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.
D-Carnitine was converted to L-carnitine by resting and permeabilized cells as well as with purified stereospecific carnitine dehydrogenases from Agrobacterium sp. With permeabilized cells only 11% of D-carnitine was converted into L-carnitine. Using highly stereospecific D- and L-carnitine dehydrogenases from Agrobacterium sp. (pH 8.5, 50 mM D-carnitine, 1 mM NAD + , 0.1 mM NADH, 25-fold excess of L-carnitine dehydrogenase) almost 50% of the D-carnitine could be converted into L-carnitine.  相似文献   

2.
The aim was to understand how interaction of the central carbon and the secondary carnitine metabolisms is affected under salt stress and its effect on the production of L-carnitine by Escherichia coli. The biotransformation of crotonobetaine into L-carnitine by resting cells of E. coli O44 K74 was improved by salt stress, a yield of nearly twofold that for the control being obtained with 0.5 M NaCl. Crotonobetaine and the L-carnitine formed acted as an osmoprotectant during cell growth and biotransformation in the presence of NaCl. The enzyme activities involved in the biotransformation process (crotonobetaine hydration reaction and crotonobetaine reduction reaction), in the synthesis of acetyl-CoA/acetate (pyruvate dehydrogenase, acetyl-CoA synthetase [ACS] and ATP/acetate phosphotransferase) and in the distribution of metabolites for the tricarboxylic acid cycle (isocitrate dehydrogenase [ICDH]) and glyoxylate shunt (isocitrate lyase [ICL]) were followed in batch with resting cells both in the presence and absence of NaCl and in perturbation experiments performed on growing cells in a high density cell recycle membrane reactor. Further, the levels of carnitine, crotonobetaine, gamma-butyrobetaine and ATP and the NADH/NAD(+) ratio were measured in order to know how the metabolic state was modified and coenzyme pools redistributed as a result of NaCl's effect on the energy content of the cell. The results provided the first experimental evidence of the important role played by salt stress during resting and growing cell biotransformation (0.5 M NaCl increased the L-carnitine production in nearly 85%), and the need for high levels of ATP to maintain metabolite transport and biotransformation. Moreover, the main metabolic pathways and carbon flow operating during cell biotransformation was that controlled by the ICDH/ICL ratio, which decreased from 8.0 to 2.5, and the phosphotransferase/ACS ratio, which increased from 2.1 to 5.2, after a NaCl pulse fivefold the steady-state level. Resting E. coli cells were seen to be made up of heterogeneous populations consisting of several types of subpopulation (intact, depolarized, and permeabilized cells) differing in viability and metabolic activity as biotransformation run-time and the NaCl concentration increased. The results are discussed in relation with the general stress response of E. coli, which alters the NADH/NAD(+) ratio, ATP content, and central carbon enzyme activities.  相似文献   

3.
The oxidation of formaldehyde by rat liver mitochondria in the presence of 50 mM phosphate was enhanced 2-fold by exogenous NAD+. Absolute requirement of NAD+ for formaldehyde oxidation was demonstrated by depleting the mitochondria of their NAD+ content (4.6 nmol/mg of protein), followed by reincorporation of the NAD+ into the depleted mitochondria. Aldehyde (formaldehyde) dehydrogenase activity was completely abolished in the depleted mitochondria, but the enzyme activity was restored to control levels following reincorporation of the pyridine nucleotide. Phosphate stimulation of formaldehyde oxidation could not be explained fully by the phosphate-induced swelling which enhances membrane permeability to NAD+, since stimulation of the enzyme activity by increased phosphate concentrations was still observed in the absence of exogenous NAD+. The Km for formaldehyde oxidation by the mitochondria was found to be 0.38 nM, a value similar to that obtained with varying concentrations of NAD+; both Vmax values were very similar, giving a value of 70 to 80 nmol/min/mg of protein. The pH optimum for the mitochondrial enzyme was 8.0. Inhibition of the enzyme activity by anaerobiosis was apparently due to the inability of the respiratory chain to oxidize the generated NADH. The inhibition of mitochondrial formaldehyde oxidation by succinate was found to be due to a lowering of the NAD+ level in the mitochondria. Succinate also inhibited acetaldehyde oxidation by the mitochondria. Malonate, a competitive inhibitor of succinic dehydrogenase, blocked the inhibitory effect of succinate. The respiratory chain inhibitors, rotenone, and antimycin A plus succinate, strongly inhibited formaldehyde oxidation by apparently the same mechanism, although the crude enzyme preparation (freed from the membrane) was slightly sensitive to rotenone. The mitochondria were subfractionated, and 85% of the enzyme activity was found in the inner membrane fraction (mitoplast). Furthermore, separation into inner membrane and matrix components indicated a distribution of aldehyde dehydrogenase activity similar to malic dehydrogenase.  相似文献   

4.
To identify the enzyme responsible for pentitol oxidation by acetic acid bacteria, two different ribitol oxidizing enzymes, one in the cytosolic fraction of NAD(P)-dependent and the other in the membrane fraction of NAD(P)-independent enzymes, were examined with respect to oxidative fermentation. The cytoplasmic NAD-dependent ribitol dehydrogenase (EC 1.1.1.56) was crystallized from Gluconobacter suboxydans IFO 12528 and found to be an enzyme having 100 kDa of molecular mass and 5 s as the sedimentation constant, composed of four identical subunits of 25 kDa. The enzyme catalyzed a shuttle reversible oxidoreduction between ribitol and D-ribulose in the presence of NAD and NADH, respectively. Xylitol and L-arabitol were well oxidized by the enzyme with reaction rates comparable to ribitol oxidation. D-Ribulose, L-ribulose, and L-xylulose were well reduced by the enzyme in the presence of NADH as cosubstrates. The optimum pH of pentitol oxidation was found at alkaline pH such as 9.5-10.5 and ketopentose reduction was found at pH 6.0. NAD-Dependent ribitol dehydrogenase seemed to be specific to oxidoreduction between pentitols and ketopentoses and D-sorbitol and D-mannitol were not oxidized by this enzyme. However, no D-ribulose accumulation was observed outside the cells during the growth of the organism on ribitol. L-Ribulose was accumulated in the culture medium instead, as the direct oxidation product catalyzed by a membrane-bound NAD(P)-independent ribitol dehydrogenase. Thus, the physiological role of NAD-dependent ribitol dehydrogenase was accounted to catalyze ribitol oxidation to D-ribulose in cytoplasm, taking D-ribulose to the pentose phosphate pathway after being phosphorylated. L-Ribulose outside the cells would be incorporated into the cytoplasm in several ways when need for carbon and energy sources made it necessary to use L-ribulose for their survival. From a series of simple experiments, membrane-bound sugar alcohol dehydrogenase was concluded to be the enzyme responsible for L-ribulose production in oxidative fermentation by acetic acid bacteria.  相似文献   

5.
The kinetics of purified beef heart mitochondrial carnitine palmitoyltransferase have been extensively investigated with a semiautomated system and the computer program TANKIN and shown to be sigmoidal with both acyl-CoA and L-carnitine. In contrast, Michaelis-Menten kinetics were found with carnitine octanoyltransferase. The catalytic activity of carnitine palmitoyltransferase is strongly pH dependent. The K0.5 and Vmax are both greater at lower pH. The K0.5 for palmitoyl-CoA is 1.9 and 24.2 microM at pH 8 and 6, respectively. The K0.5 for L-carnitine is 0.2 and 2.9 mM at pH 8 and 6, respectively. Malonyl-CoA (20-600 microM) had no effect on the kinetic parameters for palmitoyl-CoA at both saturating and subsaturating levels of L-carnitine. We conclude that malonyl-CoA is not a competitive inhibitor of carnitine palmitoyltransferase. The purified enzyme contained 18.9 mol of bound phospholipid/mol of enzyme which were identified as cardiolipin, phosphatidylethanolamine, and phosphatidylcholine by thin-layer chromatography. The data are consistent with the conclusion that native carnitine palmitoyltransferase exhibits different catalytic properties on either side of the inner membrane of mitochondria due to its non-Michaelis-Menten kinetic behavior, which can be affected by pH differences and differences in membrane environment.  相似文献   

6.
Livers of genetically obese Zucker rats showed, compared with lean controls, hypertrophy and enrichment in triacylglycerols, indicating that fatty acid metabolism was directed towards lipogenesis and esterification rather than towards fatty acid oxidation. Mitochondrial activities of cytochrome c oxidase and monoamine oxidase were significantly lower when expressed per g wet wt. of liver, whereas peroxisomal activities of urate oxidase and palmitoyl-CoA-dependent NAD+ reduction were unchanged. Liver mitochondria were able to oxidize oleic acid at the same rate in both obese and lean rats. For reactions occurring inside the mitochondria, e.g. octanoate oxidation and palmitoyl-CoA dehydrogenase, no difference was found between both phenotypes. Total carnitine palmitoyl-, octanoyl- and acetyl-transferase activities were slightly higher in mitochondria from obese rats, whereas the carnitine content of both liver tissue and mitochondria was significantly lower in obese rats compared with their lean littermates. The carnitine palmitoyltransferase I activity was slightly higher in liver mitochondria from obese rats, but this enzyme was more sensitive to malonyl-CoA inhibition in obese than in lean rats. The above results strongly suggest that the impaired fatty acid oxidation observed in the whole liver of obese rats is due to the diminished transport of fatty acids across the mitochondrial inner membrane via the carnitine palmitoyltransferase I. This effect could be reinforced by the decreased mitochondrial content per g wet wt. of liver. The depressed fatty acid oxidation may explain in part the lipid infiltration of liver observed in obese Zucker rats.  相似文献   

7.
L-Carnitine (L-beta-hydroxy-gamma-N,N,N-trimethylaminobutyric acid) plays an essential role in fatty acid transport in the mitochondrion. Conditions that appear to benefit from exogenous supplementation of L-carnitine include anorexia, chronic fatigue, cardiovascular disease, hypoglycemia, male infertility, muscular myopathies, renal failure and dialysis. D-Carnitine is not biologically active and might interfere with the proper utilization of the L isomer, and so there are claims that the racemic mixture (DL-carnitine) should be avoided. Despite the fact that it is known about the systemic manifestations of oral intake of this compound, oral supplementation with DL-carnitine for treatment of primary and secondary carnitine deficiency syndromes has been used in Russia for 25 years. The purpose of the present review was to contrast the differences in pharmacokinetics, phannacodynamics, biochemistry, and toxicity between treatments of L- and DL-carnitine. There is some evidence that L-carnitine and D-carnitine compete for uptake in small intestine and tubular re-absorption in kidneys. After intestinal absorption, L- and D-carnitine is transferred to organs whose metabolism is dependent on fatty acid oxidation, such as heart and skeletal muscle, and D-carnitine competitively depletes muscle level of L-carnitine. Whereas L-carnitine is found to be essential for the oxidation of fatty acids, D-carnitine causes a depletion of L-carnitine, and hindered fatty acid oxidation and energy formation. Pharmacological effects of carnitine are stereospecific, since L-carnitine was effective in various animals and clinical studies, while D- and DL-carnitine was found to be ineffective or toxic, for example, to muscle cells and to the myocardium. DL-Carnitine causes symptoms of myasthenia and cardiac arrhythmias, which disappeared after L-carnitine administration. Clinically toxic effect of D-carnitine was described in patients with renal failure on long-term haemodialysis, in adriamycin (doxorubicin) cardiotoxicity and in stable angina pectoris.  相似文献   

8.
The effects of L-carnitine on respiratory chain enzymes in muscle of long distance runners were studied in 14 athletes. These subjects received placebo or L-carnitine (2 g orally b.i.d.) during a 4-week period of training. Athletes receiving L-carnitine showed a significant increase (p < 0.01) in the activities of rotenone-sensitive NADH cytochrome c reductase, succinate cytochrome c reductase and cytochrome oxidase. In contrast, succinate dehydrogenase and citrate synthase were unchanged. No significant changes were observed after placebo administration. The levels of both total and free carnitine from athletes receiving placebo were significantly decreased (p < 0.01) after treatment. By contrast, total and free carnitine levels were markedly increased (p < 0.01) after supplementation with L-carnitine. Our results suggest that L-carnitine induces an increase of the respiratory chain enzyme activities in muscle, probably by mechanisms involving mitochondrial DNA.  相似文献   

9.
In double-blind cross-over experiments, ten moderately trained male subjects were submitted to two bouts of maximal cycle ergometer exercise separated by a 3 day interval. Each subject was randomly given either L-carnitine (2 g) or placebo orally 1 h before the beginning of each exercise session. At rest L-carnitine supplementation resulted in an increase of plasma-free carnitine without a change in acid-soluble carnitine esters. Treatment with L-carnitine induced a significant post-exercise decrease of plasma lactate and pyruvate and a concurrent increase of acetylcarnitine. The determination of the individual carnitine esters in urine collected for 24 h after the placebo exercise trial revealed a decrease of acetyl carnitine and a parallel increase of a C4 carnitine ester, probably isobutyrylcarnitine. Conversely, acetylcarnitine was strongly increased and C4 compounds were almost suppressed in the L-carnitine loading trial. These results suggest that L-carnitine administration prior to high-intensity exercise stimulates pyruvate dehydrogenase activity, thus diverting pyruvate from lactate to acetylcarnitine formation.  相似文献   

10.
The inhibition of carnitine acetyltransferase (EC 2.3.1.7) by metabolites of 4-pentenoic acid was studied. 3-Keto-4-pentenoyl-CoA, a beta-oxidation metabolite of 4-pentenoic acid, was found to be an effective inhibitor of the enzyme in the presence, but not in the absence of L-carnitine. Since acetyl-CoA protects the enzyme against this inhibition, 3-keto-4-pentenoyl-CoA seems to be an active site-directed inhibitor. 3-Keto-4-pentenoyl-CoA, which is a substrate of carnitine acetyltransferase, causes the irreversible inactivation of the enzyme. All observations together lead to the suggestion that 3-keto-4-pentenoyl-CoA is a mechanism-based inhibitor of carnitine acetyltransferase.  相似文献   

11.
Human brain short chain L-3-hydroxyacyl-CoA dehydrogenase (SCHAD) was found to catalyze the oxidation of 17beta-estradiol and dihydroandrosterone as well as alcohols. Mitochondria have been demonstrated to be the proper location of this NAD+-dependent dehydrogenase in cells, although its primary structure is identical to an amyloid beta-peptide binding protein reportedly associated with the endoplasmic reticulum (ERAB). This fatty acid beta-oxidation enzyme was identified as a novel 17beta-hydroxysteroid dehydrogenase responsible for the inactivation of sex steroid hormones. The catalytic rate constant of the purified enzyme was estimated to be 0.66 min-1 with apparent Km values of 43 and 50 microM for 17beta-estradiol and NAD+, respectively. The catalytic efficiency of this enzyme for the oxidation of 17beta-estradiol was comparable with that of peroxisomal 17beta-hydroxysteroid dehydrogenase type 4. As a result, the human SCHAD gene product, a single-domain multifunctional enzyme, appears to function in two different pathways of lipid metabolism. Because the catalytic functions of human brain short chain L-3-hydroxyacyl-CoA dehydrogenase could weaken the protective effects of estrogen and generate aldehydes in neurons, it is proposed that a high concentration of this enzyme in brain is a potential risk factor for Alzheimer's disease.  相似文献   

12.
It is shown that the unusual NAD(P)+-independent quinoprotein alcohol dehydrogenase, said previously to be responsible for oxidation of ethanol during growth of Acinetobacter calcoaceticus LMD 79.39, was in fact isolated from an unidentified organism which contained cytochrome c and which has now been lost. Several genuine strains of A. calcoaceticus do not contain cytochrome c nor do they contain a quinoprotein alcohol dehydrogenase. The enzyme responsible for ethanol oxidation in these bacteria is an inducible NAD+-linked alcohol dehydrogenase.  相似文献   

13.
In this work, an attempt was made to identify the reasons of impaired long-chain fatty acid utilization that waspreviously described in volume-overloaded rat hearts. The most significant data are the following: (1) The slowing down of long-chain fatty acid oxidation in severely hypertrophied hearts cannot be related to a feedback inhibition of carnitine palmitoyltransferase I from an excessive stimulation of glucose oxidation since, because of decreased tissue levels of L-carnitine, glucose oxidation also declines in volume-overloaded hearts. (2) While, in control hearts, the estimated intracellular concentrations of free carnitine are in the range of the respective Km of mitochondrial CPT I, a kinetic limitation of this enzyme could occur in hypertrophied hearts due to a 40% decrease in free carnitine. (3) However, the impaired palmitate oxidation persists upon the isolation of the mitochondria from these hearts even in presence of saturating concentrations of L-carnitine. In contrast, the rates of the conversion of both palmitoyl-CoA and palmitoylcarnitine into acetyl-CoA are unchanged. (4) The kinetic analyses of palmitoyl-CoA synthase and carnitine palmitoyltransferase I reactions do not reveal any differences between the two mitochondrial populations studied. On the other hand, the conversion of palmitate into palmitoylcarnitine proves to be substrate inhibited already at physiological concentrations of exogenous palmitate. The data presented in this work demonstrate that, during the development of a severe cardiac hypertrophy, a fragilization of the mitochondrial outer membrane may occur. The functional integrity of this membrane seems to be further deteriorated by increasing concentrations of free fatty acids which gives rise to an impaired functional cooperation between palmitoyl-CoA synthase and carnitine palmitoyltransferase I. In intact myocardium, the utilization of the generated in situ palmitoyl-CoA can be further slowed down by decreased intracellular concentrations of free carnitine.  相似文献   

14.
Laties GG 《Plant physiology》1983,72(4):953-958
The oxidation isotherms for citrate and isocitrate by potato (Solanum tuberosum var. Russet Burbank) mitochondria in the presence of NAD differ markedly. Citrate oxidation shows positively cooperative kinetics with a sigmoid isotherm, whereas isocitrate oxidation shows Michaelis-Menten kinetics at concentrations up to 3 millimolar, and cooperative kinetics thereafter up to 30 millimolar. In the absence of exogenous NAD, the isocitrate isotherm is sigmoid throughout. The dual isotherm for isocitrate oxidation in the presence of exogenous NAD reflects the operation of two forms of isocitrate dehydrogenase, one in the matrix and one associated with the inner mitochondrial membrane. Whereas in intact mitochondria the activity of the membrane-bound enzyme is insensitive to rotenone, and to butylmalonate, an inhibitor of organic acid transport, isocitrate oxidation by the soluble matrix enzyme is inhibited by both. The membrane-bound isocitrate dehydrogenase does not operate through the NADH dehydrogenase on the outer face of the inner mitochondrial membrane, and is thus considered to face inward. The regulatory potential of isocitrate dehydrogenase in potato mitochondria may be realized by the apportionment of the enzyme between its soluble and bound forms.  相似文献   

15.
Malonyl-CoA significantly increased the Km for L-carnitine of overt carnitine palmitoyltransferase in liver mitochondria from fed rats. This effect was observed when the molar palmitoyl-CoA/albumin concentration ratio was low (0.125-1.0), but not when it was higher (2.0). In the absence of malonyl-CoA, the Km for L-carnitine increased with increasing palmitoyl-CoA/albumin ratios. Malonyl-CoA did not increase the Km for L-carnitine in liver mitochondria from 24h-starved rats or in heart mitochondria from fed animals. The Km for L-carnitine of the latent form of carnitine palmitoyltransferase was 3-4 times that for the overt form of the enzyme. At low ratios of palmitoyl-CoA/albumin (0.5), the concentration of malonyl-CoA causing a 50% inhibition of overt carnitine palmitoyltransferase activity was decreased by 30% when assays with liver mitochondria from fed rats were performed at 100 microM-instead of 400 microM-carnitine. Such a decrease was not observed with liver mitochondria from starved animals. L-Carnitine displaced [14C]malonyl-CoA from liver mitochondrial binding sites. D-Carnitine was without effect. L-Carnitine did not displace [14C]malonyl-CoA from heart mitochondria. It is concluded that, under appropriate conditions, malonyl-CoA may decrease the effectiveness of L-carnitine as a substrate for the enzyme and that L-carnitine may decrease the effectiveness of malonyl-CoA to regulate the enzyme.  相似文献   

16.
Palmitoyl-L carnitine decreases the oxidation of isocitrate in rat liver mitochondria in state 3 by 25-30%. Palmitoyl-L-carnitine acts as an additional substrate raising the rate of oxidative phosphorylation, NAD reduction and ATP/ADP ratio in mitochondria. Palmitoyl-CoA added to mitochondria oxidizing isocitrate in state 3 causes a strong inhibition of isocitrate oxidation and of oxidative phosphorylation and a considerable elevation of intramitochondrial NADH/NAD and ATP/ADP ratios. The effect of palmitoyl-CoA is dependent on its concentration and is competitive with ADP. Carnitine restores only oxidative phosphorylation, but the oxidation of isocitrate remains inhibited. Evidence is presented that the transport of isocitrate is not affected by palmitoyl-CoA is due to the inhibition of adenine nucleotide translocation. The kinetic studies of NAD-dependent isocitrate dehydrogenase in the soluble fraction of sonicated mitochondria revealed that the enzyme is very sensitive towards the inhibition by NADH and only very slightly affected by ATP (Ki for NADH and ATP are 0.017 and 3.6 mM respectively). On the basis of the kinetic data the relative contribution of NADH and ATP in the inhibition of isocitrate oxidation by fatty acids was calculated. It is concluded that the inhibition of isocitrate oxidation caused by palmitoyl-L-carnitine and palmitoyl-CoA is primarily due to the increased reduction of NAD, whereas the increase of ATP/ADP ratio is much less important.  相似文献   

17.
12-L-hydroxy-5,8,10-heptadecatrienoic acid (HHT) was found to be an excellent substrate for NAD+ dependent 15-hydroxyprostaglandin dehydrogenase from porcine kidney. Kcat/Km value of HHT was comparable to that of prostaglandin E although HHT is not a prostanoic acid derivative. Product of enzyme catalyzed oxidation of HHT was identified as 12-keto-5,8,10-heptadecatrienoic acid by gas chromatography-mass spectrometry. The fact that HHT is an excellent substrate for 15-hydroxyprostaglandin dehydrogenase suggest that HHT may have profound unrecognized biological actions and its inactivation may be via oxidation of the hydroxyl group.  相似文献   

18.
L-Propionyl carnitine has been shown to improve the heart's mechanical recovery and other metabolic parameters after ischemia-reperfusion. However, the mechanism of protection is unknown. The two dominating hypotheses are: (i) L-propionyl carnitine can serve as an energy source for heart muscle cells by being enzymatically converted to propionyl-CoA and subsequently utilized in the Krebs cycle (a metabolic hypothesis), and (ii) it can act as an antiradical agent, protecting myocardial cells from oxidative damage (a free radical hypothesis). To test the two possible pathways, we compared the protection afforded to the ischemia-reperfused hearts by L-propionyl carnitine and its optical isomer, D-propionyl carnitine. The latter cannot be enzymatically utilized as an energy source. The Langendorff perfusion technique was used and the hearts were subjected to 40 min of ischemia and 20 min of reperfusion. In analysis of ischemia-reperfused hearts, a strong correlation was found between the recovery of mechanical function and the presence of protein oxidation products (protein carbonyls). Both propionyl carnitines efficiently prevented protein oxidation but L-propionyl carnitine-perfused hearts had two times greater left ventricular developed pressure. The results indicate that both metabolic and antiradical pathway are involved in the protective mechanism of L-propionyl carnitine. To obtain a better insight of the antiradical mechanism of L-propionyl carnitine, we compared the ability of L- and D-propionyl carnitines, L-carnitine, and deferoxamine to interact with: (i) peroxyl radicals, (ii) oxygen radicals, and (iii) iron. We found that none of the carnitine derivatives were able to scavenge peroxyl radicals or superoxide radicals. L- and D-propionyl carnitine and deferoxamine (not L-carnitine) suppressed hydroxyl radical production in the Fenton system, probably by chelating the iron required for the generation of hydroxyl radicals. We suggest that L-propionyl carnitine protects the heart by a dual mechanism: it is an efficient fuel source and an antiradical agent.  相似文献   

19.
The early initiation phase of acute inflammation is anabolic and primarily requires glycolysis with reduced mitochondrial glucose oxidation for energy, whereas the later adaptation phase is catabolic and primarily requires fatty acid oxidation for energy. We reported previously that switching from the early to the late acute inflammatory response following TLR4 stimulation depends on NAD(+) activation of deacetylase sirtuin 1 (SirT1). Here, we tested whether NAD(+) sensing by sirtuins couples metabolic polarity with the acute inflammatory response. We found in TLR4-stimulated THP-1 promonocytes that SirT1 and SirT 6 support a switch from increased glycolysis to increased fatty acid oxidation as early inflammation converts to late inflammation. Glycolysis enhancement required hypoxia-inducing factor-1α to up-regulate glucose transporter Glut1, phospho-fructose kinase, and pyruvate dehydrogenase kinase 1, which interrupted pyruvate dehydrogenase and reduced mitochondrial glucose oxidation. The shift to late acute inflammation and elevated fatty acid oxidation required peroxisome proliferator-activated receptor γ coactivators PGC-1α and β to increase external membrane CD36 and fatty acid mitochondrial transporter carnitine palmitoyl transferase 1. Metabolic coupling between early and late responses also required NAD(+) production from nicotinamide phosphoryltransferase (Nampt) and activation of SirT6 to reduce glycolysis and SirT1 to increase fatty oxidation. We confirmed similar shifts in metabolic polarity during the late immunosuppressed stage of human sepsis blood leukocytes and murine sepsis splenocytes. We conclude that NAD(+)-dependent bioenergy shifts link metabolism with the early and late stages of acute inflammation.  相似文献   

20.
The effect of rotenone on respiration in pea cotyledon mitochondria   总被引:7,自引:7,他引:0       下载免费PDF全文
Respiration utilizing NAD-linked substrates in mitochondria isolated from cotyledons of etiolated peas (Pisum sativum L. var. Homesteader) by sucrose density gradient centrifugation exhibited resistance to rotenone. The inhibited rate of α-ketoglutarate oxidation was equivalent to the recovered rate of malate oxidation. (The recovered rate is the rate following the transient inhibition by rotenone.) The inhibitory effect of rotenone on malate oxidation increased with increasing respiratory control ratios as the mitochondria developed. The cyanide-resistant and rotenone-resistant pathways followed different courses of development as cotyledons aged. The rotenone-resistant pathway transferred reducing equivalents to the cyanide-sensitive pathway. Malic enzyme was found to be inhibited competitively with respect to NAD by rotenone concentrations as low as 1.67 micromolar. In pea cotyledon mitochondria, rotenone was transformed into elliptone. This reduced its inhibitory effect on intact mitochondria. Malate dehydrogenase was not affected by rotenone or elliptone. However, elliptone inhibited malic enzyme to the same extent that rotenone did when NAD was the cofactor. The products of malate oxidation reflected the interaction between malic enzyme and malate dehydrogenase. Rotenone also inhibited the NADH dehydrogenase associated with malate dehydrogenase. Thus, rotenone seemed to exert its inhibitory effect on two enzymes of the electron transport chain of pea cotyledon mitochondria.  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