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.
The decomposition of lipid hydroperoxides (LOOH) into peroxyl radicals is a potential source of singlet molecular oxygen ((1)O(2)) in biological systems. Recently, we have clearly demonstrated the generation of (1)O(2) in the reaction of lipid hydroperoxides with biologically important oxidants such as metal ions, peroxynitrite and hypochlorous acid. The approach used to unequivocally demonstrate the generation of (1)O(2) in these reactions was the use of an isotopic labeled hydroperoxide, the (18)O-labeled linoleic acid hydroperoxide, the detection of labeled compounds by HPLC coupled to tandem mass spectrometry (HPLC-MS/MS) and the direct spectroscopic detection and characterization of (1)O(2) light emission. Using this approach we have observed the formation of (18)O-labeled (1)O(2) by chemical trapping of (1)O(2) with anthracene derivatives and detection of the corresponding labeled endoperoxide by HPLC-MS/MS. The generation of (1)O(2) was also demonstrated by direct spectral characterization of (1)O(2) monomol light emission in the near-infrared region (lambda = 1270 nm). In summary, our studies demonstrated that LOOH can originate (1)O(2). The experimental evidences indicate that (1)O(2) is generated at a yield close to 10% by the Russell mechanism, where a linear tetraoxide intermediate is formed in the combination of two peroxyl radicals. In addition to LOOH, other biological hydroperoxides, including hydroperoxides formed in proteins and nucleic acids, may also participate in reactions leading to the generation (1)O(2). This hypothesis is currently being investigated in our laboratory.  相似文献   

2.
Site-specific induction of lipid peroxidation by iron in charged micelles   总被引:1,自引:0,他引:1  
Generation of hydroxyl radicals by the Fenton reaction resulted in lipid peroxidation of linoleic acid (LA) (H2O2-Fe2+-induced lipid peroxidation) in positively charged tetradecyltrimethylammonium bromide (TTAB) micelles, but not in negatively charged sodium dodecyl sulfate (SDS) micelles. However, more OH radicals formed via the Fenton reaction were trapped by N-t-butyl-alpha-phenylnitrone (PBN) in SDS micelles than in TTAB micelles. When detergent-dispersed LA was contaminated with linoleic acid hydroperoxide (LOOH), lipid peroxidation was catalyzed by Fe2+ via reductive cleavage of LOOH (LOOH-Fe2+-induced lipid peroxidation), and Fe2+ was oxidized simultaneously in SDS micelles, even when H2O2 was not present. In contrast, LOOH-Fe2+-induced lipid peroxidation and simultaneous oxidation of Fe2+ were not observed in TTAB micelles. An ESR spectrum presumed to be due to an alkoxy radical trapped by PBN was also detected in SDS micelles, but not in TTAB micelles in the LOOH-Fe2+-induced lipid peroxidation system. The results are discussed in the light of the localization of iron, the unsaturated bonding moiety of LA, the OOH-group of LOOH, and the trapping site of PBN in different charged micelles.  相似文献   

3.
Human A549 lung epithelial cells were challenged with 18O-labeled hydrogen peroxide ([18O]-H2O2), the total RNA and DNA extracted in parallel, and analyzed for 18O-labeled 8-oxo-7,8-dihydroguanosine ([18O]-8-oxoGuo) and 8-oxo-7,8-dihydro-2'-deoxyguanosine ([18O]-8-oxodGuo) respectively, using high-performance liquid chromatography electrospray ionization tandem mass spectrometry (HPLC-MS/MS). [18O]-H2O2 exposure resulted in dose-response formation of both [18O]-8-oxoGuo and [18O]-8-oxodGuo and 18O-labeling of guanine in RNA was 14-25 times more common than in DNA. Kinetics of formation and subsequent removal of oxidized nucleic acids adducts were also monitored up to 24 h. The A549 showed slow turnover rates of adducts in RNA and DNA giving half-lives of approximately 12.5 h for [18O]-8-oxoGuo in RNA and 20.7 h for [18O]-8-oxodGuo in DNA, respectively.  相似文献   

4.
alpha-Tocopherol inhibited H2O2-Fe2+-induced lipid peroxidation of linoleic acid (LA) by scavenging OH radicals in tetradecyltrimethylammonium bromide (TTAB) micelles. The inhibiting ability of alpha-tocopherol was much greater than that of OH-radical scavengers mannitol and t-butanol. In contrast, alpha-tocopherol enhanced linoleic acid hydroperoxide (LOOH)-Fe2+-induced lipid peroxidation through regeneration of Fe2+ in sodium dodecyl sulfate (SDS) micelles containing LA. alpha-Tocopherol was oxidized by Fenton's reagent (FeSO4 + H2O2) at a higher rate in SDS micelles than in TTAB micelles. The likely oxidants were OH radicals in the former and Fe3+ in the latter. Both reagents formed in the Fenton reaction. Ferrous ion catalyzed in a dose-dependent manner the decomposition of LOOH and conjugated diene compounds in SDS but not in TTAB micelles. alpha-Tocopherol and Fe3+ individually had no effect on the decomposition of LOOH, but together were quite effective. The rate of the decomposition was a function of the concentration of alpha-tocopherol. The mechanism of "site-specific" antioxidant action of alpha-tocopherol in charged micelles is discussed.  相似文献   

5.
Selected reaction monitoring (SRM)-MS is an emerging technology for high throughput targeted protein quantification and verification in biomarker discovery studies; however, the cost associated with the application of stable isotope-labeled synthetic peptides as internal standards can be prohibitive for screening a large number of candidate proteins as often required in the preverification phase of discovery studies. Herein we present a proof of concept study using an (18)O-labeled proteome reference as global internal standards (GIS) for SRM-based relative quantification. The (18)O-labeled proteome reference (or GIS) can be readily prepared and contains a heavy isotope ((18)O)-labeled internal standard for every possible tryptic peptide. Our results showed that the percentage of heavy isotope ((18)O) incorporation applying an improved protocol was >99.5% for most peptides investigated. The accuracy, reproducibility, and linear dynamic range of quantification were further assessed based on known ratios of standard proteins spiked into the labeled mouse plasma reference. Reliable quantification was observed with high reproducibility (i.e. coefficient of variance <10%) for analyte concentrations that were set at 100-fold higher or lower than those of the GIS based on the light ((16)O)/heavy ((18)O) peak area ratios. The utility of (18)O-labeled GIS was further illustrated by accurate relative quantification of 45 major human plasma proteins. Moreover, quantification of the concentrations of C-reactive protein and prostate-specific antigen was illustrated by coupling the GIS with standard additions of purified protein standards. Collectively, our results demonstrated that the use of (18)O-labeled proteome reference as GIS provides a convenient, low cost, and effective strategy for relative quantification of a large number of candidate proteins in biological or clinical samples using SRM.  相似文献   

6.
It is generally accepted that the availability of vitamin K in vivo depends on its homologues, the biological activities of which would differ among organs. To test this hypothesis, we examined the uptake, metabolism, and utilization of menaquinone-4 (MK-4) and phylloquinone (PK) using 18O-labeled compounds in two cultured human cell lines (HepG2 and MG-63). Lipid extracts were prepared from the cells and media after 1, 3, and 6h of incubation. The detection of the vitamin K analogues (18O-, 16O-quinone, and epoxide forms) was carried out with LC-APCI-MS/MS as previously reported. The 18O of vitamin K was replaced with atmospheric 16O2 during the formation of vitamin K epoxide with a carboxylative catalytic reaction. As a result, a significant difference was observed between MK-4 and PK in the amounts taken up into the cells. The 18O-labeled MK-4 was rapidly and remarkably well absorbed into the cells and metabolized to the epoxide form via a hydroquinone form as compared to the 18O-labeled PK. The difference in uptake of MK-4 and PK was not affected by treatment with warfarin although the metabolism of both compounds was markedly inhibited. This methodology should be utilized to clarify some of the actions of vitamin K in target cells and facilitate the development of new vitamin K drugs.  相似文献   

7.
Turbulent flow chromatograph (TFC) is a technique for the direct and efficient analysis of drugs and metabolites in biological matrices. We report here TFC on-line with an HPLC-MS/MS assay for the determination of 5-[2,4-dioxothiazolidin-5-yl)methyl]-2-methoxy-N-[[(4-trifluoromethyl)phenyl]methyl]benzamide (I, MK-0767, KRP297, Fig. 1) in plasma. Samples were transferred using an automated system followed by the addition of internal standard (II), prepared in 0.1 M ammonium acetate (pH 4.0). The plasma samples were directly injected onto a C18 turbulent flow column on-line with an HPLC-MS/MS system, and the analytical column used was a ThermoHypersil Keystone C18. Detection was achieved by MS/MS, using positive ionization on a TurboIonSpray probe, operated in multiple reaction monitoring (MRM) mode. The linear range was 4-2000 ng/mL for I when using 50 microL of plasma. The method exhibited good linearity and reproducibility. The method also showed good selectivity and ruggedness when applied to clinical samples, and was successfully cross-validated with a conventional off-line SPE, LC-MS/MS method.  相似文献   

8.
《Inorganica chimica acta》2006,359(9):2842-2849
The reaction between TpOs(N)Cl2 (1) [Tp = hydrotris(1-pyrazolyl)borate] and aqueous (nBu4N)(OH) in THF-d8 forms the nitrosyl complex TpOs(NO)Cl2 (5) among other products, suggesting an initial hydroxide attack at the nitrido ligand. In contrast, the reaction of the acetate complex TpOs(N)(OAc)2 (2) with NaOH in Me2CO/H2O yields the osmium bis-hydroxide complex TpOs(N)(OH)2 (3), which has been structurally characterized by single-crystal X-ray diffraction. Acetate for hydroxide exchange could occur by ligand substitution or by nucleophilic attack at the carbonyl carbon of the acetate ligands (saponification). Reacting 2 with Na18OH in H218O/CD3CN yields predominantly doubly 18O-labeled TpOs(N)(18OH)2 (3-18O2) and unlabeled acetate, by ESI/MS and 13C{1H} NMR. This indicates that hydroxide reacts by substitution rather than by attack at the ligand. The reaction of 2 with the softer nucleophile thiosulfate occurs at the nitrido ligand, giving the thionitrosyl complex TpOs(NS)(OAc)2 (4). Reacting 4 with NaOH in (CD3)2CO/D2O also generates the bis-hydroxide complex 3.  相似文献   

9.
It has been claimed that the sole H(2)O(2)-scavenging system in the cyanobacterium Synechococcus sp. PCC 7942 is a cytosolic catalase-peroxidase. We have measured in vivo activity of a light-dependent peroxidase in Synechococcus sp. PCC 7942 and UTEX 625. The addition of small amounts of H(2)O(2) (2.5 microM) to illuminated cells caused photochemical quenching (qP) of chlorophyll fluorescence that was relieved as the H(2)O(2) was consumed. The qP was maximal at about 50 microM H(2)O(2) with a Michaelis constant of about 7 microM. The H(2)O(2)-dependent qP strongly indicates that photoreduction can be involved in H(2)O(2) decomposition. Catalase-peroxidase activity was found to be almost completely inhibited by 10 microM NH(2)OH with no inhibition of the H(2)O(2)-dependent qP, which actually increased, presumably due to the light-dependent reaction now being the only route for H(2)O(2)-decomposition. When (18)O-labeled H(2)O(2) was presented to cells in the light there was an evolution of (16)O(2), indicative of H(2)(16)O oxidation by PS 2 and formation of photoreductant. In the dark (18)O(2) was evolved from added H(2)(18)O(2) as expected for decomposition by the catalase-peroxidase. This evolution was completely blocked by NH(2)OH, whereas the light-dependent evolution of (16)O(2) during H(2)(18)O(2) decomposition was unaffected.  相似文献   

10.
Two gadolinium polyoxometalates, Gd(2)P(2)W(18)O(62) and K(15)[(GdO)(3)(PW(9)O(34))(2)], have been evaluated by in vivo as well as in vitro experiments as the candidates of tissue-specific magnetic resonance imaging (MRI) contrast agents. T(1)-relaxivities of 28.4 mM(-1).s(-1) for Gd(2)P(2)W(18)O(62) and 11.2 mM(-1).s(-1) for K(15)[(GdO)(3)(PW(9)O(34))(2)] (400 MHz, 25 degrees C) were higher than that of the commercial MRI contrast agent (GdDTPA). Their relaxivities in bovine serum albumin and human serum transferrin were also reported. The favorable liver-specific contrast enhancement and renal excretion capability in in vivo MRI with Sprague-Dawley rats after i.v. administration of K(15)[(GdO)(3)(PW(9)O(34))(2)] was demonstrated. In vivo and in vitro assay showed that K(15)[(GdO)(3)(PW(9)O(34))(2)] is a promising liver-specific MRI contrast agent. However, Gd(2)P(2)W(18)O(62) did not show the favorable quality in vivo as expected from its high relaxivity in vitro, which was attributed to low bioavailability, indicating that it is of limited value as tissue-specific MRI contrast agent.  相似文献   

11.
Oxidation of linoleic acid (LA) in tetradecyltrimethylammonium bromide micelles was induced by ferrous- and ferric-chelates in the presence of linoleic acid hydroperoxide (LOOH). Ferrous-chelates also induced lipid peroxidation in the presence of H2O2, but ferric-chelates did not, thought they could generate OH-radicals in the presence of H2O2, resulting in deoxyribose degradation. Of the chelators tested, nitrilotriacetic acid (NTA) chelated with iron showed the highest activity for induction of H2O2- and LOOH-dependent lipid peroxidations and H2O2-dependent deoxyribose degradation. NTA with ferrous ion, but not with ferric ion, also initiated oxidation of LA after a short lag period in the absence of peroxides such as H2O2 and LOOH, but other chelators with ferrous ion did not. The peroxide-independent lipid peroxidation and associated oxidation of ferrous-NTA to ferric-NTA progressed in two steps: an induction step in a lag period and then a propagation step. Ferrous ion complexed with NTA was autoxidized pH-dependently and synchronously with oxygen uptake. The rates of both reactions increased with increase of pH, but were not related to the length of the lag period, which was also dependent on pH, and was shortest at pH 4.2. The EPR spectrum of the ferric-NTA complex prepared directly from ferric salt was different from that of the complex prepared from ferrous salt, confirming that some ferric-type active oxygen participated in induction of peroxide-independent lipid peroxidation. From these results, we propose a possible mechanism of lipid peroxidation induced by ferrous-NTA without peroxides. The finding that iron-NTA had the highest activity for induction of the oxidations of LA and deoxyribose is discussed in relation to the carcinogenic and nephrotoxic effects of this chelating agent.  相似文献   

12.
The excited triplet state of 1-nitronaphthalene ((3)1NN*) reacts with OH(-) with a second-order reaction rate constant of (1.66 ± 0.08)×10(7) M(-1) s(-1) (μ±σ). The reaction yields the ˙OH radical and the radical anion 1NN(-)˙. In aerated solution, the radical 1NN(-)˙ would react with O(2) to finally produce H(2)O(2) upon hydroperoxide/superoxide disproportionation. The photolysis of H(2)O(2) is another potential source of ˙OH, but such a pathway would be a minor one in circumneutral (pH 6.5) or in basic solution ([OH(-)] = 0.3-0.5 M). The oxidation of H(2)O by (3)1NN*, with rate constant 3.8 ± 0.3 M(-1) s(-1), could be the main ˙OH source at pH 6.5.  相似文献   

13.
Lima ES  Di Mascio P  Rubbo H  Abdalla DS 《Biochemistry》2002,41(34):10717-10722
Nitric oxide (*NO) is a pervasive free radical species that concentrates in lipophilic compartments to serve as a potent inhibitor of lipid and low-density lipoprotein oxidation processes. In this study, we synthesized, characterized, and detected nitrated derivatives of linoleic acid (18:2) in human blood plasma using high-pressure liquid chromatography coupled with electrospray ionization tandem mass spectrometry. While the reaction of nitronium tetrafluoroborate with 18:2 presented ions with a mass/charge (m/z) ratio of 324 in the negative ion mode, characteristic of nitrolinoleate (LNO(2)), the reaction of nitrite (NO(2)(-)) with linoleic acid hydroperoxide yielded nitrohydroxylinoleate (LNO(2)OH, m/z 340). Further analysis by MS/MS gave a major fragment at m/z 46, characteristic of a nitro group (-NO(2)) present in the parent ion. This was confirmed by using [(15)N]O(2), which gave products of m/z 325 and 341, that after fragmentation yielded a daughter ion at m/z 47. Moreover, a C-NO(2) structure was also demonstrated in LNO(2)OH by nuclear magnetic resonance spectroscopy ((15)N NMR, delta 375.9), as well as by infrared analysis in both LNO(2)OH (nu(max) = 3427, 1553, and 1374 cm(-1)) and LNO(2) (nu(max) = 1552 and 1373 cm(-1)). Stable products with m/z of 324 and 340, which possessed the same chromatographic characteristics and fragmentation pattern as synthesized standards, were found in human plasma of normolipidemic and hyperlipidemic donors. The presence of these novel nitrogen-containing oxidized lipid adducts in human plasma could represent "footprints" of the antioxidant action of *NO on lipid oxidation and/or a pro-oxidant and nitrating action of *NO-derived species.  相似文献   

14.
16O/18O labeling is one differential proteomics technology among many that promises diagnostic and prognostic biomarkers of disease. Although the incorporation of 18O in the C-terminal carboxyl group during endoproteinase digestion in the presence of H2 18O makes the process of labeling facile, the ease and effectiveness of label incorporation have in some regards been outweighed by the difficulties in interpreting the resulting spectra. Complex isotope patterns result from the composition of unlabeled (18O(0)), singly labeled (18O(1)), and doubly labeled species (18O(2)) as well as contributions from the naturally occurring isotopes (e.g. 13C and 15N). Moreover because labeling is enzymatic, the number of 18O atoms incorporated can vary from peptide to peptide. Finally it is difficult to distinguish highly up-regulated from highly down-regulated or C-terminal peptides. We have developed an algorithm entitled regression analysis applied to mass spectrometry (RAAMS) that automatically, rapidly, and confidently interprets spectra of 18O-labeled peptides without requiring chemical composition information derived from product ion spectra. The algorithm is able to measure the effective 18O incorporation rate due to variable enzyme substrate specificity of the pseudosubstrate during the isotope exchange reaction and corrects for the 18O(0) abundance that remains in the labeled sample when using a two-step digestion/labeling procedure. We have also incorporated a method for distinguishing pure 18O(0) from pure 18O(2) peptides utilizing impure H2 18O. The algorithm operates on centroided peak lists and is therefore very fast: nine chromatograms of, on average, 1,168 spectra and containing, on average, 6,761 isotopic clusters were interpreted in, on average, 45 s per chromatogram. RAAMS is fast enough (average, 38 ms/spectrum) to allow the possibility of performing information-dependent MS/MS on a chromatographic time scale on species exceeding predetermined ratio thresholds. We describe in detail the operation of the algorithm and demonstrate its use on datasets with known and unknown ratios.  相似文献   

15.
Rat liver phenylalanine hydroxylase must be in a reduced form to be catalytically active (Marota, J.J. A., and Shiman, R. (1984) Biochemistry 23, 1303-1311). In this communication we show that a fatty acid hydroperoxide, 13-hydroperoxylinoleic acid (LOOH), can efficiently oxidize the reduced enzyme. In the process, the hydroperoxide is decomposed, oxygen consumed, and hydrogen peroxide formed. Enzyme reduction by the tetrahydropterin cofactor and reoxidation by LOOH can occur as two single steps or, when the enzyme concentration is low compared to that of the substrates, as part of a catalytic cycle. In this latter case, phenylalanine hydroxylase is a hydroperoxide-dependent tetrahydropterin oxidase. The reaction requires 1.0 mol of O2, 1.0 mol of tetrahydropterin, and 0.5 mol of LOOH to yield 1.0 mol of quinonoid dihydropterin, 0.4 mol of H2O2, and fatty acid products. Thus far, the catalytic and single-step reactions appear the same in all properties, consistent with the steady-state reaction following a ping-pong mechanism. Phenylalanine hydroxylase is an excellent catalyst for this reaction: the turnover number with LOOH is slightly greater than with phenylalanine; the Km(app) for LOOH is 11 +/- 4 microM; and the kcat/Km ratio for LOOH is about 25 times greater than for phenylalanine. LOOH and phenylalanine appear to react at different sites on phenylalanine appear to react at different sites on phenylalanine hydroxylase, and the reaction of LOOH is inhibited only slightly by phenylalanine and not at all by 5-deaza-6-methyltetrahydropterin, a competitive inhibitor of phenylalanine hydroxylation. The reaction of LOOH with phenylalanine hydroxylase strongly resembles the nonenzymatic reaction of LOOH with hematin, implying similar mechanisms for the two reactions and implicating the enzyme's non-heme iron as both the site of reaction of LOOH and of electron transfer during oxidation and reduction. The formation of hydrogen peroxide during a reaction of phenylalanine hydroxylase is unusual. Indirect evidence indicates a reduced oxygen species, formed on the enzyme during the reduction step, is (partially) released as H2O2 when the hydroperoxide reacts.  相似文献   

16.
Asparagine (Asn) deamidation and aspartic acid (Asp) isomerization are spontaneous and common alterations occurring in pharmaceutical protein drugs in solution. Because those reactions may cause functional changes, it is important to identify the product-related substances, especially when biopharmaceuticals are under development. In this study, we used H(2)(18)O to identify Asn deamidation and Asp isomerization sites on a recombinant humanized monoclonal antibody (mAb) by using high-performance liquid chromatography-mass spectrometry (HPLC-MS). This strategy takes advantage of reactions whereby (18)O is incorporated into the protein molecule. The mAb was lyophilized and reconstituted in H(2)O or H(2)(18)O, followed by incubation at 50 degrees C for 1 month. Samples were reduced/carboxymethylated and digested by trypsin and then subjected to HPLC-MS and HPLC-tandem mass spectrometry (MS/MS) analysis. Among all of the peptide fragments analyzed, there were two in which deamidation and/or isomerization was observed. In one peptide fragment, an obvious mass shift ( approximately 3Da) at Asn was observed in the newly produced peptide when the mAb was incubated in H(2)(18)O, whereas it was barely feasible to identify this mass shift in H(2)O. In the other peptide fragment, isomerization of Asp was identified after incubation in H(2)(18)O, although it was impossible to distinguish when using H(2)O. By means of this procedure, identification of deamidation and isomerization sites can be accomplished easily even when they are difficult or impossible to detect by the usual peptide mapping.  相似文献   

17.
The aim of the present work was to evaluate the potential for (1)O(2) to induce oxidation of cellular DNA. For this purpose cells were incubated in the presence of a water-soluble endoperoxide whose thermal decomposition leads to the formation of singlet oxygen. Thereafter, DNA was extracted and the level of several modified DNA bases was determined by HPLC analysis coupled to a tandem mass spectrometric detection. A significant increase in the level of 8-oxo-7,8-dihydro-2'-deoxyguanosine was observed upon incubation of the cells with the chemical generator of (1)O(2), whereas the level of the other DNA bases measured remained unchanged. To demonstrate that singlet oxygen is directly involved in the formation of 8-oxo-7, 8-dihydro-2'-deoxyguanosine, the corresponding (18)O-labeled endoperoxide was used. Incubation of the cells with such a generator of (18)O-labeled singlet oxygen results in the formation of (18)O-labeled 8-oxo-7,8-dihydro-2'-deoxyguanosine in the nuclear DNA. This result clearly demonstrates that singlet oxygen, when released within cells, is able to directly oxidize cellular DNA.  相似文献   

18.
Singlet oxygen induces oxidation of cellular DNA   总被引:4,自引:0,他引:4  
The aim of the present work was to evaluate the potential for (1)O(2) to induce oxidation of cellular DNA. For this purpose cells were incubated in the presence of a water-soluble endoperoxide whose thermal decomposition leads to the formation of singlet oxygen. Thereafter, DNA was extracted and the level of several modified DNA bases was determined by HPLC analysis coupled to a tandem mass spectrometric detection. A significant increase in the level of 8-oxo-7,8-dihydro-2'-deoxyguanosine was observed upon incubation of the cells with the chemical generator of (1)O(2), whereas the level of the other DNA bases measured remained unchanged. To demonstrate that singlet oxygen is directly involved in the formation of 8-oxo-7, 8-dihydro-2'-deoxyguanosine, the corresponding (18)O-labeled endoperoxide was used. Incubation of the cells with such a generator of (18)O-labeled singlet oxygen results in the formation of (18)O-labeled 8-oxo-7,8-dihydro-2'-deoxyguanosine in the nuclear DNA. This result clearly demonstrates that singlet oxygen, when released within cells, is able to directly oxidize cellular DNA.  相似文献   

19.
This study was conducted to determine whether provision of preformed dietary docosapentaenoic acid (DPAn-6) can replace docosahexaenoic acid (DHA) for brain function as assessed by spatial task performance. A newly modified artificial rearing method was employed to generate n-3 fatty acid-deficient rats. Newborn pups were separated from their mothers at 2 days of age and given artificial rat milk containing linoleic acid (LA), or LA supplemented with 1% DHA (DHA), 1% DPAn-6 (DPA) or 1% DHA plus 0.4% DPAn-6 (DHA/DPA). The animals were then weaned onto similar pelleted diets. At adulthood, behavioural tasks were administered and then the brains were collected for fatty acid analysis. The LA and DPA groups showed a lower (63-65%) brain DHA than the dam-reared, DHA and DHA/DPA groups and this loss was largely compensated for by an increase in brain DPAn-6. The brain fatty acid composition in the DPA group was the same as that in the LA group at adulthood. In the Morris water maze, the LA and DPA groups exhibited a longer escape latency than the dam-reared and DHA groups and had a defect in spatial retention. In conclusion, DPAn-6 could not replace DHA for brain function, indicating a highly specific structural requirement for DHA.  相似文献   

20.
Nitrobacter agilis, which contains a very active nitrite dehydrogenase, was studied in vivo under anaerobic conditions by the 15N NMR technique. When incubated with equimolar 15NO3- and unlabeled nitrite (or 15NO2- and unlabeled nitrate) the bacterium catalyzed an isotope exchange reaction at rates about 10% those observed in the nitrite oxidase assay. When incubated with 18O-labeled 15NO2- and 18O-labeled 15NO3-, the 18O was observed to exchange at similar rates from both species into water. Finally, when incubated with equimolar [18O]nitrate and 15NO2-, intermolecular 18O transfer was observed to result in formation of double labeled nitrate and nitrite at similar rates. 18O was transferred from nitrate to a 15N species or to water at approximately equal rates under the conditions of the experiments. It is argued that the enzyme responsible for these exchange reactions is nitrite dehydrogenase and not nitrate reductase. This work and the related experiments of DiSpirito and Hooper (DiSpirito, A.A., and Hooper, A.B. (1986) J. Biol. Chem. 261, 10534-10537) represent the first demonstrations of intermolecular oxygen atom transfer among oxotransferases. Mechanistic implications are discussed.  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