首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
Glutathione S-transferases class zeta (GSTζ) is involved in the detoxification of xenobiotic compounds and catalyzes the biotransformation of a variety of α-haloacids including dichloroacetic acid and chlorofluoroacetic acid. It has been reported that, in mice, deficiency of Gstz1 (a member of GSTζ) resulted in the generation of a constant level of oxidative stress. The present study was carried out to investigate the association between genetic polymorphisms of GSTZ1 (in promoter site G-1002A and in coding sites Glu32Lys and Gly42Arg) and risk of breast cancer. We included 106 females with breast cancer and 106 healthy females frequency matched for age. The study polymorphisms were not associated with risk of breast cancer (p>0.05). The polymorphisms of GSTZ1 showed strong linkage disequilibrium among cancer patients and control subjects (p<0.0001). There was no significant difference between cancer patients and controls for frequencies of the GSTZ1 haplotypes (p>0.05). It seems there is no meaningful relationship between the genetic polymorphisms of GSTZ1 and risk of breast cancer.  相似文献   

2.
GABA transporters accumulate GABA to inactivate or reutilize it. Transporter-mediated GABA release can also occur. Recent findings indicate that GABA transporters can perform additional functions. We investigated how activation of GABA transporters can mediate release of glycine. Nerve endings purified from mouse cerebellum were prelabeled with [(3)H]glycine in presence of the glycine GlyT1 transporter inhibitor NFPS to label selectively GlyT2-bearing terminals. GABA was added under superfusion conditions and the mechanisms of the GABA-evoked [(3)H]glycine release were characterized. GABA stimulated [(3)H]glycine release in a concentration-dependent manner (EC(50) = 8.26 μM). The GABA-evoked release was insensitive to GABA(A) and GABA(B) receptor antagonists, but it was abolished by GABA transporter inhibitors. About 25% of the evoked release was dependent on external Ca(2+) entering the nerve terminals through VSCCs sensitive to ω-conotoxins. The external Ca(2+)-independent release involved mitochondrial Ca(2+), as it was prevented by the Na(+)/Ca(2+) exchanger inhibitor CGP37157. The GABA uptake-mediated increases in cytosolic Ca(2+) did not trigger exocytotic release because the [(3)H]glycine efflux was insensitive to clostridial toxins. Bafilomycin inhibited the evoked release likely because it reduced vesicular storage of [(3)H]glycine so that less [(3)H]glycine can become cytosolic when GABA taken up exchanges with [(3)H]glycine at the vesicular inhibitory amino acid transporters shared by the two amino acids. The GABA-evoked [(3)H]glycine efflux could be prevented by niflumic acid or NPPB indicating that the evoked release occurred essentially by permeation through anion channels. In conclusion, GABA uptake into GlyT2-bearing cerebellar nerve endings triggered glycine release which occurred essentially by permeation through Ca(2+)-dependent anion channels. Glial GABA release mediated by anion channels was proposed to underlie tonic inhibition in the cerebellum; the present results suggest that glycine release by neuronal anion channels also might contribute to tonic inhibition.  相似文献   

3.
Myocardial ischaemia is associated with the generation of lipid peroxidation products such as HNE (4-hydroxy-trans-2-nonenal); however, the processes that predispose the ischaemic heart to toxicity by HNE and related species are not well understood. In the present study, we examined HNE metabolism in isolated aerobic and ischaemic rat hearts. In aerobic hearts, the reagent [(3)H]HNE was glutathiolated, oxidized to [(3)H]4-hydroxynonenoic acid, and reduced to [(3)H]1,4-dihydroxynonene. In ischaemic hearts, [(3)H]4-hydroxynonenoic acid formation was inhibited and higher levels of [(3)H]1,4-dihydroxynonene and [(3)H]GS-HNE (glutathione conjugate of HNE) were generated. Metabolism of [(3)H]HNE to [(3)H]4-hydroxynonenoic acid was restored upon reperfusion. Reperfused hearts were more efficient at metabolizing HNE than non-ischaemic hearts. Ischaemia increased the myocardial levels of endogenous HNE and 1,4-dihydroxynonene, but not 4-hydroxynonenoic acid. Isolated cardiac mitochondria metabolized [(3)H]HNE primarily to [(3)H]4-hydroxynonenoic acid and minimally to [(3)H]1,4-dihydroxynonene and [(3)H]GS-HNE. Moreover, [(3)H]4-hydroxynonenoic acid was extruded from mitochondria, whereas other [(3)H]HNE metabolites were retained in the matrix. Mitochondria isolated from ischaemic hearts were found to contain 2-fold higher levels of protein-bound HNE than the cytosol, as well as increased [(3)H]GS-HNE and [(3)H]1,4-dihydroxynonene, but not [(3)H]4-hydroxynonenoic acid. Mitochondrial HNE oxidation was inhibited at an NAD(+)/NADH ratio of 0.4 (equivalent to the ischaemic heart) and restored at an NAD(+)/NADH ratio of 8.6 (equivalent to the reperfused heart). These results suggest that HNE metabolism is inhibited during myocardial ischaemia owing to NAD(+) depletion. This decrease in mitochondrial metabolism of lipid peroxidation products and the inability of the mitochondria to extrude HNE metabolites could contribute to myocardial ischaemia/reperfusion injury.  相似文献   

4.
拟南芥谷胱甘肽S-转移酶Zeta类(AtGSTZ)是一种与细胞代谢和环境净化密切相关的多功能酶.应用易错PCR和多轮DNA洗牌技术构建了AtGSTZ随机突变文库;再利用pH指示剂颜色改变法对突变文库进行筛选,获得了9个二氯乙酸脱氯活性提高的突变子.其中,NN23含25个氨基酸突变,比活力提高120%,NN20含24个氨基酸突变,比活力提高102%,EC1含2个氨基酸突变,比活力提高47%,其他6个为单点突变,比活力分别提高9%~60%.酶学分析显示,所有进化酶对底物二氯乙酸的催化效率和对谷胱甘肽的亲和力以及个别进化酶的复性能力都得到不同程度的提高,但热稳定性均没有明显改善.同时,对一系列与AtGSTZ空间折叠及催化活性相关位点进行了讨论.  相似文献   

5.
1. A single oral dose of either [(14)C]Prefix or 2,6-dichlorobenzo[(14)C]nitrile to rats is almost entirely eliminated in 4 days: 84.8-100.5% of (14)C from [(14)C]Prefix is excreted, 67.3-79.7% in the urine, and 85.8-97.2% of (14)C from 2,6-dichlorobenzo-[(14)C]nitrile is excreted, 72.3-80.7% in the urine. Only 0.37+/-0.03% of the dose of [(14)C]Prefix and 0.25+/-0.03% of the dose of 2,6-dichlorobenzo[(14)C]nitrile are present in the carcass plus viscera after removal of the gut. Rats do not show sex differences in the pattern of elimination of the respective metabolites of the two herbicides. The rates of elimination of (14)C from the two compounds in the 24hr. and 48hr. urines are not significantly different (P >0.05) from one another. 2. After oral administration to dogs, 85.9-106.1% of (14)C from [(14)C]Prefix is excreted, 66.6-80.9% in the urine, and 86.8-92.5% of (14)C from 2,6-dichlorobenzo[(14)C]nitrile is excreted, 60.0-70.1% in the urine. Dogs do not show sex differences in the pattern of eliminating the metabolites of either Prefix or 2,6-dichlorobenzonitrile. 3. Dogs and rats do not show species differences in the patterns of elimination of the two herbicides. 4. Prefix and 2,6-dichlorobenzonitrile are completely metabolized; unchanged Prefix and 2,6-dichlorobenzonitrile are absent from the urine and faeces, and from the carcasses when elimination is complete. In the hydrolysed urine of rats dosed with either [(14)C]Prefix or 2,6-dichlorobenzo[(14)C]nitrile, 2,6-dichloro-3-hydroxybenzonitrile accounts for approx. 42% of the (14)C, a further 10-11% is accounted for by 2,6-dichlorobenzamide, 2,6-dichlorobenzoic acid, 2,6-dichloro-3- and -4-hydroxybenzoic acid and 2,6-dichloro-4-hydroxybenzonitrile collectively, and 25-30% by six polar constituents, of which two are sulphur-containing amino acids. 5. In the unhydrolysed urines of rats dosed with either [(14)C]Prefix or 2,6-dichlorobenzo[(14)C]nitrile, there are present free 2,6-dichloro-3- and -4-hydroxybenzonitrile, their glucuronide conjugates, ester glucuronides of the principal aromatic acids that are present in the hydrolysed urines, and two sulphur-containing metabolites analogous to mercapturic acids or premercapturic acids. 6. Prefix is thus extensively transformed into 2,6-dichlorobenzonitrile: R.CS.NH(2)-->R.CN+H(2)S, where R=C(6)H(3)Cl(2). However, the competitive reaction: R.CS.NH(2)+H(2)O-->R.CO.NH(2)+H(2)S takes place to a very limited extent.  相似文献   

6.
谷胱甘肽S-转移酶Zeta类基因在酿酒酵母中的表达   总被引:1,自引:0,他引:1  
贾向东  陈喜文  陈德富  陈洁 《遗传》2006,28(5):551-556
谷胱甘肽S-转移酶Zeta类基因在酿酒酵母中的表达 贾向东1,陈喜文1,陈德富1,陈洁2 (1.南开大学生命科学学院,生物活性材料教育部重点实验室,天津300071;2.湖南怀化市铁路第一中学,怀化418000) 摘要:谷胱甘肽S-转移酶Zeta类(GSTZ)是一种重要的多功能酶,与细胞生化代谢、环境净化等密切相关。将拟南芥、甘蓝型油菜品系陕2B与垦C1的GSTZ基因克隆到大肠杆菌—酿酒酵母穿梭表达载体pYES2的多克隆位点,筛选到重组子后,提取重组质粒并将其转入酿酒酵母营养缺陷型菌株INCSc1细胞中,经SC-U培养基选择得到重组酵母Y2At、Y2BnB和Y2BnC。重组酵母在含棉子糖和半乳糖的诱导培养基中,表达出了具有二氯乙酸脱氯活力的谷胱甘肽S-转移酶Zeta类,且主要以可溶状态存在于酵母细胞中。不同碳源比较发现,使用半乳糖为唯一碳源时,与棉子糖和半乳糖共同使用相比,酵母生长虽受到轻微影响,但表达的GSTZ比活力几乎不受任何影响。0~96h诱导时间的优化实验表明,36h诱导下呈现最高比活力。同时也对不同GSTZ的Km值进行了比较。  相似文献   

7.
The effect of alpha-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid (AMPA), a selective glutamate receptor agonist, on the release of previously incorporated [(3)H]GABA was examined in superfused striatal slices of the rat. The slices were loaded with [(3)H]GABA in the presence of beta-alanine (1 mM) and superfused with Krebs-bicarbonate buffer containing nipecotic acid (0.1 mM) and aminooxyacetic acid (0.1 mM) to inhibit GABA uptake and metabolism. AMPA (0.01 to 3 mM) increased basal [(3)H]GABA outflow and nipecotic acid potentiated this effect. The [(3)H]GABA releasing effect of AMPA was an external Ca(2+)-dependent process in the absence but not in the presence of nipecotic acid. Cyclothiazide (0.03 mM), a positive modulator of AMPA receptors, failed to evoke [(3)H]GABA release by itself, but it dose-dependently potentiated the [(3)H]GABA releasing effect of AMPA. The AMPA (0.3 mM)-induced [(3)H]GABA release was antagonized by NBQX (0.01 mM) in a competitive fashion (pA(2) 5.08). The negative modulator of AMPA receptors, GYKI-53784 (0.01 mM) reversed the AMPA-induced [(3)H]GABA release by a non-competitive manner (pD'(2) 5.44). GYKI-53784 (0. 01-0.1 mM) also decreased striatal [(3)H]GABA outflow on its own right, this effect was stereoselective and was not influenced by concomitant administration of 0.03 mM cyclothiazide. GYKI-52466 (0. 03-0.3 mM), another negative modulator at AMPA receptors, also inhibited basal [(3)H]GABA efflux whereas NBQX (0.1 mM) by itself was ineffective in alteration of [(3)H]GABA outflow.The present data indicate that AMPA evokes GABA release from the vesicular pool in neostriatal GABAergic neurons. They also confirm that multiple interactions may exist between the agonist binding sites and the positive and negative modulatory sites but no such interaction was detected between the positive and negative allosteric modulators. Since GYKI-53784, but not NBQX, inhibited [(3)H]GABA release by itself, AMPA receptors located on striatal GABAergic neurons may be in sensitized state and phasically controlled by endogenous glutamate. It is also postulated that these AMPA receptors are located extrasynaptically on GABAergic striatal neurons.  相似文献   

8.
We have investigated the role of endothelial cells in the metabolism of 20-hydroxyeicosatetraenoic acid (20-HETE), a vasoactive mediator synthesized from arachidonic acid by cytochrome P450 omega-oxidases. Porcine coronary artery endothelial cells (PCEC) incorporated 20-[(3)H]HETE primarily into the sn-2 position of phospholipids through a coenzyme A-dependent process. The incorporation was reduced by equimolar amounts of arachidonic, eicosapentaenoic or 8,9-epoxyeicosatrienoic acids, but some uptake persisted even when a 10-fold excess of arachidonic acid was available. The retention of 20-[(3)H]HETE increased substantially when methyl arachidonoyl fluorophosphonate, but not bromoenol lactone, was added, suggesting that a Ca(2+)-dependent cytosolic phospholipase A(2) released the 20-HETE contained in PCEC phospholipids. Addition of calcium ionophore A23187 produced a rapid release of 20-[(3)H]HETE from the PCEC, a finding that also is consistent with a Ca(2+)-dependent mobilization process. PCEC also converted 20-[(3)H]HETE to 20-carboxy-arachidonic acid (20-COOH-AA) and 18-, 16-, and 14-carbon beta-oxidation products. 20-COOH-AA produced vasodilation in porcine coronary arterioles, but 20-HETE was inactive. These results suggest that the incorporation of 20-HETE and its subsequent conversion to 20-COOH-AA in the endothelium may be important in modulating coronary vascular function.  相似文献   

9.
Defects in Niemann-Pick, Type C-1 protein (NPC1) cause cholesterol, sphingolipids, phospholipids, and glycolipids to accumulate in lysosomes of liver, spleen, and brain. In cultured fibroblasts, NPC1 deficiency causes lysosomal retention of lipoprotein-derived cholesterol after uptake by receptor-mediated endocytosis. NPC1 contains 1278 amino acids that form 13 membrane-spanning helices and three large loops that project into the lumen of lysosomes. We showed earlier that NPC1 binds cholesterol and oxysterols. Here we localize the binding site to luminal loop-1, a 240-amino acid domain with 18 cysteines. When produced in cultured cells, luminal loop-1 was secreted as a soluble dimer. This loop bound [(3)H]cholesterol (K(d), 130 nM) and [(3)H]25-hydroxycholesterol (25-HC, K(d), 10 nM) with one sterol binding site per dimer. Binding of both sterols was competed by oxysterols (24-, 25-, and 27-HC). Unlabeled cholesterol competed strongly for binding of [(3)H]cholesterol, but weakly for [(3)H]25-HC binding. Binding of [(3)H]cholesterol but not [(3)H]25-HC was inhibited by detergents. We also studied NPC2, a soluble protein whose deficiency causes a similar disease phenotype. NPC2 bound cholesterol, but not oxysterols. Epicholesterol and cholesteryl sulfate competed for [(3)H]cholesterol binding to NPC2, but not NPC1. Glutamine 79 in luminal loop-1 of NPC-1 is important for sterol binding; a Q79A mutation abolished binding of [(3)H]cholesterol and [(3)H]25-HC to full-length NPC1. Nevertheless, the Q79A mutant restored cholesterol transport to NPC1-deficient Chinese hamster ovary cells. Thus, the sterol binding site on luminal loop-1 is not essential for NPC1 function in fibroblasts, but it may function in other cells where NPC1 deficiency produces more complicated lipid abnormalities.  相似文献   

10.
The effects of saturated long-chain (C: 16-22) N-acylethanolamines and a series of saturated fatty acids with the same length of carbon chains were investigated on depolarization-induced (45)Ca(2+) fluxes mediated by voltage-dependent Ca(2+) channels in transverse tubule membrane vesicles from rabbit skeletal muscle. Vesicles were loaded with (45)Ca(2+) and membrane potentials were generated by establishing potassium gradients across the vesicle using the ionophore valinomycin. Arachidonoylethanolamide and docosaenoylethanolamide but not palmitoylethanolamide and stearoylethanolamide (all 10 microM) caused a significant inhibition of depolarization-induced (45)Ca(2+) fluxes and specific binding of [(3)H]Isradipine to transverse tubule membranes. On the other hand, saturated fatty acids including palmitic, stearic, arachidic, and docosanoic acids (all 10 microM) were ineffective in functional and radioligand binding experiments. Additional experiments using endocannabinoid metabolites suggested that whereas ethanolamine and arachidic acids were ineffective, arachidonoylethanolamide inhibited Ca(2+) effluxes and specific binding of [(3)H]Isradipine. Further studies indicated that only those fatty acids containing ethanolamine as a head group and having a chain length of more than 18 carbons were effective in inhibiting depolarization-induced Ca(2+) effluxes and specific binding of [(3)H]Isradipine. In conclusion, results indicate that depending on the chain length and the head group of fatty acid, N-acylethanolamines have differential effects on the function of voltage-dependent Ca(2+) channels and on the specific binding of [(3)H]Isradipine in skeletal muscle membranes.  相似文献   

11.
1. Fatty acid synthesis was studied in microsomal preparations from germinating pea (Pisum sativum). 2. The preparations synthesized a mixture of saturated fatty acids up to a chain length of C(24) from [(14)C]malonyl-CoA. 3. Whereas hexadecanoic acid was made de novo, octadecanoic acid and icosanoic acid were synthesized by elongation. 4. The products formed during [(14)C]malonyl-CoA incubation were analysed, and unesterified fatty acids and polar lipids were found to be major products. [(14)C]Palmitic acid represented a high percentage of the acyl-carrier protein esters, whereas (14)C-labelled very-long-chain fatty acids were mainly present as unesterified fatty acids. CoA esters were minor products. 5. The addition of exogenous lipids to the incubation system usually resulted in stimulation of [(14)C]malonyl-CoA incorporation into fatty acids. The greatest stimulation was obtained with dipalmitoyl phosphatidylcholine. Both exogenous palmitic acid and dipalmitoyl phosphatidylcholine increased the amount of [(14)C]-stearic acid synthesized, relative to [(14)C]palmitic acid. Addition of stearic acid increased the amount of [(14)C]icosanoic acid formed. 6. [(14)C]Stearic acid was elongated more effectively to icosanoic acid than [(14)C]stearoyl-CoA, and its conversion was not decreased by addition of unlabelled stearoyl-CoA. 7. Incorporation of [(14)C]malonyl-CoA into fatty acids was markedly decreased by iodoacetamide and 5,5'-dithiobis-(2-nitrobenzoic acid). Palmitate elongation was sensitive to arsenite addition, and stearate elongation to the presence of Triton X-100 or fluoride. The action of fluoride was not, apparently, due to chelation. 8. The microsomal preparations differed from soluble fractions from germinating pea in (a) synthesizing very-long-chain fatty acids, (b) not utilizing exogenous palmitate-acyl-carrier protein as a substrate for palmitate elongation and (c) having fatty acid synthesis stimulated by the addition of certain complex lipids.  相似文献   

12.
The whole-cell phospholipid composition of the six known polymorphic species of Tetrahymena has been examined by [(3)H]acetate and [(3)H]myristic acid radiolabeling, and by gas-liquid chromatography of total phospholipid-bound fatty acids. Five of the polymorphic species contained similar phospholipid profiles following radiolabeling in that phosphatidylethanolamine (PE) was the predominant phospholipid; however, in cells of Tetrahymena patula LFF, aminoethylphosphonolipid was present in amounts nearly equal to PE. Tetrahymena patula LFF contained an unusually large percentage of sphingolipid (16.2% by [(3)H]acetate radiolabeling). Substantial differences were found in the fatty acid profiles of the polymorphic species, which included the degree of fatty acid unsaturation and relative weight percentages of odd-chain fatty acids. Tetrahymena vorax contained a low ratio of unsaturated C(18) fatty acids to saturated C(18) fatty acids as compared with all other species examined. The differentiating species generally contained a lesser percentage of monoenoic fatty acids and a lower ratio of unsaturated C(16) fatty acids to saturated C(16) fatty acids as compared with the two monomorphic species examined.  相似文献   

13.
Oxalic acid metabolism and calcium oxalate formation in Lemna minor L.   总被引:6,自引:0,他引:6  
Abstract Axenic Lemna minor plants, which form numerous calcium oxalate crystals, were exposed to [14C]-glycolic acid, -glyoxylic acid, -oxalic acid and -ascorbic acid and prepared for microautoradiography by a technique that preserves only insoluble label to determine specifically the pathway leading to oxalic acid used for crystal formation. Label from glycolic, glyoxylic, and oxalic acids was incorporated into crystals. Label from oxalic acid was also found in starch when exposure to label was done in the light but not dark, while plastids specialized for lipid storage were heavily labelled under both conditions. Incorporation of label from glycolic and glyoxylic acids, but not oxalic acid, was inhibited in the presence of the glycolate oxidase inhibitors, αHPMS (2-pyridylhydroxy methanesulphonic acid) and mHBA (methyl 2-hydroxy-3-butynoic acid), and inhibition of labelling was not due to an effect on uptake. These studies show that the glycolate oxidase pathway to oxalic acid is operational in L. minor and that the product is available for crystal formation. Dark-grown plants form almost four times as many crystal cells (idioblasts) as do light-grown plants, indicating crystal formation is not in response to photorespiratory glycolate production. Label from [1-14C]ascorbic acid was also incorporated into crystals and labelling was inhibited by mHBA, indicating glycolic acid and/or glyoxylic acid are possible intermediates of ascorbic acid catabolism. The effect of nitrogen source on crystal formation was also investigated. Significantly more crystal idioblasts were formed, on a surface area basis, by plants grown on ammonium than by plants grown on nitrate nitrogen. When grown with mixed ammonium and nitrate, an intermediate number of crystal idioblasts were formed.  相似文献   

14.
Previous studies examined the bioavailability and first-pass biotransformation of 3-hydroxy[(3)H]benzo[a]pyrene ([(3)H]-3-OHBaP) in an isolated perfused catfish intestinal model. This work showed that 3-OHBaP, or a metabolite formed in intestine, bound covalently to blood protein. In this study, the blood adducts were characterized in vitro by incubating bovine ferric hemoglobin or albumin with [(3)H]-3OHBaP under various conditions. Incubation of 2 microM [(3)H]-3-OHBaP with hemoglobin for 1 h resulted in 7.49 pmol bound/mg protein, while albumin binding was 1.37 pmol/mg protein. Mild acid hydrolysis released only 5% of the radioactivity from 3-OHBaP-hemoglobin adducts. After gel filtration, the 3-OHBaP-hemoglobin adducts were examined by HPLC analysis. A single peak of radioactivity was detected at the same retention time as the heme component of hemoglobin. Unbound 3-OHBaP was oxidized to BaP-3,6-dione during incubation with ferric hemoglobin. Treatment of hemoglobin with ascorbic acid decreased the formation of hemoglobin adducts by 33%, while hydrogen peroxide treatment increased adduct formation by 44%. Incubation of [(3)H]-BaP-3-beta-D-glucuronide (BaP-3G) with hemoglobin and beta-glucuronidase resulted in greater binding to hemoglobin than incubation with [(3)H]-3-OHBaP alone. The hemoglobin adduct obtained from [(3)H]-BaP-3G also co-migrated with heme. These results indicate that an oxidative process is involved in formation of the heme adduct and that 3-OHBaP or BaP-3G might be a precursor of the bound metabolite.  相似文献   

15.
1. Chloroplasts isolated from spinach leaves by using the low-ionic-strength buffers of Nakatani & Barber [(1977) Biochim. Biophys. Acta.461, 510-512] had higher rates of HCO(3) (-)-dependent oxygen evolution (up to 369mumol/h per mg of chlorophyll) and higher rates of [1-(14)C]acetate incorporation into long-chain fatty acids (up to 1500nmol/h per mg of chlorophyll) than chloroplasts isolated by using alternative procedures. 2. Acetate appeared to be the preferred substrate for fatty acid synthesis by isolated chloroplasts, although high rates of synthesis were also measured from H(14)CO(3) (-) in assays permitting high rats of photosynthesis. Incorporation of H(14)CO(3) (-) into fatty acids was decreased by relatively low concentrations of unlabelled acetate. Acetyl-CoA synthetase activity was present 3-4 times in excess of that required to account for rates of [1-(14)C]acetate incorporation into fatty acids, but pyruvate dehydrogenase was either absent or present in very low activity in spinach chloroplasts. 3. Rates of long-chain-fatty acid synthesis from [1-(14)C]acetate in the highly active chloroplast preparations, compared with those used previously, were less dependent on added cofactors, but showed a greater response to light. The effects of added CoA plus ATP, Triton X-100 and sn-glycerol 3-phosphate on the products of [1-(14)C]acetate incorporation were similar to those reported for less active chloroplast preparations. 4. Endogenous [(14)C]acetyl-CoA plus [(14)C]malonyl-CoA was maintained at a constant low level even when fatty acid synthesis was limited by low HCO(3) (-) concentrations. Endogenous [(14)C]acyl-(acyl-carrier protein) concentrations increased with increasing HCO(3) (-) concentration and higher rates of fatty acid synthesis, but were slightly lower in the presence of Triton X-100. It is proposed that rates of long-chain-fatty acid synthesis in isolated chloroplasts at saturating [1-(14)C]acetate concentrations and optimal HCO(3) (-) concentrations may be primarily controlled by rates of removal of the products of the fatty acid synthetase.  相似文献   

16.
The biosynthetic pathway of an unusual amino acyl [(2S,3R)-3-amino-2-hydroxy-4-phenylbutanoyl (AHP)] moiety which is contained in bestatin has been studied by testing the incorporation of potential precursors. L-[U-14C]-Phenylalanine, L-[U-14C]leucine, and [U-14C]acetic acid were efficiently incorporated into bestatin, but the radioactivity of L-[1-14C]phenylalanine, [1-14C]glyoxylic acid, and [14C]oxalic acid were not incorporated. Incorporation of acetic acid into 1- and 2-carbon of the AHP moiety was confirmed by incorporation of [13C]acetic acid. Thus, the AHP moiety was shown to be biosynthesized from L-phenylalanine and two carbon atoms of acetic acid, accompanied by decarboxylation of the phenylalanine.  相似文献   

17.
d-Amino acids are now recognized to be widely present in mammals. Renal d-amino-acid oxidase (DAO) is associated with conversion of d-amino acids to the corresponding alpha-keto acids, but its contribution in vivo is poorly understood because the alpha-keto acids and/or l-amino acids formed are indistinguishable from endogenous compounds. First, we examined whether DAO is indispensable for conversion of d-amino acids to their alpha-keto acids by using the stable isotope tracer technique. After a bolus intravenous administration of d-[(2)H(7)]leucine to mutant mice lacking DAO activity (ddY/DAO(-)) and normal mice (ddY/DAO(+)), elimination of d-[(2)H(7)]leucine and formation of alpha-[(2)H(7)]ketoisocaproic acid ([(2)H(7)]KIC) and l-[(2)H(7)]leucine in plasma were determined. The ddY/DAO(-) mice, in contrast to ddY/DAO(+) mice, failed to convert d-[(2)H(7)]leucine to [(2)H(7)]KIC and l-[(2)H(7)]leucine. This result clearly revealed that DAO was indispensable for the process of chiral inversion of d-leucine. We further investigated the effect of renal mass reduction by partial nephrectomy on elimination of d-[(2)H(7)]leucine and formation of [(2)H(7)]KIC and l-[(2)H(7)]leucine. Renal mass reduction slowed down the elimination of d-[(2)H(7)]leucine. The fraction of conversion of d-[(2)H(7)]leucine to [(2)H(7)]KIC in sham-operated rats was 0.77, whereas that in five-sixths-nephrectomized rats was 0.25. The elimination behavior of d-[(2)H(7)]leucine observed in rats suggested that kidney was the principal organ responsible for converting d-leucine to KIC.  相似文献   

18.
Neutral cholesteryl ester hydrolase (CEH)-mediated hydrolysis of cellular cholesteryl esters (CEs) is required not only to generate free cholesterol (FC) for efflux from macrophages but also to release FC from lipoprotein-delivered CE in the liver for bile acid synthesis or direct secretion into the bile. We hypothesized that hepatic expression of CEH would regulate the hydrolysis of lipoprotein-derived CE and enhance reverse cholesterol transport (RCT). Adenoviral-mediated CEH overexpression led to a significant increase in bile acid output. To assess the role of hepatic CEH in promoting flux of cholesterol from macrophages to feces, cholesterol-loaded and [(3)H]cholesterol-labeled J774 macrophages were injected intraperitoneally into mice and the appearance of [(3)H]cholesterol in gallbladder bile and feces over 48 h was quantified. Mice overexpressing CEH had significantly higher [(3)H]cholesterol radiolabel in bile and feces, and it was associated with bile acids. This CEH-mediated increased movement of [(3)H]cholesterol from macrophages to bile acids and feces was significantly attenuated in SR-BI(-/-) mice. These studies demonstrate that similar to macrophage CEH that rate-limits the first step, hepatic CEH regulates the last step of RCT by promoting the flux of cholesterol entering the liver via SR-BI and increasing hepatic bile acid output.  相似文献   

19.
The site of inhibition of peptidoglycan synthesis during the stringent response in Escherichia coli was determined in strains which were auxotrophic for both lysine and diaminopimelic acid (DAP). Cells were labeled with [(3)H]DAP for 30 to 60 min in the presence and absence of required amino acids, and the cellular distribution of [(3)H]DAP was determined. In both stringent (rel(+)) and relaxed (relA) strains, amino acid deprivation did not inhibit the incorporation of [(3)H]DAP into the nucleotide precursor and lipid intermediate fractions. The amount of [(3)H]DAP incorporated into the peptidoglycan fraction by the amino acid-deprived relA strain was over 70% of the amount incorporated in the presence of required amino acids. In contrast, the amounts of labeled peptidoglycan in amino acid-deprived rel(+) strains were only 20 to 44% of the amounts synthesized in the presence of amino acids. These results indicate that a late step in peptidoglycan synthesis is inhibited during the stringent response. The components of the lipid intermediate fraction synthesized by rel(+) strains in the presence and absence of required amino acids were quantitated. Amino acid deprivation did not inhibit the synthesis of either the monosaccharide-pentapeptide or the disaccharide-pentapeptide derivatives of the lipid intermediate. Thus, the reaction which is most likely inhibited during the stringent response is the terminal one involving the incorporation of the disaccharide-pentapeptide into peptidoglycan.  相似文献   

20.
We investigated the role of caveolae in uptake and intracellular trafficking of long chain fatty acids (LCFA) in HepG2 human hepatoma cells. The uptake of [(3)H]oleic acid and [(3)H]stearic acid into HepG2 cells was measured by radioactive assays and internalization of the non-metabolizable fluorescent fatty acid 12-(N-methyl)-N-[(7-nitrobenz-2-oxa-1,3-diazol-4-yl)amino] (12-NBD) stearate into single HepG2 cells was semi-quantitatively assessed by laser scanning microscopy. The initial rate of [(3)H]oleic acid uptake (V(0)) in HepG2 cells exhibited saturable transport kinetics with increasing concentrations of free oleic acid (V(max) 854 +/- 46 pmol mg protein(-1) min(-1), K(m) 100 +/- 14 nmol/l). While inhibition of clathrin coated pits did not influence LCFA uptake in HepG2, inhibition of caveolae formation by filipin III, cyclodextrin, and caveolin-1 antisense oligonucleotides resulted in reduction of [(3)H]oleic acid uptake by 54%, 45%, and 23%, respectively. Furthermore, filipin III inhibited the uptake of [(3)H]stearic acid and its fluorescent derivative 12-NBD stearate by 44% and 50%, respectively. Transfection studies with alpha-caveolin-1/cyanofluorescent protein chimeras showed significant colocalization of caveolae and internalized 12-NBD stearate. In conclusion, these data suggest a significant role for caveolae mediated uptake and intracellular trafficking of LCFA in HepG2 cells.  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