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.
Ho KK  Hurley TD  Weiner H 《Biochemistry》2006,45(31):9445-9453
Random mutagenesis followed by a filter-based screening assay has been used to identify a mutant of human class 1 aldehyde dehydrogenase (ALDH1) that was no longer inhibited by Mg(2+) ions but was activated in their presence. Several mutants possessed double, triple, and quadruple amino acid substitutions with a total of seven different residues being altered, but each had a common T244S change. This point mutation proved to be responsible for the Mg(2+) ion activation. An ALDH1 T244S mutant was recombinantly expressed and was used for mechanistic studies. Mg(2+) ions have been shown to increase the rate of deacylation. Consistent with the rate-limiting step for ALDH1 being changed from coenzyme dissociation to deacylation was finding that chloroacetaldehyde was oxidized more rapidly than acetaldehyde. Furthermore, Mg(2+) ions only in the presence of NAD(H) increased the rate of hydrolysis of p-nitrophenyl acetate showing that the metal only affects the binary complex. Though the rate-limiting step for the T244S mutant was different from that of the native enzyme, the catalytic efficiency of the mutant was just 20% that of the native enzyme. The basis for the change in the rate-limiting step appears to be related to NAD(+) binding. Using the structure of a sheep class 1 ALDH, it was possible to deduce that the interaction between the side chain of T244 and its neighboring residues with the nicotinamide ring of NAD(+) were an essential determinant in the catalytic action of ALDH1.  相似文献   

2.
Aldehyde dehydrogenase 2 (ALDH2) catalyzes oxidation of toxic aldehydes to carboxylic acids. Physiologic levels of Mg(2+) ions influence ALDH2 activity in part by increasing NADH binding affinity. Traditional fluorescence measurements monitor the blue shift of the NADH fluorescence spectrum to study ALDH2-NADH interactions. By using time-resolved fluorescence spectroscopy, we have resolved the fluorescent lifetimes (τ) of free NADH (τ=0.4 ns) and bound NADH (τ=6.0 ns). We used this technique to investigate the effects of Mg(2+) on the ALDH2-NADH binding characteristics and enzyme catalysis. From the resolved free and bound NADH fluorescence signatures, the K(D) for NADH with ALDH2 ranged from 468 μM to 12 μM for Mg(2+) ion concentrations of 20 to 6000 μM, respectively. The rate constant for dissociation of the enzyme-NADH complex ranged from 0.4s(-1) (6000 μM Mg(2+)) to 8.3s(-1) (0 μM Mg(2+)) as determined by addition of excess NAD(+) to prevent re-association of NADH and resolving the real-time NADH fluorescence signal. The apparent NADH association/re-association rate constants were approximately 0.04 μM(-1)s(-1) over the entire Mg(2+) ion concentration range and demonstrate that Mg(2+) ions slow the release of NADH from the enzyme rather than promoting its re-association. We applied NADH fluorescence lifetime analysis to the study of NADH binding during enzyme catalysis. Our fluorescence lifetime analysis confirmed complex behavior of the enzyme activity as a function of Mg(2+) concentration. Importantly, we observed no pre-steady state burst of NADH formation. Furthermore, we observed distinct fluorescence signatures from multiple ALDH2-NADH complexes corresponding to free NADH, enzyme-bound NADH, and, potentially, an abortive NADH-enzyme-propanal complex (τ=11.2 ns).  相似文献   

3.
Aldehyde dehydrogenases are general detoxifying enzymes, but there are also isoenzymes that are involved in specific metabolic pathways in different organisms. Two of these enzymes are Escherichia coli lactaldehyde (ALD) and phenylacetaldehyde dehydrogenases (PAD), which participate in the metabolism of fucose and phenylalanine, respectively. These isozymes share some properties with the better characterized mammalian enzymes but have kinetic properties that are unique. It was possible to thread the sequences into the known ones for the mammalian isozymes to better understand some structural differences. Both isozymes were homotetramers, but PAD used both NAD+ and NADP+ but with a clear preference for NAD, while ALD used only NAD+. The rate-limiting step for PAD was hydride transfer as indicated by the primary isotopic effect and the absence of a pre-steady-state burst, something not previously found for tetrameric enzymes from other organisms where the rate-limiting step is related to both deacylation and coenzyme dissociation. In contrast, ALD had a pre-steady-state burst indicating that the rate-limiting step was located after the NADH formation, but the rate-limiting step was a combination of deacylation and coenzyme dissociation. Both enzymes possessed esterase activity that was stimulated by NADH; NAD+ stimulated the esterase activity of PAD but not of ALD. Finding enzymes that structurally are similar to the well-characterized mammalian enzymes but have a different rate-limiting step might serve as models to allow us to determine what regulates the rate-limiting step.  相似文献   

4.
Usual human livers contain two major aldehyde dehydrogenase isozymes, cytosolic ALDH1 component and mitochondrial ALDH2 component, while human livers with "atypical" phenotype have only ALDH1 isozyme and are missing ALDH2 isozyme. Approximately 50% of orientals are atypical in respect to ALDH isozymes. We previously demonstrated an existence of enzymatically inactive but immunologically cross-reactive material (CRM) in atypical oriental livers. ALDH1 and ALDH2 isozymes were purified to homogeneity from usual livers, and ALDH1 and CRM were purified from atypical oriental livers. Amino acid compositions of ALDH1 and ALDH2 were similar to, but not identical with, each other. Amino acid compositions of ALDH2 and CRM were identical within analytical errors. Subunit molecular size of ALDH1 estimated by sodium dodecyl sulfate-polyacrylamide gel electrophoresis was 56,200 daltons, and that of ALDH2 was 52,600 daltons. The two isozymes did not contain a common subunit. Subunit molecular weight of CRM was identical with that of ALDH2. Double immunodiffusion precipitation revealed that ALDH1 and ALDH2 were immunologically analogous but not identical, and that CRM and ALDH2 were immunologically indistinguishable. These results support the genetic model that CRM is an abnormal defective protein resulting from a mutation of the ALDH2 locus.  相似文献   

5.
V A Rizzoli  C R Rossi 《Enzyme》1988,39(1):28-43
In intact rat liver mitochondria acetaldehyde is oxidized by three functionally distinct dehydrogenase systems. Two of these reduce intramitochondrial nicotinamide adenine dinucleotide (NAD): one is operative with micromolar acetaldehyde concentrations and is stimulated by Mg2+, the other is operative with millimolar acetaldehyde concentrations and is stimulated by adenosine 5'-triphosphate (ATP). The third system reduces added NAD and is stimulated by rotenone. Connected to these systems, three aldehyde dehydrogenase isozymes (ALDH) have been purified: a low-Km ALDH activated by Mg2+, a high-Km ALDH activated by ATP and Mg2+, a high-Km ALDH activated by rotenone. The properties of some isozymes are affected by detergents. Thus, deoxycholate augments the stimulation of low-Km isozyme by Mg2+ and confers sensitivity to Mg2+ and ATP on one of the high-Km isozymes. A fourth isozyme has been purified. Its affinity for acetaldehyde is so low that it is very unlikely that acetaldehyde is the physiological substrate.  相似文献   

6.
Crystallographic analysis revealed that the nicotinamide ring of NAD can bind with multiconformations to aldehyde dehydrogenase (ALDH) (Ni, L., Zhou, J., Hurley, T. D., and Weiner, H. (1999) Protein Sci. 8, 2784-2790). Electron densities can be defined for two conformations, neither of which appears to be compatible with the catalytic reaction. In one conformation, it would prevent glutamate 268 from functioning as a general base needed to activate the catalytic nucleophile, cysteine 302. In the other conformation, the nicotinamide is too far from the enzyme-substrate adduct for efficient hydride transfer. In this study, NMR and fluorescence spectroscopies were used to demonstrate that NAD and NADH bind to human liver cytosol and mitochondrial ALDH such that the nicotinamide samples a population of conformations while the adenosine region remains relatively immobile. Although the nicotinamide possesses extensive conformational heterogeneity, the catalyzed reaction leads to the stereospecific transfer of hydride to the coenzyme. Mobility allows the nicotinamide to move into position to be reduced by the enzyme-substrate adduct. Although the reduced nicotinamide ring retains mobility after NADH formation, the extent of the motion is less than that of NAD. It appears that after reduction the population of favored nicotinamide conformations shifts toward those that do not interfere with the ability of the enzyme to release the reaction product. In the case of the mitochondrial, but not the cytosolic, enzyme this change in conformational preference is promoted by the presence of Mg2+ ions. Coenzyme conformational mobility appears to be beneficial to catalysis by ALDH throughout the catalytic cycle.  相似文献   

7.
Timed assays in which GTP and GDP were separated and quantitated by HPLC were developed and used to study the metal activation of the mitochondrial and cytosolic isozymes of phosphoenolpyruvate carboxykinase purified from rabbit liver. These assays allowed both directions of catalysis to be studied under similar conditions and in the absence of coupling enzymes. The mitochondrial enzyme is rapidly inactivated by preincubation with Fe2+, as had been shown previously for the cytosolic isozyme. The greatest activation by Fe2+ was obtained by adding micromolar Fe2+ immediately after enzyme to form the complete assay mixture that also contained millimolar Mg2+. In the direction of synthesis of OAA from Pep, the K0.5 values for Mn2+ and Fe2+ were in the 3-7 microM range when a nonchelating buffer, Hepes, was used. The buffer used strongly affected activation by Fe2+ at pH 7.4; activation was eliminated in the case of phosphate and K0.5 increased several-fold over that obtained with Hepes when imidazole was used. In non-chelating buffer, the pH optimum was near 7.4 for both isozymes and for both directions of catalysis. However, the near optimal pH range extended below 7.4 for the direction of oxaloacetate synthesis while the range was above 7.4 for Pep synthesis. In the direction of oxaloacetate synthesis: (1) Both isozymes required the presence of micromolar Mn2+ or Fe2+ in addition to millimolar Mg2+ in order to shown significant activity. (2) Fe2+ was as effective an activator as Mn2+ at pH 7 and below. In the direction of Pep synthesis: (1) Micromolar Mn2+ was a much better activator than Fe2+ at the higher pH values needed for optimal activity in this direction. (2) With increasing pH, decreasing activation was obtained with Fe2+ while the activity supported by Mg2+ alone increased. The results demonstrate the potential for regulation of either isozyme of Pep carboxykinase by the availability of iron or manganese.  相似文献   

8.
The binding of NADH and NAD+ to the human liver cytoplasmic, E1, and mitochondrial, E2, isozymes at pH 7.0 and 25 degrees C was studied by the NADH fluorescence enhancement technique, the sedimentation technique, and steady-state kinetics. The binding of radiolabeled [14C]NADH and [14C]NAD+ to the E1 isozyme when measured by the sedimentation technique yielded linear Scatchard plots with a dissociation constant of 17.6 microM for NADH and 21.4 microM for NAD+ and a stoichiometry of ca. two coenzyme molecules bound per enzyme tetramer. The dissociation constant, 19.2 microM, for NADH as competitive inhibitor was found from steady-state kinetics. With the mitochondrial E2 isozyme, the NADH fluorescence enhancement technique showed only one, high-affinity binding site (KD = 0.5 microM). When the sedimentation technique and radiolabeled coenzymes were used, the binding studies showed nonlinear Scatchard plots. A minimum of two binding sites with lower affinity was indicated for NADH (KD = 3-6 microM and KD = 25-30 microM) and also for NAD+ (KD = 5-7 microM and KD = 15-30 microM). A fourth binding site with the lowest affinity (KD = 184 microM for NADH and KD = 102 microM for NAD+) was observed from the steady-state kinetics. The dissociation constant for NAD+, determined by the competition with NADH via fluorescence titration, was found to be 116 microM. The number of binding sites found by the fluorescence titration (n = 1 for NADH) differs from that found by the sedimentation technique (n = 1.8-2.2 for NADH and n = 1.2-1.6 for NAD+).(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

9.
The variation with pH of the kinetic parameters for the alcohol and acetaldehyde reactions were studied for the alleloenzyme AdhS from Drosophila melanogaster. The variation of Ki (KEO,I) with pH for two ethanol-competitive inhibitors, pyrazole and 2,2,2-trifluoroethanol, was also studied. Both alcohol oxidation and acetaldehyde reduction follow a compulsory ordered pathway, with coenzyme binding first. The rate-limiting step for ethanol oxidation is complex and involves at least hydride transfer and dissociation of the enzyme-NADH complex (ER). In contrast with this, the rate-limiting step for the back reaction, i.e. the reduction of acetaldehyde, is dissociation of the enzyme-NAD+ complex (EO). A rate-limiting ER dissociation appears in the oxidation of the secondary alcohol propan-2-ol, whereas for the back reaction, i.e. acetone reduction, hydride transfer in the ternary complexes is rate-limiting. There is one group in the free enzyme, with a pK of approx. 8.0, that regulates the kon velocity for NADH, whereas for NAD+ several groups seem to be involved. A group in the enzyme is drastically perturbed by the formation of the binary EO complex. Protonation of this group with a pK of 7.6 in the EO complex resulted in weakened alcohol and inhibitor binding, in addition to an increased dissociation rate of NAD+ from the binary EO complex. Neither the binding of acetaldehyde nor the dissociation rate of NADH from the binary ER complex varied within the pH region studied.  相似文献   

10.
Polymorphism of aldehyde dehydrogenase and alcohol sensitivity   总被引:6,自引:0,他引:6  
H W Goedde  D P Agarwal 《Enzyme》1987,37(1-2):29-44
The metabolism of acetaldehyde has received considerable attention in the past years owing to its acute and chronic toxic effects in humans. Aldehyde dehydrogenase (ALDH) catalyzes the oxidation of acetaldehyde in liver and other organs. Two major isozymes of hepatic ALDH (ALDH I or E2 and ALDH II or E1), which differ in their structural and functional properties, have been characterized in humans. The ALDH I with a low Km for acetaldehyde is predominantly of mitochondrial origin and ALDH II which has a relatively higher Km is of cytosolic origin. An inherited deficiency of ALDH I isozyme has been found among Japanese and Chinese which is primarily responsible for producing acute alcohol sensitivity symptoms (flushing response) after drinking mild doses of alcohol. Biochemical, immunochemical and molecular genetics data indicate a structural mutation in the ALDH I isozyme gene responsible for the loss in catalytic activity. Population genetic studies indicate a wide prevalence of this ALDH polymorphism among individuals of Mongoloid race. Flushing response to alcohol shows familial resemblances and preliminary family data from Japan, China and Korea hint to an autosomal codominant inheritance for ALDH I isozyme deficiency. The ALDH polymorphism is apparently responsible for the low incidence of alcoholism in Japanese, Chinese and Koreans. Alcohol-induced sensitivity due to ALDH isozyme deficiency may act as an inhibitory factor against excessive alcohol drinking thereby imparting a protection against alcoholism.  相似文献   

11.
M Slatner  B Nidetzky  K D Kulbe 《Biochemistry》1999,38(32):10489-10498
To characterize catalysis by NAD-dependent long-chain mannitol 2-dehydrogenases (MDHs), the recombinant wild-type MDH from Pseudomonas fluorescens was overexpressed in Escherichia coli and purified. The enzyme is a functional monomer of 54 kDa, which does not contain Zn(2+) and has B-type stereospecificity with respect to hydride transfer from NADH. Analysis of initial velocity patterns together with product and substrate inhibition patterns and comparison of primary deuterium isotope effects on the apparent kinetic parameters, (D)k(cat), (D)(k(cat)/K(NADH)), and (D)(k(cat)/K(fructose)), show that MDH has an ordered kinetic mechanism at pH 8.2 in which NADH adds before D-fructose, and D-mannitol and NAD are released in that order. Isomerization of E-NAD to a form which interacts with D-mannitol nonproductively or dissociation of NAD from the binary complex after isomerization is the slowest step (>/=110 s(-)(1)) in D-fructose reduction at pH 8.2. Release of NADH from E-NADH (32 s(-)(1)) is the major rate-limiting step in mannitol oxidation at this pH. At the pH optimum for D-fructose reduction (pH 7.0), the rate of hydride transfer contributes significantly to rate limitation of the catalytic cascade and the overall reaction. (D)(k(cat)/K(fructose)) decreases from 2.57 at pH 7.0 to a value of 相似文献   

12.
The dehydrogenase activity of the mitochondrial isozyme (E2) of human liver aldehyde dehydrogenase was stimulated about 2-fold by the presence of low concentrations (about 120-140 microM) of Mg2+ in the assay at pH 7.0 using propionaldehyde as substrate. The stimulation was totally reversible by treatment with EDTA. Maximum stimulation was dependent on the concentration of NAD+ used in the assay; an increase in Km value of NAD+ was observed to parallel the increase in maximal velocity with increasing Mg2+ concentration, indicating that alterations in the catalytic properties of the E2 isozyme occur in the presence of Mg2+. The presteady state burst of NADH product was observed to decrease in the presence of Mg2+, suggesting that the rate-limiting step of the dehydrogenase reaction is altered by Mg2+. No evidence for Mg2+-induced alterations in the molecular weight properties of the E2 isozyme was observed using gel filtration column chromatography and fluorescence polarization techniques. In addition, no alterations in the inactivating properties of iodoacetamide or disulfiram were produced by Mg2+. These results suggest that the mechanism by which human mitochondrial aldehyde dehydrogenase (E2) is stimulated by Mg2+ is different from that of the horse enzyme, representing a significant species difference.  相似文献   

13.
Transient-kinetic studies of pig muscle lactate dehydrogenase   总被引:9,自引:8,他引:1  
1. The very fast pre-steady-state formation of NADH catalysed by pig M(4) lactate dehydrogenase was equivalent to the enzyme-site concentration at pH values greater than 8.0 and to one-half the site concentration at pH6.8. 2. The rate of dissociation of NADH from the enzyme at pH8.0 (450s(-1)) in the absence of other substrates is faster than the steady-state oxidation of lactate (80s(-1)). The latter process is therefore controlled by a step before NADH dissociation but subsequent to the hydride transfer. 3. The oxidation of enzyme-NADH by excess of pyruvate was studied as a first-order process at pH9.0. There was no effect of NADD on this reaction and it was concluded that the ternary complex undergoes a rate-limiting change before the hydride-transfer step. 4. Some conclusions about the reactions catalysed by the M(4) isoenzyme were drawn from a comparison of these results with those obtained with the H(4) isoenzyme and liver alcohol dehydrogenase.  相似文献   

14.
1. Lactate oxidation catalysed by pig heart lactate dehydrogenase was studied in the presence of inhibitory concentrations of pyruvate. Experimental results show the presence of an intermediate which occurs immediately after the hydride transfer step, but before the dissociation of pyruvate and the H+ produced by the reaction. The rate constant for pyruvate dissociation and the dissociation constant for pyruvate from the ternary complex differ from those obtained in pyruvate reduction experiments. 2.In single-turnover pyruvate reduction by pig heart lactate dehydrogenase at pH8.0 pyruvate can bind to the enzyme before a H+ is taken up, and the subsequent uptake of a H+ is governed by a step that is also rate-limiting for single-turnover and steady-state NADH oxidation. 3. Observation of various intermediates in the single-turnover pyruvate reduction experiments has made it possible to determine separately the dissociation constant and Km value for pyruvate at pH8.0, and also the catalytic turnover rate and Km for pyruvate under first-order conditions at different pH values. 4. Further studies on single-turnover pyruvate reduction carried out in 2H2O, or in water at low temperature, show another step which, under these conditions, is slower than that controlling H+ uptake and rate-limiting for NADH oxidation. A scheme is presented which explains these results.  相似文献   

15.
4-trans-(N,N-dimethylamino)cinnamaldehyde (DACA) is a chromophoric and fluorogenic substrate of aldehyde dehydrogenase. Fluorescence of DACA is enhanced by binding to aldehyde dehydrogenase in the absence of catalysis both in the presence and absence of the coenzyme analogue 5'AMP. DACA binds to aldehyde dehydrogenase with a dissociation constant of 1-3 microM and stoichiometry of 2 mol mol(-1) enzyme. Incorporation of DACA during catalysis was also investigated and found to be 2 mol DACA mol(-1) enzyme. Effect of pH on the stoichiometry of DACA incorporation during catalysis has shown that DACA incorporation remained constant at 2 mol DACA mol(-1) enzyme, despite a 74-fold velocity enhancement between pH 5.0 and 9.0. Increase of pH increased decomposition of enzyme-acyl intermediate without affecting the rate-limiting step of the reaction. At pH 7.0 the pH stimulated velocity enhancement was 10-fold over that at pH 5.0; further velocity enhancement (11.5-fold that of pH 7.0) was achieved by 150 microM Mg(2+) ions. The velocity at pH 7.0 with Mg(2+) exceeded that of pH 9.0, and that at maximal pH stimulation at pH 9.5. It was observed that level of intermediate decreased to about 1 mol mol(-1) enzyme, indicating that Mg(2+) ions increased the rate of decomposition of the enzyme-acyl intermediate and shifted the rate-limiting step of the reaction to another step in the reaction sequence.  相似文献   

16.
The synthesis of glutamic oxaloacetic transaminase isozymes in rat liver explants was studied using specific antisera against the cytosolic and mitochondrial isozymes. The pulse-labeled cytosolic isozyme was detected in the cytosolic fraction and remained there in pulse-chase experiments. On the other hand, the pulse-labeled mitochondrial isozyme was detected as a larger precursor in the cytosolic fraction. During chase, the amount of pulse-labeled precursor of the mitochondrial isozyme decreased and labeled mature mitochondrial isozyme appeared in the mitochondrial fraction.  相似文献   

17.
Oral administration of 7.0 mg/kg calcium carbimide (calcium cyanamide, CC) to the rat produced differential inhibition of hepatic aldehyde dehydrogenase (ALDH) isozymes, as indicated by the time-course profiles of enzyme activity. The low-Km mitochondrial ALDH was most susceptible to inhibition following CC administration, with complete inhibition occurring at 0.5 h and return to control activity at 96 h. The low-Km cytosolic and high-Km mitochondrial, cytosolic, and microsomal ALDH isozymes were inhibited to a lesser degree and (or) for a shorter duration compared with the mitochondrial low-Km enzyme. The time course of carbimide, the hydrolytic product of CC, was determined in plasma following oral administration of 7.0 mg/kg CC to the rat. The maximum plasma carbimide concentration (102 ng/mL) occurred at 1 h and the apparent elimination half-life in plasma was 1.5 h. Carbimide was not measurable in the liver during the 6.5 h time interval when carbimide was present in the plasma. There were negative, linear correlations between plasma carbimide concentration and hepatic low-Km mitochondrial, low-Km cytosolic, and high-Km microsomal ALDH activities. In vitro studies demonstrated that carbimide, at concentrations obtained in plasma following oral CC administration, produced only 19% inhibition of low-Km mitochondrial ALDH and no inhibition of low-Km cytosolic and high-Km microsomal ALDH isozymes. These data demonstrate that carbimide, itself, is not primarily responsible for hepatic ALDH inhibition in vivo following oral CC administration. It would appear that carbimide must undergo metabolic conversion in vivo to inhibit hepatic ALDH enzymes, which is supported by the observation of no measurable carbimide in the liver when ALDH was maximally inhibited following oral CC administration.  相似文献   

18.
The kcat value for the oxidation of propionaldehyde by sheep liver cytosolic aldehyde dehydrogenase increased 3-fold, from 0.16 s-1 at pH 7.6 to 0.49 s-1 at pH 5.2, in parallel with the increase in the rate of displacement of NADH from binary enzyme.NADH complexes. A burst in nucleotide fluorescence was observed at all pH values consistent with the rate of isomerization of binary enzyme.NADH complexes constituting the rate-limiting step in the steady state. No substrate activation by propionaldehyde was observed at pH 5.2, but the enzyme exhibited dissociation/association behavior. The inactive dissociated form of the enzyme was favored by low enzyme concentration, low pH, and low ionic strength. Propionaldehyde protected the enzyme against dissociation.  相似文献   

19.
Acetaldehyde and biogenic aldehydes were used as substrates to investigate the subcellular distribution of aldehyde dehydrogenase activity in autopsied human brain. With 10 microM acetaldehyde as substrate, over 50% of the total activity was found in the mitochondrial fraction and 38% was associated with the cytosol. However, with 4 microM 3,4-dihydroxyphenylacetaldehyde and 10 microM indoleacetaldehyde as substrates, 40-50% of the total activity was found in the soluble fraction, the mitochondrial fraction accounting for only 15-30% of the total activity. These data suggested the presence of distinct aldehyde dehydrogenase isozymes in the different compartments. The mitochondrial and cytosolic fractions were, therefore, subjected to salt fractionation and ion-exchange chromatography to purify further the isozymes present in both fractions. The kinetic data on the partially purified isozymes revealed the presence of a low Km isozyme in both the mitochondria and the cytosol, with Km values for acetaldehyde of 1.7 microM and 10.2 microM, respectively. However, the cytosolic isozyme exhibited lower Km values for the biogenic aldehydes. Both isozymes were activated by Mg2+ and Ca2+ in phosphate buffers (pH 7.4). Also, high Km isozymes were found in the mitochondria and in the microsomes.  相似文献   

20.
The role of Ca(2+) as a cytosolic signaling molecule between porcine cardiac sarcoplasmic reticulum (SR) ATPase and mitochondrial ATP production was evaluated in vitro. The Ca(2+) sensitivity of these processes was determined individually and in a reconstituted system with SR and mitochondria in a 0.5:1 protein-to-cytochrome aa(3) ratio. The half-maximal concentration (K(1/2)) of SR ATPase was 335 nM Ca(2+). The ATP synthesis dependence was similar with a K(1/2) of 243 nM for dehydrogenases and 114 nM for overall ATP production. In the reconstituted system, Ca(2+) increased thapsigargin-sensitive ATP production (maximum approximately 5-fold) with minimal changes in mitochondrial reduced nicotinamide adenine dinucleotide (NADH). NADH concentration remained stable despite graded increases in NADH turnover induced over a wide range of Ca(2+) concentrations (0 to approximately 500 nM). These data are consistent with a balanced activation of SR ATPase and mitochondrial ATP synthesis by Ca(2+) that contributes to a homeostasis of energy metabolism metabolites. It is suggested that this balanced activation by cytosolic Ca(2+) is partially responsible for the minimal alteration in energy metabolism intermediates that occurs with changes in cardiac workload in vivo.  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