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.
L C Costello  R Franklin  R Stacey 《Enzyme》1976,21(6):495-506
Mitochondrial preparations isolated from rat ventral prostate were capable of oxidizing isocitrate by way of NADP isocitrate dehydrogenase (NADP-IDH) and NAD-IDH. NAD-IDH activity required ADP for activation. The pH responses for NAD-IDH and NADP-IDH were quite different. The results indicated that two different enzymes were involved in the NAD- and NADP-IDH activities. Indirect evidence indicated that NADPH-NAD transhydrogenase activity might also be involved in the mitochondrial pathway for isocitrate oxidation. NADP-IDH activity was significantly greater than NAD-IDH activity. The oxidation of isocitrate through IDH activity was coupled to the cytochrome system by NADPH- and NADH-cytochrome c reductase activities. Citrate, via isocitrate, oxidation proceeded at a much slower rate suggesting that aconitase activity could be limiting in the oxidation of citrate. In comparison to other tissues, the prostate oxidative enzyme activities are considerably lower. The results suggest that the accumulation of high prostate citrate levels is not due to a limitation imposed by a lack of IDH activity in prostate mitochondria.  相似文献   

2.
In the green alga Chlamydomonas reinhardtii , nitrogen staravation induced a reversible increase (2-fold) in NAD-isocitrate dehydrogenase (NAD-IDH; EC 1.1.1.41) and NADP-isocitrate dehydrogenase (NADP-IDH; EC 1.1.1.42) activities. Both enzymes were not affected by the concentration of CO2, the dark or the nature of the nitrogen source (nitrate, nitrite, or ammonium). When cells growing autotrophically were transferred to heterotrophic conditions, a 40% reduction of the NAD-IDH activity was detected, a 2-fold increase of NADP-IDH was observed and isocitrate lyase (ICL; EC 4.1.3.1) activity was induced. The replacement of autotrophic conditions led to the initial activity levels. NAD- and NADP-IDH activities showed markedly different patterns of increase in synchronous cultures of this alga obtained by 12 h light/12 h dark transitions. While NAD-IDH increased in the last 4 h of the dark period, NADP-IDH increased during the last 4 h of the light period, remaining constant for the rest of the cycle.  相似文献   

3.
Combined effects of cadmium (Cd) and temperature on key mitochondrial enzymes [including Complexes I-IV of electron transport chain and Krebs cycle enzymes citrate synthase (CS), and NAD- and NADP-dependent isocitrate dehydrogenases (NAD-IDH and NADP-IDH)] were studied in a marine ectotherm, Crassostrea virginica in order to better understand the mechanisms of Cd-induced impairment of mitochondrial function. Matrix enzymes including CS and isocitrate dehydrogenases were the most sensitive to Cd making Krebs cycle a likely candidate to explain Cd-induced impairment of mitochondrial substrate oxidation. CS and NAD-IDH had IC(50) of 26 and 65 microM at the acclimation temperature (15 degrees C) and 65 (CS) and 1.5 (NAD-IDH) microM at elevated temperature (25 degrees C), respectively. Mitochondrial NADP-IDH was the most sensitive to Cd with IC(50) of 14 and 3.4 microM at 15 degrees and 25 degrees C, respectively. Electron transport chain (ETC) complexes were significantly less sensitive to the direct effects of Cd with IC(50) ranging from 260 to >400 microM. Temperature increase led to a higher sensitivity of mitochondrial enzymes to the inhibitory effects of Cd as indicated by a decline in IC(50) with the exception of Complex III from gills and CS from gills and hepatopancreas. Cd exposure also resulted in a decrease in activation energy of mitochondrial enzymes suggesting that mitochondria from Cd-exposed oysters could exhibit reduced capacity to respond to temperature rise with an adequate increase in the substrate flux. These interactive effects of Cd and temperature on mitochondrial enzymes could negatively affect metabolic performance of oysters and possibly other ectotherms in polluted environments during temperature increase such as expected during the global climate change and/or tidal or seasonal warming in estuarine and coastal waters.  相似文献   

4.
Summary The subcellular distribution and substrate kinetics of soluble and particulate-associated bovine adrenal glycerol kinase have been investigated. Whole adrenal, adrenal cortex and adrenal medulla were examined for distribution of glycerol kinase between soluble and particulate fractions. No major differences in distribution were noted between these tissues; of the total homogenate activity, 0–20% sedimented with the nuclear fraction, 24–36% sedimented with the post-nuclear fraction and 62–69% remained soluble. Steadystate kinetic parameters of glycerol kinase activity were compared in the soluble and mitochondrial fractions. The Km for glycerol in the soluble fraction was 6.3 ± 0.1 M and in the mitochondrial fraction was 4.0 = 0.3 M. The Km for ATP in soluble fraction was 12.8 1.5 and in the mitochondrial fraction was 5.3 ± 1.6. Release of adrenal glycerol kinase from the mitochondria) fraction was investigated using inorganic phosphate, ATP and glycerol 3-phosphate. Of these compounds, only ATP and glycerol 3-phosphate were effective in releasing particulate-associated glycerol kinase. Inorganic phosphate had no effect upon release. Particulate-associated glycerol kinase activity of the mitochondrial fraction was stimulated by addition of succinate and ADP and was inhibited by addition of atractyloside. The data presented here indicate that bound glycerol kinase found within the mitochondrial fraction is kinetically distinct from soluble glycerol kinase and binding to mitochondria is responsive to substrate and product levels within the physiological range.  相似文献   

5.
This study revealed that cytosolic aconitase (ACO, EC 4.2.1.3) and isocitrate lyase (ICL, EC 4.1.3.1, marker of the glyoxylate cycle) are active in germinating protein seeds of yellow lupine. The glyoxylate cycle seems to function not only in the storage tissues of food-storage organs, but also in embryonic tissue of growing embryo axes. Sucrose (60 mM) added to the medium of in vitro culture of embryo axes and cotyledons decreased activity of lipase (LIP, EC 3.1.1.3) and activity of glutamate dehydrogenase (NADH-GDH, EC 1.4.1.2). The opposite effect was caused by sucrose on activity of cytosolic ACO, ICL as well as NADP+-dependent (EC 1.1.1.42) and NAD+-dependent (EC 1.1.1.41) isocitrate dehydrogenase (NADP-IDH and NAD-IDH, respectively); activity of these enzymes was clearly stimulated by sucrose. Changes in the activity of LIP, ACO, NADP-IDH, and NAD-IDH caused by sucrose were based on modifications in gene expression because corresponding changes in the enzyme activities and in the mRNA levels were observed. The significance of cytosolic ACO and NADP-IDH in carbon flow from storage lipid to amino acids, as well as the peculiar features of storage lipid breakdown during germination of lupine seeds are discussed.  相似文献   

6.
The role of Asp-328 and Ile-329 as a cofactor discrimination site of the NAD-dependent isocitrate dehydrognase (NAD-IDH) from Pyrococcus furiosus has been verified by replacing these residues with Lys and Tyr, respectively, which are the corresponding residues in NADP-IDH from Escherichia coli. The Asp-328-Lys mutant showed dual coenzyme specificity, whereas introduction of the double mutation, Asp-328-Lys/Ile-329-Tyr shifted the cofactor preference from NAD to NADP. NADP-dependent P. furiosus IDH retained thermostability and thermoactivity compared with NAD-IDH.  相似文献   

7.
NADP-Utilizing Enzymes in the Matrix of Plant Mitochondria   总被引:9,自引:4,他引:5       下载免费PDF全文
Purified potato tuber (Solanum tuberosum L. cv Bintie) mitochondria contain soluble, highly latent NAD+- and NADP+-isocitrate dehydrogenases, NAD+- and NADP+-malate dehydrogenases, as well as an NADPH-specific glutathione reductase (160, 25, 7200, 160, and 16 nanomoles NAD(P)H per minute and milligram protein, respectively). The two isocitrate dehydrogenase activities, but not the two malate dehydrogenase activities, could be separated by ammonium sulfate precipitation. Thus, the NADP+-isocitrate dehydrogenase activity is due to a separate matrix enzyme, whereas the NADP+-malate dehydrogenase activity is probably due to unspecificity of the NAD+-malate dehydrogenase. NADP+-specific isocitrate dehydrogenase had much lower Kms for NADP+ and isocitrate (5.1 and 10.7 micromolar, respectively) than the NAD+-specific enzyme (101 micromolar for NAD+ and 184 micromolar for isocitrate). A broad activity optimum at pH 7.4 to 9.0 was found for the NADP+-specific isocitrate dehydrogenase whereas the NAD+-specific enzyme had a sharp optimum at pH 7.8. Externally added NADP+ stimulated both isocitrate and malate oxidation by intact mitochondria under conditions where external NADPH oxidation was inhibited. This shows that (a) NADP+ is taken up by the mitochondria across the inner membrane and into the matrix, and (b) NADP+-reducing activities of malate dehydrogenase and the NADP+-specific isocitrate dehydrogenase in the matrix can contribute to electron transport in intact plant mitochondria. The physiological relevance of mitochondrial NADP(H) and soluble NADP(H)-consuming enzymes is discussed in relation to other known mitochondrial NADP(H)-utilizing enzymes.  相似文献   

8.
An antibody that inhibits over 95% of the cytosolic NADP+-dependent gamma-hydroxybutyrate (GHB) dehydrogenase activity of either rat brain or kidney was found to inhibit only approximately 50% of the conversion of [1-14C]GHB to 14CO2 by rat kidney homogenate. A similar result was obtained with sodium valproate, a potent inhibitor of GHB dehydrogenase. The mitochondrial fraction from rat brain and kidney was found to catalyze the conversion of [1-14C]GHB to 14CO2. The dialyzed mitochondrial fraction also catalyzed the oxidation of GHB to succinic semialdehyde (SSA) in a reaction that did not require added NAD+ or NADP+ and which was not inhibited by sodium valproate. The enzyme from the mitochondrial fraction which converts GHB to SSA appears to be distinct from the NADP+-dependent cytosolic oxidoreductase which catalyzes this reaction.  相似文献   

9.
1. The NAD(P) transhydrogenase activity of the soluble fraction of sonicated rat liver mitochondrial preparations was greater than the NAD-linked isocitrate dehydrogenase activity, and the NAD-linked and NADP-linked isocitrate dehydrogenase activities were not additive. The NAD-linked isocitrate dehydrogenase activity was destroyed by an endogenous autolytic system or by added nucleotide pyrophosphatase, and was restored by a catalytic amount of NADP. 2. We concluded that the isocitrate dehydrogenase of rat liver mitochondria was exclusively NADP-specific, and that the oxoglutarate/isocitrate couple could therefore be used unequivocally as redox reactant for NADP in experiments designed to operate only the NAD(P) transhydrogenase (or loop 0) segment of the respiratory chain in intact mitochondria. 3. During oxidation of isocitrate by acetoacetate in intact, anaerobic, mitochondria via the rhein-sensitive, but rotenone- and arsenite-insensitive, NAD(P) transhydrogenase, measurements of the rates of carbonyl cyanide p-trifluoromethoxyphenylhydrazone-sensitive and carbonyl cyanide p-trifluoromethoxyphenylhydrazone-insensitive pH change in the presence of various oxoglutarate/isocitrate concentration ratios gave an -->H(+)/2e(-) quotient of 1.94+/-0.12 for outward proton translocation by the NAD(P) transhydrogenase. 4. Measurements with a K(+)-sensitive electrode confirmed that the electrogenicity of the NAD(P) transhydrogenase reaction corresponded to the translocation of one positive charge per acid equivalent. 5. Sluggish reversal of the NAD(P) transhydrogenase reaction resulted in a significant inward proton translocation. 6. The possibility that isocitrate might normally be oxidized via loop 0 at a redox potential of -450mV, or even more negative, is discussed, and implies that a P/O quotient of 4 for isocitrate oxidation might be expected.  相似文献   

10.
NAD+-isocitrate dehydrogenase and 2-oxoglutarate dehydrogenase in extracts of mitochondria from the highly malignant AS-30D rat hepatoma cell line demonstrate Ca2+ sensitivities and affinities for substrates similar to those of normal liver mitochondria. However, the maximal activities of NAD+- and NADP+-dependent isocitrate dehydrogenase were found to be 8 and 3.5 fold higher in hepatoma mitochondrial extracts than those of liver mitochondria, whereas maximal activities of succinate and 2-oxoglutarate dehydrogenases were similar in the two tissues. At pyridine nucleotide concentrations giving the lowest physiological NADH/NAD+ ratio, NAD+-isocitrate dehydrogenase activity in hepatoma mitochondrial extracts was completely inhibited at subsaturating concentrations of Ca2+, substrate, and NAD+, in contrast to rat liver mitochondrial extracts which retained significant activity.  相似文献   

11.
X-irradiation (100-1500 r) administered to the heads of rats 8-30 days of age inhibited the development of glycerol phosphate dehydrogenase (l-glycerol 3-phosphate-NAD oxidoreductase, EC 1.1.1.8) in the brain stem and cerebral hemispheres. At 40 days of age and older no effect was observed. This inhibition was a delayed phenomenon, dose-dependent and with no recovery. It is proposed that the inhibition of enzyme formation is related to radiation damage caused to DNA. Actinomycin D inhibited the development of glycerol phosphate dehydrogenase in a manner similar to ionizing radiation. Four other dehydrogenases also showed age-dependent radiosensitivities. ;Malic enzyme' (EC 1.1.1.40), lactate dehydrogenase (EC 1.1.1.27) and malate dehydrogenase (EC 1.1.1.37) ceased to be radiosensitive at about 8 days of age and isocitrate dehydrogenase (NADP) (EC 1.1.1.42) at 16 days. The correlation between developmental increase in enzyme activity and radiosensitivity held closely for glycerol phosphate dehydrogenase and isocitrate dehydrogenase and to a smaller extent for the others.  相似文献   

12.
Effects of 1-methyl-4-phenylpyridinium ion (MPP+) on the activities of NAD+- or NADP+-linked dehydrogenases in the TCA cycle were studied using mitochondria prepared from mouse brains. Activities of NAD+- and NADP+-linked isocitrate dehydrogenases, NADH- and NADPH-linked glutamate dehydrogenases, and malate dehydrogenase were little affected by 2 mM of MPP+. However, alpha-ketoglutarate dehydrogenase activity was significantly inhibited by MPP+. Kinetic analysis revealed a competitive type of inhibition. Inhibition of alpha-ketoglutarate dehydrogenase may be one of the important mechanisms of MPP+-induced inhibition of mitochondrial respiration, and of neuronal degeneration.  相似文献   

13.
A soluble NAD+-linked isocitrate dehydrogenase has been isolated from Crithidia fasciculata. The enzyme was purified 128-fold, almost to homogeneity, and was highly specific for NAD+ as the coenzyme. There is also a cytoplasmic NADP+-linked and a mitochondrial isocitrate dehydrogenase in the organism. Studies of the physical and kinetic properties of the soluble NAD+-isocitrate dehydrogenase from this organism showed that it resembled microbial NADP+-isocitrate dehydrogenases in general, all of which are cytoplasmic enzymes. The enzyme appeared not to be related to other NAD+-isocitrate dehydrogenases, which are found in the mitochondria of eukaryotic cells. The molecular weight of the soluble NAD+-isocitrate dehydrogenase was 105,000 which is within the range of the values for microbial NADP+-isocitrate dehydrogenases. Similar to the NADP+-isocitrate dehydrogenase in this organism, the enzyme was inhibited in a concerted manner by glyoxalate plus oxalacetate. Kinetic analysis revealed that Mn2+ was involved in the binding of isocitrate to the enzyme. Inhibition of the NAD+-linked isocitrate dehydrogenase by p-chloromercuribenzoate could be prevented by prior incubation of the enzyme with both Mn2+ and isocitrate; however, neither ion alone conferred protection. Free isocitrate, free Mn2+, and the Mn2+-isocitrate complex could all bind to the enzyme. Four different mechanisms with respect to the binding of isocitrate to the enzyme were tested. Of these, the formation of the active enzyme-Mn2+-isocitrate complex from (a) the random binding of Mn2+, isocitrate, and the Mn2+-isocitrate complex, or (b) the binding of Mn2+-isocitrate with free Mn2+ and isocitrate acting as dead-end competitors were both in agreement with these data.  相似文献   

14.
The net uptake of 45Ca into mitochondria from pancreatic islets is stimulated by substrates that transfer reducing equivalents to various sites of the respiratory chain, such as succinate or glycerol 3-phosphate (site II), malate plus pyruvate (site I) or ascorbate plus TMPD (site III). Diazoxide, a known inhibitor of insulin release in vivo and in vitro, strongly inhibited net 45Ca uptake supported by glycerol phosphate and succinate and weakly inhibited 45Ca uptake supported by the other substrates. These results suggest that diazoxide, although not completely specific, is predominately an inhibitor at site II of the respiratory chain. This result is consistent with previous work that showed diazoxide inhibits the enzyme activity of the mitochondrial glycerol phosphate dehydrogenase in islets. Sodium ion inhibited the net accumulation of 45Ca by islet mitochondria suggesting a similarity between islet mitochondria and those of heart and some other endocrine tissues.  相似文献   

15.
SUCCINATE DEHYDROGENASE (SUCCINATE: phenazine methosulfate oxidoreductase, EC 1.3.99.1) activity in crude mitochondrial fraction from pea (var. Alaska) cotyledons increased during seed imbibition to reach a maximum after about 12 hours. The increase was not inhibited by either cycloheximide or d(-)threo-chloramphenicol. The postmicrosomal fraction from dry cotyledons, but not that from fully imbibed ones, contained a soluble form of succinate dehydrogenase. The soluble enzyme was partially purified by ammonium sulfate fractionation and diethylaminoethyl-cellulose and Sepharose 6B column chromatography. The enzyme showed no succinate-coenzyme Q oxidoreductase activity and had a molecular mass of about 100,000 daltons. The soluble enzyme seemed to differ only slightly from succinate dehydrogenase solubilized from the mitochondrial inner membrane from fully imbibed cotyledons by a detergent. It is proposed that the soluble succinate dehydrogenase is associated with an inert mitochondrial inner membrane in dry cotyledons to form an active one during seed imbibition.  相似文献   

16.
Homogenates of isolated pancreatic islets contain 40-70 times as much flavin-linked glycerol-3-phosphate dehydrogenase (EC 1.1.99.5) as homogenates of whole pancreas, liver, heart, or skeletal muscle when the activity is assayed with either iodonitrotetrazolium or with dichloroindophenol as an electron acceptor. Intact mitochondria from islets release 3HOH from [2-3H]glycerol phosphate 7 times faster than do skeletal muscle mitochondria. The activity of the cytosolic, NAD-linked, glycerol phosphate dehydrogenase (EC 1.1.1.8) in pancreatic islets is comparable to that of the mitochondrial dehydrogenase so a glycerol phosphate shuttle is possible in pancreatic islets. Diazoxide, an inhibitor of insulin release in vivo and in vitro, inhibits the islet mitochondrial glycerol phosphate dehydrogenase in all three of the assays mentioned above at concentrations that inhibit insulin release and CO2 formation from glucose by isolated pancreatic islets. Diazoxide does not inhibit the dehydrogenase in mitochondria from skeletal muscle, liver, and heart. A slight inhibition in mitochondria from whole pancreas can be accounted for as inhibition of the islet dehydrogenase because no inhibition is observed in mitochondria from pancreas of rats treated with alloxan, an agent that causes diabetes by destroying pancreatic beta cells. The results of this study are compatible with the hypothesis that the mitochondrial glycerol phosphate dehydrogenase has a key role in stimulus-secretion coupling in the pancreatic beta cell during glucose-induced insulin release.  相似文献   

17.
The yeast Candida parapsilosis possesses two routes of electron transfer from exogenous NAD(P)H to oxygen. Electrons are transferred either to the classical cytochrome pathway at the level of ubiquinone through an NAD(P)H dehydrogenase, or to an alternative pathway at the level of cytochrome c through another NAD(P)H dehydrogenase which is insensitive to antimycin A. Analyses of mitoplasts obtained by digitonin/osmotic shock treatment of mitochondria purified on a sucrose gradient indicated that the NADH and NADPH dehydrogenases serving the alternative route were located on the mitochondrial inner membrane. The dehydrogenases could be differentiated by their pH optima and their sensitivity to amytal, butanedione and mersalyl. No transhydrogenase activity occurred between the dehydrogenases, although NADH oxidation was inhibited by NADP+ and butanedione. Studies of the effect of NADP+ on NADH oxidation showed that the NADH:ubiquinone oxidoreductase had Michaelis-Menten kinetics and was inhibited by NADP+, whereas the alternative NADH dehydrogenase had allosteric properties (NADH is a negative effector and is displaced from its regulatory site by NAD+ or NADP+).  相似文献   

18.
Plant mitochondria have the unique ability to directly oxidize exogenous NAD(P)H. We recently separated two NAD(P)H dehydrogenase activities from maize (Zea mays L.) mitochondria using anion-exchange (Mono Q) chromatography. The first peak of activity oxidized only NADH, whereas the second oxidized both NADH and NADPH. In this paper we describe the purification of the first peak of activity to a 32-kD protein. Polyclonal antibodies to the 32-kD protein were used to show that it was present in mitochondria from several plant species. Two-dimensional gel analysis of the 32-kD NADH dehydrogenase indicated that it consisted of two major and one minor isoelectric forms. Immunoblot analysis of submitochondrial fractions indicated that the 32-kD protein was enriched in the soluble protein fraction after mitochondrial disruption and fractionation; however, some association with the membrane fraction was observed. The membrane-impermeable protein cross-linking agent 3,3[prime] -dithiobis-(sulfosuccinimidylpropionate) was used to further investigate the submitochondrial location of the 32-kD NADH dehydrogenase. The 32-kD protein was localized to the outer surface of the inner mitochondrial membrane or to the intermembrane space. The pH optimum for the enzyme was 7.0. The activity was found to be severely inhibited by p-chloromercuribenzoic acid, mersalyl, and dicumarol, and stimulated somewhat by flavin mononucleotide.  相似文献   

19.
Subcellular distribution of pentose-phosphate cycle enzymes in rat liver was investigated, using differential and isopycnic centrifugation. The activities of the NADP+-dependent dehydrogenases of the pentose-phosphate pathway (glucose-6-phosphate dehydrogenase and phosphogluconate dehydrogenase) were detected in the purified peroxisomal fraction as well as in the cytosol. Both dehydrogenases were localized in the peroxisomal matrix. Chronic administration of the hypolipidemic drug clofibrate (ethyl-alpha-p-chlorophenoxyisobutyrate) caused a 1.5-2.5-fold increase in the amount of glucose-6-phosphate and phosphogluconate dehydrogenases in the purified peroxisomes. Clofibrate decreased the phosphogluconate dehydrogenase, but did not alter glucose-6-phosphate dehydrogenase activity in the cytosolic fraction. The results obtained indicate that the enzymes of the non-oxidative segment of the pentose cycle (transketolase, transaldolase, triosephosphate isomerase and glucose-phosphate isomerase) are present only in a soluble form in the cytosol, but not in the peroxisomes or other particles, and that ionogenic interaction of the enzymes with the mitochondrial and other membranes takes place during homogenization of the tissue in 0.25 M sucrose. Similar to catalase, glucose-6-phosphate dehydrogenase and phosphogluconate dehydrogenase are present in the intact peroxisomes in a latent form. The enzymes have Km values for their substrates in the millimolar range (0.2 mM for glucose-6-phosphate and 0.10-0.12 mM for 6-phosphogluconate). NADP+, but not NAD+, serves as a coenzyme for both enzymes. Glucose-6-phosphate dehydrogenase was inhibited by palmitoyl-CoA, and to a lesser extent by NADPH. Peroxisomal glucose-6-phosphate and phosphogluconate dehydrogenases have molecular mass of 280 kDa and 96 kDa, respectively. The putative functional role of pentose-phosphate cycle dehydrogenases in rat liver peroxisomes is discussed.  相似文献   

20.
The reversible, membrane-associated transhydrogenase that catalyzes hydride-ion transfer between NADP(H) and NAD(H) was evaluated and compared to the corresponding NADH oxidase and succinate dehydrogenase activities in midgut and fat body mitochondria from fifth larval instar Manduca sexta. The developmentally significant NADPH-forming transhydrogenation occurs as a nonenergy- or energy-linked activity with energy for the latter derived from either electron transport-dependent NADH or succinate utilization, or ATP hydrolysis by Mg++-dependent ATPase. In general, the plant flavonoids examined (chyrsin, juglone, morine, quercetin, and myricetin) affected all reactions in a dose-dependent fashion. Differences in the responses to the flavonoids were apparent, with the most notable being inhibition of midgut, but stimulation of fat body transhydrogenase by morin, and myricetin as also noted for NADH oxidase and succinate dehydrogenase. Although quercetin inhibited or stimulated transhydrogenase activity depending on the origin of mitochondria, it was without effect on either midgut or fat body NADH oxidase or succinate dehydrogenase. Observed sonication-dependent increases in flavonoid inhibition may well reflect an alteration in membrane configuration, resulting in increased exposure of the enzyme systems to the flavonoids. The effects of flavonoids on the transhydrogenation, NADH oxidase, and succinate dehydrogenase reactions suggest that compounds of this nature may prove valuable in the control of insect populations by affecting these mitochondrial enzyme components.  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