hydroperoxide formation at 234 nm, after high-pressure liquid chromatography (HPLC) separation of phospholipid hydroperoxides. The radical initiators, water-soluble 2,2'-azobis(2-amidinopropane) (AAPH) and lipid-soluble 2,2'-azobis(2,4-dimethylvaleronitrile (AMVN), when thermally decomposed at 37 degrees C under air, produced a constant rate of lipid peroxidation in microsomes and lag times inversely related to their concentrations. Using 25 mM AAPH, beta-carotene suppressed lipid peroxidation at a concentration of 50 nmol/mg protein; using 24 mM AMVN, an inhibition of MDA formation was observed at a concentration of only 5 nmol/mg protein. Inhibition by beta-carotene did not produce a clearly defined lag phase. During AAPH-induced lipid peroxidation, beta-carotene was consumed linearly, and high levels of the antioxidant were still present at the end of 45 min of incubation. Using NADPH/ADP/Fe+3, protection by beta-carotene was observed at 10 nmol/mg protein. alpha-Tocopherol effectively suppressed both MDA and hydroperoxide formation in a dose-dependent manner when either NADPH/ADP/Fe+3 or azocompounds were used. These effects were observed at very low concentrations of the added alpha-tocopherol, ranging from 2 to 3 nmol/mg protein. When the lag times were measurable (AAPH and AMVN), they were directly proportional to the concentration of alpha-tocopherol and revealed the presence of endogenous antioxidants in the microsomal membranes. Different temporal relationships between the loss of alpha-tocopherol and lipid peroxidation were observed in relation to the prooxidant used. A substantial depletion of about 70% of endogenous alpha-tocopherol preceded the propagation phase when induced by the azocompounds, while only 20% of antioxidant disappeared at the beginning of the peroxidation when induced by NADPH/ADP/Fe+3. Although our results show that both beta-carotene and alpha-tocopherol suppress the peroxidation of microsomal membranes, their antioxidant efficacy is influenced by several factors, including the type of radical initiator involved and the site and rate of radical production.  相似文献   

7.
Lipid peroxidation is a major factor in the pathogenesis of many disease states. To detect the initial stages of lipid peroxidation or evaluate antioxidant efficacy, cis-parinaric acid (cis-PnA) has been successfully used and thoroughly validated. However, cis-PnA is not very well suited for medium throughput screening of antioxidants in living cells. We recently introduced and validated a lipid peroxidation reporter molecule, C11-BODIPY(581/591). To further explore this probe, we evaluated the protective effect of 12 natural antioxidants in rat-1 fibroblasts subjected to 50 microM cumene-hydroperoxide using both probes. The same pecking order for the individual antioxidant efficacies was obtained: alpha-tocopherol approximately gamma-tocopherol > quercetin approximately lycopene > kaempferol > palm oil > hydroxy-tyrosol > > alpha-carotene = beta-carotene = lutein = tyrosol = chlorogenic acid. This validates the accuracy of the C11-BODIPY(581/591) method and shows that this assay is an accurate and highly flexible method for indexing lipid peroxidation or determining antioxidant efficacy in living cells in a medium throughput scenario. The antioxidant efficacy was compared with their one-electron reduction potential, hydrophobicity and Trolox C equivalent antioxidant capacity. Our results show that although these parameters are valuable for determining structure-function relationships, they have limited predictive value for antioxidant efficacy in vivo.  相似文献   

8.
Rat liver microsomal lipids in hexane solution were exposed to the lipid-soluble radical initiator, azobis-isobutyronitrile (AIBN), and the antioxidant activities of alpha-tocopherol and beta-carotene have been compared. Lipid peroxidation was monitored both by conjugated diene formation at 233 nm, and by malondialdehyde (MDA) formation in the thiobarbituric acid assay at 535 nm. Diene formation was continuous for at least 120 min in the presence of 85 micrograms/ml lipid and 4 mM AIBN. Both alpha-tocopherol and beta-carotene acted as chain-breaking antioxidants, suppressing lipid peroxidation and producing an induction period at concentrations as low as 0.5 and 8 microM, respectively. When both of these lipid-soluble antioxidants were present together, the oxidation was strongly suppressed and the induction period was the sum of the individual antioxidants, alpha-Tocopherol and beta-carotene also inhibited MDA generation. In the presence of 170 micrograms/ml lipid and 8 mM AIBN, beta-carotene exhibited an IC50 of 1.1 microM and inhibited completely at 15 microM. Using beta-carotene, an induction period was observed, although much less pronounced than with alpha-tocopherol. Furthermore, beta-carotene inhibited MDA production in a concentration-dependent manner and exhibited an IC50 of 50 microM. In addition, added beta-carotene delayed the radical-initiated destruction of the endogenous alpha-tocopherol and gamma-tocopherol in this system.  相似文献   

9.
Conjugated eicosapentaenoic acid (CEPA) and conjugated docosahexaenoic acid (CDHA) with triene structure, isomerized by alkaline treatment, showed intensive cytotoxicity with LD(50) at 12 and 16 microM, respectively, in DLD-1 cells (colorectal adenocarcinoma), while they had no effect on normal human fibroblast cell lines such as MRC-5, TIG-103, and KMS-6 cells. Cytotoxic action of CEPA and CDHA was also demonstrated in other tumor cell lines including HepG2, A549, MCF-7, and MKN-7 cells. alpha-Tocopherol suppressed cytotoxicity of CEPA and CDHA in tumor cells, and the cytotoxicity involved membrane phospholipid peroxidation. CEPA and CDHA induced DNA condensation and fragmentation in DLD-1 cells, indicating the involvement of apoptosis in this cytotoxic mechanism. Furthermore, previous reports have shown that lipid peroxidation product induces cell death, including apoptotic cell death in different cell lines. CEPA and CDHA have been demonstrated in cultured cells to cause cell death via lipid peroxidation and apoptosis in the absence of alpha-tocopherol.  相似文献   

10.
Recently we have demonstrated that extracellular ATP acts as an excitatory neurotransmitter and enhances cell death in the presence of ferrous ions. By using a newly developed cis-parinaric acid fluorescence technique, we demonstrated that ATP, in a dose dependent manner, enhanced the increased membrane lipid peroxidation in PC12 cells when cells were incubated with micromolar FeCl2/DTP. P2 purinoceptor agonists, α,β-methylene ATP and 2-methylthio-ATP, induced PC12 cell lipid peroxidation, but to a lesser extent than ATP. ATP-induced Ca2+ influx via P2 purinoceptor activation significantly increased the intracellular Ca2+ concentration, which may have triggered a free radical generating cascade(s), and led to membrane lipid peroxidation and cell death. Since oxidative stress has been implicated in certain neurodegenerative diseases such as aging, extracellular ATP may contribute to neuronal cell death by an oxidative mechanism involving lipid peroxidation.  相似文献   

11.
We investigated the growth inhibitory effect of conjugated linoleic acid (CLA) on HepG2 (human hepatoma cell line), exploring whether the inhibitory action occurs via lipid peroxidation in the cells. When the cells were incubated up to 72 h with 5-40 microM of CLA (a mixture of 9c,11t-18:2 and 10t,12c-18:2), cell proliferation was clearly inhibited in a dose and time dependent manner but such an inhibition was not confirmed with linoleic acid (LA). In order to evaluate the possible contribution of lipid peroxidation exerted by CLA to cell growth inhibition, alpha-tocopherol (5-20 microM) and BHT (1-10 microM) as potent antioxidants were added to the medium with CLA (20 microM), which did not restore cell growth at all. Furthermore, after 72 h incubation, the membranous phospholipid hydroperoxide formation in the CLA-supplemented cells was suppressed respectively to 25% and 50% of that in LA-supplemented cells and control cells. No difference was observed by a conventional lipid peroxide assay, the TBA test, between CLA-supplemented cells and LA-supplemented cells. Although the cellular lipid peroxidation was not stimulated, lipid contents (triacylglycerol, total cholesterol and free cholesterol) and fatty acid contents (palmitic acid, palmitoleic acid and stearic acid) markedly increased in CLA-supplemented cells compared with LA-supplemented and control cells. Moreover, supplementation with 20 microM LA and 20 microM arachidonic acid profoundly interfered with the inhibitory effect of CLA in HepG2. These results suggest that the growth inhibitory effect of CLA on HepG2 is due to changes in fatty acid metabolism but not to lipid peroxidation.  相似文献   

12.
Diethyl maleate (DEM) (5 mM) and ethyl methanesulfonate (EMS) (35 mM) treatments rapidly depleted cellular reduced glutathione (GSH) below detectable levels (1 nmol/10(6) cells), and induced lipid peroxidation and necrotic cell death in freshly isolated rat hepatocytes. In hepatocytes incubated with 2.5 mM DEM and 10 mM EMS, however, the complete depletion of cellular GSH observed was not sufficient to induce lipid peroxidation or cell death. Instead, DEM- and EMS-induced lipid peroxidation and cell death were dependent on increased reactive oxygen species (ROS) production as measured by increases in dichlorofluorescein fluorescence. The addition of antioxidants (vitamin E succinate and deferoxamine) prevented lipid peroxidation and cell death, suggesting that lipid peroxidation is involved in the sequence of events leading to necrotic cell death induced by DEM and EMS. To investigate the subcellular site of ROS generation, the cytochrome P450 inhibitor, SKF525A, was found to reduce EMS-induced lipid peroxidation but did not protect against the loss of cell viability, suggesting a mitochondrial origin for the toxic lipid peroxidation event. In agreement with this conclusion, mitochondrial electron transport inhibitors (rotenone, thenoyltrifluoroacetone and antimycin A) increased EMS-induced lipid peroxidation and cell death, while the mitochondrial uncoupler, carbonyl cyanide m-chlorophenylhydrazone, blocked EMS- and DEM-mediated ROS production and lipid peroxidation. Furthermore, EMS treatment resulted in the significant loss of mitochondrial alpha-tocopherol shortly after its addition, and this loss preceded losses in cellular alpha-tocopherol levels. Treatment of hepatocytes with cyclosporin A, a mitochondrial permeability transition inhibitor, oxypurinol, a xanthine oxidase inhibitor, or BAPTA-AM, a calcium chelator, provided no protection against EMS-induced cell death or lipid peroxidation. Our results indicate that DEM and EMS induce cell death by a similar mechanism, which is dependent on the induction of ROS production and lipid peroxidation, and mitochondria are the major source for this toxic ROS generation. Cellular GSH depletion in itself does not appear to be responsible for the large increases in ROS production and lipid peroxidation observed.  相似文献   

13.
The decrease in fluorescence of conjugated polyenic acyl chains is used as a sensitive assay for lipid peroxidation. The fatty acid cis-trans-trans-cis-9,11,13, 15-octadecatetraenoic acid (cis-parinaric acid) is introduced into liposomal membranes as free fatty acid or, by using the PC specific transfer protein from bovine liver, as 1-palmitoyl-2-cis-parinaroyl-sn-glycero-3-phosphocholine. The peroxidation process as monitored by the decrease in fluorescence intensity is compared with other peroxidation assay systems. Applications of the new assay system are discussed.  相似文献   

14.
Cryopreservation is associated with the production of reactive oxygen species which lead to lipid peroxidation of sperm membranes. The objective was to determine an alpha-tocopherol concentration capable of improving the quality of cryopreserved porcine semen. Boar spermatozoa frozen with 200, 500 or 1000 microg/mL alpha-tocopherol were thawed and incubated at 37 degrees C for 4 h. Routine parameters of semen quality, susceptibility to lipid peroxidation 2-thiobarbituric acid (TBARS) and oxygen uptake were evaluated. Motility was higher (P<0.05) in samples treated with different concentrations of alpha-tocopherol up to 2 h of incubation. Viability and acrosome integrity significantly decreased during incubation (no significant differences between treatments). Two hundred micrograms per milliliter alpha-tocopherol protected spermatozoa against lipid peroxidation during incubation, but 1000 microg/mL failed to protect after 2 h of incubation. There was a negative association between TBARS and motility, suggesting that lipid peroxidation affected sperm motility. Both control and 200 microg/mL alpha-tocopherol samples preserved the capacity to generate oxidative energy up to 1 h of incubation. The addition of 200 microg/mL alpha-tocopherol in the semen extender could be useful to preserve boar spermatozoa against the oxidative stress generated by cryopreservation.  相似文献   

15.
The objective of this study was to compare the effect of alpha-tocopherol and its ester, alpha tocopherol succinate, on lipid peroxidation and motility of equine spermatozoa. In experiment one, spermatozoa were incubated with dl-alpha-tocopherol (5, 25, 100 or 500 microM), DL-alpha tocopherol succinate (5, 25, 100 or 500 microM) or vehicle (0.5% ethanol) at 38 degrees C, and sperm motility was determined at 30, 60 and 120 min. In experiment two, spermatozoa loaded with the lipophilic probe, C11BODIPY(581/591), were incubated with dl-alpha-tocopherol (50 and 100 microM), DL-alpha-tocopherol succinate (50 and 100 microM) or ethanol (0.5%) and with the promoters cumene hydroperoxide, Fe2SO4, and ascorbate at 38 degrees C in 5% CO2. Lipid peroxidation was determined by changes in fluorescence of C11BODIPY(581/591), and motility was determined by CASA at 0, 15, 30 and 60 min. In experiment three, spermatozoa loaded with C11BODIPY(581/591) were incubated with dl-alpha-tocopherol (5, 25, 100 or 500 microM), DL-alpha-tocopherol succinate (5, 25, 100 or 500 microM) or ethanol (0.5%) at 38 degrees C and then submitted to a 4-hour incubation at room temperature. Motility and lipid peroxidation were determined at 1 and 4 h. In experiment four, the effect of DL alpha tocopherol (5, 25 or 500 microM), DL-alpha-tocopherol succinate (5, 25 or 500 microM) or ethanol (0.5%) on lipid peroxidation and motility were evaluated during storage at 5 degrees C in a skim-milk based extender. Although dl-alpha-tocopherol succinate appeared more effective than DL-alpha-tocopherol in preventing lipid peroxidation during short-term incubations, the succinate ester suppressed sperm motility compared to dl-alpha-tocopherol alone.  相似文献   

16.
Summary In this article, I review the current information concerning the partition of the fluorescent probes, cis-parinaric acid (9, 11, 13, 15-cis, trans, trans, cis-octadecatetraenoic acid) and trans-parinaric acid (9, 11, 13, 15-all trans-octadecatetraenoic acid) among aqueous, solid lipid, and fluid lipid phases. The association of these probes with lipid is described by a mole fraction partition coefficient whose value is typically in the range of 1–5 × 106, a reasonable value in light of partition coefficients for other fatty acids between hydrophobic phases and water. The partition coefficient, in the absence of lipid phase changes, is relatively independent of temperature and only slightly dependent on the total aqueous probe concentration.In lipid samples which contain coexisting fluid and solid phases, trans-parinaric acid preferentially partitions into the solid phase, while cis-parinaric acid distributes nearly equally between fluid and solid phases. This partition behavior probably arises from the molecular shape of the cis and trans parinaric acid isomers. From measurements of the polarization of fluorescence of cis and trans parinaric acid in mixed lipid systems or membranes it is possible to evaluate the proportion of lipid components involved in phase changes or phase separation. From fluorescence energy transfer between protein typtophan residues and the parinaric acid isomers it is possible to gain information about the organization of lipids and proteins in membranes and model systems. I close the review by considering some of the membrane research areas where these probes and their various lipid derivatives may be particularly useful.  相似文献   

17.
Fluorescence in red cells following hydrogen peroxide treatment has been attributed to lipid peroxidation of the membrane. The putative relationship between lipid peroxidation and fluorescence was questioned by the finding that BHT and alpha-tocopherol, which are thought to inhibit lipid peroxidation, do not inhibit the fluorescence detected by flow cytometry. Furthermore, lipid peroxidation induced in red cells by the Fe(III)-ADP-ascorbate system did not produce fluorescence. These results require an alternative explanation for the hydrogen peroxide-induced fluorescence. A role for reduced hemoglobin is indicated by the inhibition of fluorescence by pretreatment of cells with CO that binds strongly to ferrohemoglobin and nitrite that oxidizes ferrohemoglobin. Our earlier studies have shown the formation of fluorescent heme degradation products during the reaction of purified hemoglobin with hydrogen peroxide, which was also inhibited by CO and nitrite pretreatment. The fluorescence produced in red cells after the addition of hydrogen peroxide can, therefore, be attributed to fluorescent heme degradation products.  相似文献   

18.
Treatment of the porcine intestinal brush-border membranes with 100 microM ascorbic acid and 10 microM Fe2+ in the presence of various concentrations of tert-butyl hydroperoxide (t-BuOOH) resulted in a marked fluorescence development at 430 nm, depending on the hydroperoxide concentration. This fluorescence formation was closely related to lipid peroxidation of the membranes as assessed by formation of conjugated diene. However there is no linear relation between thiobarbituric acid-reactive substances (TBARS) and fluorescence formation. On the other hand, fluorescence formation in the membranes by treatment with ascorbic acid/Fe2+ or t-BuOOH alone was negligible. The results with antioxidants and radical scavengers suggest that ascorbic acid/Fe2+/t-BuOOH-induced lipid peroxidation of the membranes is mainly due to t-butoxyl and/or t-butyl peroxy radicals. Most TBARS produced during the peroxidation reaction were released from the membranes, but fluorescent products remained in the membrane components. The fluorescence properties of products formed by lipid peroxidation of the membranes were compared with those of products derived from the interaction of malondialdehyde (MDA) or acetaldehyde with the membranes. The fluorescence products in the acetaldehyde-modified membranes also exhibited the emission maximum at 430 nm, while the emission maximum of MDA-modified membranes was 470 nm. The fluorescence intensity of MDA-modified membranes was markedly decreased by treatment with 10 mM NaBH4 but that of the peroxidized or acetaldehyde-modified membranes was enhanced by about two-fold with the treatment. In addition, a pH dependence profile revealed that the fluorescence intensity of the peroxidized or acetaldehyde-modified membranes decreases with increasing pH of the medium, whereas that of MDA-modified ones did not change over the pH range from 5.4 to 8.0. On the basis of these results, the fluorescence properties of products formed in the intestinal brush-border membranes by lipid peroxidation are discussed.  相似文献   

19.
H M Shapiro 《Cytometry》1981,2(3):143-150
The addition of RNA content estimation to flow cytometric measurement of DNA content provides valuable information concerning cells' transitions between quiescent and proliferative states. Equilibrium staining methods employing acridine orange have been used for DNA/RNA content measurement but are difficult to apply to intact cells and impractical for use in conjunction with fluorescent antibodies or ligands for demonstration of cell surface structures. I have used a combination of Hoechst 33342 (HO342) and pyronin Y (PY) to stain intact cells for DNA/RNA content estimation with a dual source flow cytometer using UV and blue-green or green excitation, measuring HO342 fluorescence at 430--470 nm and PY fluorescence at 590--650 nm. Results obtained with cultured cells and stimulated lymphocytes are in good agreement with those obtained using acridine orange for DNA/RNA staining; about half of the PY fluorescence can be removed from ethanol-fixed cells stained with HO342 and PY by RNAse digestion. The HO342/PY method can be combined with fluorescein immunofluorescence for detection of cell surface markers. HO342 can be combined with other tricyclic heteroaromatic dyes for DNA/RNA estimation; the combination of HO342 and oxazine 1 can be excited in a dual source instrument using a mercury arc lamp and a helium-neon laser. The staining procedure is simple; cells in medium are incubated with 5 microM HO342 at 37 degrees C for 45 min, 5 microM PY (or oxazine 1) is then added and cells are analyzed without washing after an additional 45 min incubation. Suitability of these dye combinations for vital cell staining and sorting remains to be determined.  相似文献   

20.
Effect of lipid physical state on the rate of peroxidation of liposomes.   总被引:2,自引:0,他引:2  
The effect of cholesterol on the rate of peroxidation of arachidonic acid and 1-palmitoyl-2-arachidonoyl phosphatidylcholine (PAPC) in dimyristoylphosphatidylcholine (DMPC) liposomes was examined above and below the phase transition temperature (Tm) of the lipid. The rate of peroxidation of arachidonic acid was more rapid below the phase transition temperature of the host lipid. At a temperature below the Tm (4 degrees C), increasing concentrations of cholesterol reduced the rate of peroxidation of arachidonic acid as judged by the production of thiobarbituric acid reactive substances. Above Tm (37 degrees C), cholesterol increased the rate of peroxidation of the fatty acid. Similarly, PAPC was peroxidized more rapidly at 4 degrees C than at 37 degrees C. However, cholesterol had little effect on the rate of peroxidation of PAPC at 4 degrees C. The rate of peroxidation of arachidonic acid was related to the lipid bilayer fluidity as judged by fluorescence anisotropy measurements of diphenylhexatriene. The rate of peroxidation increased slowly with increasing rigidity of the probe environment when the bilayer was relatively fluid and more rapidly as the environment became more rigid. The increase in the rate of peroxidation of arachidonic acid in the less fluid host lipid was unrelated to differences in iron binding or to transfer of arachidonic acid to the aqueous phase. Decreasing the concentration of arachidonic acid in DMPC to less than 2 mol% dramatically decreased the rate of peroxidation at 4 degrees C, suggesting that formation of clusters of fatty acids at 4 degrees C is required for rapid peroxidation.(ABSTRACT TRUNCATED AT 250 WORDS)  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