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.
1. The reversible NAD(+)-linked oxidation of d-3-hydroxybutyrate to acetoacetate in 0.1m-sodium pyrophosphate buffer, pH8.5, at 25.0 degrees C, catalysed by d-3-hydroxybutyrate dehydrogenase (d-3-hydroxybutyrate-NAD(+) oxidoreductase, EC 1.1.1.30), was studied by initial-velocity, dead-end inhibition and product-inhibition analysis. 2. The reactions were carried out on (a) the soluble enzyme from Rhodopseudomonas spheroides and (b) an insoluble derivative of this enzyme prepared by its covalent attachment to DEAE-cellulose by using 2-amino-4,6-dichloro-s-triazine as coupling agent. 3. The insolubilized enzyme preparation contained 5mg of protein/g wet wt. of total material, and when freshly prepared its specific activity was 1.2mumol/min per mg of protein, which is 67% of that of the soluble dialysed enzyme. 4. The reactions catalysed by both the enzyme in solution and the insolubilized enzyme were shown to follow sequential pathways in which the nicotinamide nucleotides bind obligatorily first to the enzyme. Evidence is presented for kinetically significant ternary complexes and that the rate-limiting step(s) of both catalyses probably involves isomerization of the enzyme-nicotinamide nucleotide complexes and/or dissociation of the nicotinamide nucleotides from the enzyme. Both catalyses therefore are probably best described as ordered Bi Bi mechanisms, possibly with multiple enzyme-nicotinamide nucleotide complexes. 5. The kinetic parameters and the calculable rate constants for the catalysis by the soluble enzyme are similar to the corresponding parameters and rate constants for the catalysis by the insolubilized enzyme.  相似文献   

2.
1. The existing procedures for extraction of oxidized and reduced nicotinamide coenzymes were adapted to spermatozoa to overcome the coenzyme-degrading activity of seminal plasma. 2. The content of total NAD(+) and NADH was determined in the spermatozoa of ram, bull, boar, stallion and cock. NADP(+) and NADPH were not detected in ram spermatozoa. 3. The oxidation state of sperm NAD depended on the seminal plasma, the removal of which produced a change in the percentage oxidation state of the coenzyme, 100x[NAD(+)/(NAD(+)+NADH)], without altering the total content of NAD(+)+NADH. 4. In suspensions of washed ram spermatozoa, incubated anaerobically at 25 degrees C, the percentage oxidation state of NAD declined with increasing spermatozoa concentration. 5. When ram or boar spermatozoa that had been previously washed and resuspended in Ringer phosphate medium, were incubated anaerobically at 25 degrees C with various substances, pronounced effects on the percentage oxidation state of NAD could be observed with l-lactate, pyruvate, oxaloacetate, dihydroxyacetone, formaldehyde and glyceraldehyde; sorbitol and acetoacetate acted only on ram spermatozoa; fructose, glucose, mannose and acetaldehyde acted predominantly on boar spermatozoa. Formaldehyde lowered the (NAD(+)+NADH) content of ram spermatozoa, but none of the other substances had a comparable effect. 6. The percentage oxidation state of sperm NAD was not influenced by exogenous cysteine, cystine, ergothioneine or ascorbate. 7. A highly active sorbitol dehydrogenase could be prepared from ram, but not from boar, spermatozoa. 8. Sorbitol, acetoacetate and 3-hydroxybutyrate effectively supported the respiration of ram, but not boar, spermatozoa. 9. ;Cold shock', resulting from sudden cooling of spermatozoa, abolished motility completely and irreversibly but produced only a slow and partial decrease in the total NAD content. Slight over-heating, sufficient to produce loss of motility, had no adverse effect on the total NAD content. 10. Storage of ram sperm at 14 degrees C produced only a small decrease of NAD after 2 days, but subsequently the loss became greater.  相似文献   

3.
Effects of ischaemia on metabolite concentrations in rat liver   总被引:24,自引:21,他引:3       下载免费PDF全文
1. Changes in the concentrations of ammonia, glutamine, glutamate, 2-oxoglutarate, 3-hydroxybutyrate, acetoacetate, alanine, aspartate, malate, lactate, pyruvate, NAD(+), NADH and adenine nucleotides were measured in freeze-clamped rat liver during ischaemia. 2. Although the concentrations of most of the metabolites changed rapidly during ischaemia the ratios [glutamate]/[2-oxoglutarate][NH(4) (+)] and [3-hydroxybutyrate]/[acetoacetate] changed equally and the value of the expression [3-hydroxybutyrate][2-oxoglutarate][NH(4) (+)]/[acetoacetate][glutamate] remained approximately constant, indicating that the 3-hydroxybutyrate dehydrogenase and glutamate dehydrogenase systems were at near-equilibrium with the mitochondrial NAD(+) couple. 3. The value of the expression [alanine][oxoglutarate]/[pyruvate][glutamate] was about 0.7 in vivo and remained fairly constant during the ischaemic period of 5min, although the concentrations of alanine and oxoglutarate changed substantially. No explanation can be offered why the value of the ratio differed from that of the equilibrium constant of the alanine aminotransferase reaction, which is 1.48. 4. Injection of l-cycloserine 60min before the rats were killed increased the concentration of alanine in the liver fourfold and decreased the concentration of the other metabolites measured, except that of pyruvate. During ischaemia the concentration of alanine did not change but that of aspartate almost doubled. 5. After treatment with l-cycloserine the value in vivo of the expression [alanine][oxoglutarate]/[pyruvate][glutamate] rose from 0.7 to 2.4. During ischaemia the value returned to 0.8. 6. The effects of l-cycloserine are consistent with the assumption that it specifically inhibits alanine aminotransferase. 7. Most of the alanine formed during ischaemia is probably derived from pyruvate and from ammonia released by the deamination of adenine nucleotides and glutamine. The alanine is presumably formed by the combined action of glutamate dehydrogenase and alanine aminotransferase. 8. The rate of anaerobic glycolysis, calculated from the increase in the lactate concentration, was 1.3mumol/min per g fresh wt. 9. Although the concentrations of the adenine nucleotides changed rapidly during ischaemia, the ratio [ATP][AMP]/[ADP](2) remained constant at 0.54, indicating that adenylate kinase established near-equilibrium under these conditions.  相似文献   

4.
Abstract Poly-β-hydroxybutyrate biosynthesis was studied in Alcaligenes eutrophus under various nutrient-limiting conditions. When the cells were cultivated in nitrogen-limited media, both the levels of NAD(P)H and the ratios of NAD(P)H/NAD(P) were higher than those under nitrogen-sufficient conditions. The specific poly-β-hydroxybutyrate production rate was found to increase with the values of both NADH/NAD and NADPH/NADP, indicating that poly-β-hydroxybutyrate synthesis is directly regulated by the ratios of nicotinamide nucleotides. The effects of nicotinamide nucleotides on poly-β-hydroxybutyrate biosynthesis was investigated with regard to enzyme kinetics. Citrate synthase activity was significantly inhibited by NADH and NADPH, indicating that poly-β-hydroxybutyrate accumulation could be enhanced by facilitating the metabolic flux of acetyl-CoA to poly-β-hydroxybutyrate synthetic pathway. It was also found that cellular NADPH was a limiting substrate for NADPH-linked reductase, controlling the overall biosynthetic activity of poly-/3-hydroxybutyrate in this strain.  相似文献   

5.
Nidetzky B  Klimacek M  Mayr P 《Biochemistry》2001,40(34):10371-10381
Microbial xylose reductase, a representative aldo-keto reductase of primary sugar metabolism, catalyzes the NAD(P)H-dependent reduction of D-xylose with a turnover number approximately 100 times that of human aldose reductase for the same reaction. To determine the mechanistic basis for that physiologically relevant difference and pinpoint features that are unique to the microbial enzyme among other aldo/keto reductases, we carried out stopped-flow studies with wild-type xylose reductase from the yeast Candida tenuis. Analysis of transient kinetic data for binding of NAD(+) and NADH, and reduction of D-xylose and oxidation of xylitol at pH 7.0 and 25 degrees C provided estimates of rate constants for the following mechanism: E + NADH right arrow over left arrow E.NADH right arrow over left arrow E.NADH + D-xylose right arrow over left arrow E.NADH.D-xylose right arrow over left arrow E.NAD(+).xylitol right arrow over left arrow E.NAD(+) right arrow over left arrow E.NAD(+) right arrow over left arrow E + NAD(+). The net rate constant of dissociation of NAD(+) is approximately 90% rate limiting for k(cat) of D-xylose reduction. It is controlled by the conformational change which precedes nucleotide release and whose rate constant of 40 s(-)(1) is 200 times that of completely rate-limiting E.NADP(+) --> E.NADP(+) step in aldehyde reduction catalyzed by human aldose reductase [Grimshaw, C. E., et al. (1995) Biochemistry 34, 14356-14365]. Hydride transfer from NADH occurs with a rate constant of approximately 170 s(-1). In reverse reaction, the E.NADH --> E.NADH step takes place with a rate constant of 15 s(-1), and the rate constant of ternary-complex interconversion (3.8 s(-1)) largely determines xylitol turnover (0.9 s(-1)). The bound-state equilibrium constant for C. tenuis xylose reductase is estimated to be approximately 45 (=170/3.8), thus greatly favoring aldehyde reduction. Formation of productive complexes, E.NAD(+) and E.NADH, leads to a 7- and 9-fold decrease of dissociation constants of initial binary complexes, respectively, demonstrating that 12-fold differential binding of NADH (K(i) = 16 microM) vs NAD(+) (K(i) = 195 microM) chiefly reflects difference in stabilities of E.NADH and E.NAD(+). Primary deuterium isotope effects on k(cat) and k(cat)/K(xylose) were, respectively, 1.55 +/- 0.09 and 2.09 +/- 0.31 in H(2)O, and 1.26 +/- 0.06 and 1.58 +/- 0.17 in D(2)O. No deuterium solvent isotope effect on k(cat)/K(xylose) was observed. When deuteration of coenzyme selectively slowed the hydride transfer step, (D)()2(O)(k(cat)/K(xylose)) was inverse (0.89 +/- 0.14). The isotope effect data suggest a chemical mechanism of carbonyl reduction by xylose reductase in which transfer of hydride ion is a partially rate-limiting step and precedes the proton-transfer step.  相似文献   

6.
(1) d(-)-3-Hydroxybutyrate dehydrogenase activity from guinea pig, rat, and bovine heart and from guinea pig liver is inhibited by malonate and tartronate, and more potently by the analogs methylmalonate, bromomalonate, chloromalonate, and mesoxalate. Little or no inhibitory effect was found for aminomalonate, ethylmalonate, dimethylmalonate, succinate, glutarate, oxaloacetate, malate, propionate, pyruvate, d- and l-lactate, n-butyrate, isobutyrate, and cyclopropanecarboxylate. (2) In initial velocity kinetics at pH 8.1 with a soluble enzyme preparation from bovine heart, the inhibition by the active malonate derivatives is competitive with respect to 3-hydroxybutyrate and uncompetitive with respect to acetoacetate, NAD+ or NADH. With d-3-hydroxybutyrate as the variable reactant (Km app = 0.26 mM) the inhibition constant of methylmalonate (Kis) was 0.09 mm. (3) The rate of utilization of d-3-hydroxybutyrate (78 μm) by coupled rat heart mitochondria in the presence of ADP was inhibited 50% by 150 μm methylmalonate. (4) With coupled guinea pig liver mitochondria oxidizing n-octanoate in the absence of added ADP, methylmalonate (1–3 mm) depressed 3-hydroxybutyrate formation substantially more than total ketone production. However, the intramitochondrial NADH (or NADPH) levels were unchanged by the addition of methylmalonate, indicating that the changes in ratios of accumulated 3-hydroxybutyrate and acetoacetate were caused by direct inhibition of 3-hydroxybutyrate dehydrogenase. Methylmalonate had the same effect on 3-hydroxybutyrate/acetoacetate ratios and ketone body formation with pyruvate or acetate as the source of acetyl groups. Similar results were obtained with malonate (10 mm) although the inhibition of total ketone formation from octanoate was more severe.  相似文献   

7.
The content of coenzyme A-SH (CoASH) and acetyl-CoA of suspensions of rat heart mitochondria was stabilized by the addition of DL-carnitine and acetyl-DL-carnitine, in the presence of the respiratory inhibitor rotenone. The mitochondrial content of NAD+ and NADH was similarly stabilized by the addition of acetoacetate and DL-3-hydroxybutyrate, and the content of ADP and ATP was imposed by the addition of these nucleotides to the mitochondrial suspension, in the presence of uncoupling agent and oligomycin, to inhibit ATPase. Under these conditions, mitochondrial CoASH/acetyl-CoA, NAD+/ NADH, and ADP/ATP ratios could be varied independently, and the effect on the interconversion of active and inactive pyruvate dehydrogenase could be studied. Decreases in both CoASH/acetyl-CoA and NAD+/NADH ratios were shown to be inhibitory to the steady state activity of pyruvate dehydrogenase, and this effect is described at three different ADP/ATP ratios and different concentrations of added MgCl2. A new steady state level of activity was achieved within 10 min of a change in either CoASH/acetyl-CoA or NAD+/NADH ratio; the rate of inactivation was much higher than the rate of reactivation under these conditions. Effects of CoASH/acetyl-CoA and NAD+/NADH may be additive but are still quantitatively lesser than the changes in activity of pyruvate dehydrogenase induced by changes in ADP/ATP ratio. The variation in activity of pyruvate dehydrogenase with ADP/ATP ratio is described in the absence of changes in the other two ratios, conditions which were not met in earlier studies which employed the oxidation of different substrates to generate changes in all three ratios.  相似文献   

8.
1. The inhibition of alkaline phosphatase by NAD(+), NADH, adenosine and nicotinamide was studied. 2. All of these substances except NAD(+) act as uncompetitive inhibitors, i.e. double-reciprocal plots are parallel. NAD(+), however, is a ;mixed' inhibitor of alkaline phosphatase and is less potent than NADH. 3. Inhibition studies with pairs of the inhibitors suggest that, in spite of the difference in type of inhibition, NAD(+) and NADH bind to alkaline phosphatase at a common site. Adenosine and nicotinamide also seem to bind at the NAD site and the binding of adenosine is facilitated by nicotinamide, and vice versa. 4. The facilitation may indicate the occurrence of an induced fit for NAD(+) and NADH. Attempts to desensitize alkaline phosphatase to NAD(+) and NADH inhibition by partial denaturation were unsuccessful. 5. The results are discussed in terms of a two-site model in which separate, but interacting, regions exist on the enzyme to accommodate the adenosine and nicotinamide moieties of NAD, and a single-site model in which the adenosine part of the molecule is bound preferentially and this interacts with the nicotinamide fraction. 6. The activity of alkaline phosphatase can be changed fourfold by alteration of the NAD(+)/NADH ratio. This sensitivity to the redox state of the coenzyme could be a means of controlling phosphatase activity.  相似文献   

9.
1. Halothane at clinically effective concentrations [2.5 and 4% (v/v) of the gas phase of the incubation flask] was found to inhibit significantly lipogenesis from endogenous substrates, e.g., glycogen, or from added lactate plus pyruvate. This was accompanied by a decrease in the ratio of the free [NAD+]/[NADH] of the mitochondrion and the cytoplasm, as shown by the [3-hydroxybutyrate]/[acetoacetate] ratio and the [lactate]/[pyruvate] ratio. 2. Acetoacetate or pyruvate decreased the inhibitory effect of halothane and restored lipogenesis to control rates. They were reduced rapidly by 3-hydroxybutyrate dehydrogenase or lactate dehydrogenase respectively, with the concomitant oxidation of NADH and the generation of NAD+. 3. These results suggest that the mechanism by which halothane inhibits lipogenesis from glycogen or lactate is by inhibition of the oxidation of NADH; this results in inhibition of flux of carbon through pyruvate dehydrogenase and a shortage of acetyl-CoA for fatty acid synthesis. Thus when NADH acceptors are added in the presence of halothane, the concentration of mitochondrial NAD+ is raised so that the flux of carbon through pyruvate dehydrogenase increases and lipogenesis is restored.  相似文献   

10.
Burgos ES  Schramm VL 《Biochemistry》2008,47(42):11086-11096
Human nicotinamide phosphoribosyltransferase (NAMPT, EC 2.4.2.12) catalyzes the reversible synthesis of nicotinamide mononucleotide (NMN) and inorganic pyrophosphate (PP i) from nicotinamide (NAM) and alpha- d-5-phosphoribosyl-1-pyrophosphate (PRPP). NAMPT, by capturing the energy provided by its facultative ATPase activity, allows the production of NMN at product:substrate ratios thermodynamically forbidden in the absence of ATP. With ATP hydrolysis coupled to NMN synthesis, the catalytic efficiency of the system is improved 1100-fold, substrate affinity dramatically increases ( K m (NAM) from 855 to 5 nM), and the K eq shifts -2.1 kcal/mol toward NMN formation. ADP-ATP isotopic exchange experiments support the formation of a high-energy phosphorylated intermediate (phospho-H247) as the mechanism for altered catalytic efficiency during ATP hydrolysis. NAMPT captures only a small portion of the energy generated by ATP hydrolysis to shift the dynamic chemical equilibrium. Although the weak energetic coupling of ATP hydrolysis appears to be a nonoptimized enzymatic function, closer analysis of this remarkable protein reveals an enzyme designed to capture NAM with high efficiency at the expense of ATP hydrolysis. NMN is a rate-limiting precursor for recycling to the essential regulatory cofactor, nicotinamide adenine dinucleotide (NAD (+)). NMN synthesis by NAMPT is powerfully inhibited by both NAD (+) ( K i = 0.14 muM) and NADH ( K i = 0.22 muM), an apparent regulatory feedback mechanism.  相似文献   

11.
In an attempt to trace the source of phosphate activation of the enzyme-catalysed pyruvate-lactate interconversion by rabbit muscle lactate dehydrogenase, equilibrium constants were measured to examine the effects of phosphate on interactions pertinent to the enzymic process. Frontal gel-chromatographic studies of the binding of NADH to the enzyme established that the intrinsic association constant is doubled in the presence of 50 mM-phosphate in the buffer (pH 7.4, I0.15). From kinetic studies of the competition between NAD+ and NADH for the coenzyme-binding sites of the enzyme it is concluded that the binding of oxidized nicotinamide nucleotide is also doubled in the presence of 50 mM-phosphate. Competitive-inhibition studies and fluorescence-quenching measurements indicated the lack of a phosphate effect on ternary-complex formation between enzyme-NADH complex and oxamate, a substrate analogue of pyruvate. The equilibrium constant for the interaction between enzyme-NAD+ complex and oxalate, an analogue of lactate, was also shown, by difference spectroscopy, to be insensitive to phosphate concentration. Provided that the effects observed with the substrate analogues mimic those operative in the kinetic situation, the equilibrium constant governing the isomerization of ternary complex is also independent of phosphate concentration. It is concluded that enhanced coenzyme binding is the source of phosphate activation of the rabbit muscle lactate dehydrogenase system.  相似文献   

12.
The involvement of nicotinamide adenine nucleotides (NAD(+), NADH) in the regulation of glycolysis in Lactococcus lactis was investigated by using (13)C and (31)P NMR to monitor in vivo the kinetics of the pools of NAD(+), NADH, ATP, inorganic phosphate (P(i)), glycolytic intermediates, and end products derived from a pulse of glucose. Nicotinic acid specifically labeled on carbon 5 was synthesized and used in the growth medium as a precursor of pyridine nucleotides to allow for in vivo detection of (13)C-labeled NAD(+) and NADH. The capacity of L. lactis MG1363 to regenerate NAD(+) was manipulated either by turning on NADH oxidase activity or by knocking out the gene encoding lactate dehydrogenase (LDH). An LDH(-) deficient strain was constructed by double crossover. Upon supply of glucose, NAD(+) was constant and maximal (approximately 5 mm) in the parent strain (MG1363) but decreased abruptly in the LDH(-) strain both under aerobic and anaerobic conditions. NADH in MG1363 was always below the detection limit as long as glucose was available. The rate of glucose consumption under anaerobic conditions was 7-fold lower in the LDH(-) strain and NADH reached high levels (2.5 mm), reflecting severe limitation in regenerating NAD(+). However, under aerobic conditions the glycolytic flux was nearly as high as in MG1363 despite the accumulation of NADH up to 1.5 mm. Glyceraldehyde-3-phosphate dehydrogenase was able to support a high flux even in the presence of NADH concentrations much higher than those of the parent strain. We interpret the data as showing that the glycolytic flux in wild type L. lactis is not primarily controlled at the level of glyceraldehyde-3-phosphate dehydrogenase by NADH. The ATP/ADP/P(i) content could play an important role.  相似文献   

13.
1. The effects of injecting nicotinamide, 5-methylnicotinamide, ethionine, nicotinamide+5-methylnicotinamide and nicotinamide+ethionine on concentrations in rat liver of NAD, NADP and ATP were investigated up to 5hr. after injection. 2. Nicotinamide induced three- to four-fold increases in hepatic NAD concentration even in the presence of 5-methylnicotinamide or ethionine, whereas 5-methylnicotinamide or ethionine alone did not cause marked changes in hepatic NAD concentration. 3. Nicotinamide alone also induced a twofold increase in hepatic NADP concentration. However, in the presence of 5-methylnicotinamide+nicotinamide, the NADP concentration decreased by 25% after 5hr., and in the presence of nicotinamide+ethionine by 30% in the same time. In the presence of 5-methylnicotinamide or ethionine alone hepatic NADP concentrations fell by 50% after 5hr. 4. 5-Methylnicotinamide inhibited the microsomal NAD(+) glycohydrolase (EC 3.2.2.6) by 60% at a concentration of 1mm and the NADP(+) glycohydrolase by 40% at the same concentration. 5. The rat liver NAD(+) kinase (EC 2.7.1.23) was found to have V(max.) 4.83mumoles/g. wet wt./hr. and K(m) (NAD(+)) 5.8mm. This enzyme was also inhibited by 5-methylnicotinamide in a ;mixed' fashion. 6. The results are discussed with respect to the control of NAD synthesis. It is suggested that in vivo the NAD(P)(+) glycohydrolases are effectively inactive and that the increased NAD concentrations induced by nicotinamide are due to increased substrate concentration available to both the nicotinamide and nicotinic acid pathways of NAD formation.  相似文献   

14.
Pyridine Nucleotide Transhydrogenase from Azotobacter vinelandii   总被引:5,自引:0,他引:5       下载免费PDF全文
A method is described for the partial purification of pyridine nucleotide transhydrogenase from Azotobacter vinelandii (ATCC 9104) cells. The most highly purified preparation catalyzes the reduction of 300 mumoles of nicotinamide adenine dinucleotide (NAD(+)) per min per mg of protein under the assay conditions employed. The enzyme catalyzes the reduction of NAD(+), deamino-NAD(+), and thio-NAD(+) with reduced nicotinamide adenine dinucleotide phosphate (NADPH) as hydrogen donor, and the reduction of nicotinamide adenine dinucleotide phosphate (NADP(+)) and thio-NAD(+) with reduced NAD (NADH) as hydrogen donor. The reduction of acetylpyridine AD(+), pyridinealdehyde AD(+), acetylpyridine deamino AD(+), and pyridinealdehydedeamino AD(+) with NADPH as hydrogen donor was not catalyzed. The enzyme catalyzes the transfer of hydrogen more readily from NADPH than from NADH with different hydrogen acceptors. The transfer of hydrogen from NADH to NADP(+) and thio-NAD(+) was markedly stimulated by 2'-adenosine monophosphate (2'-AMP) and inhibited by adenosine diphosphate (ADP), adenosine triphosphate (ATP), and phosphate ions. The transfer of hydrogen from NADPH to NAD(+) was only slightly affected by phosphate ions and 2'-AMP, except at very high concentrations of the latter reagent. In addition, the transfer of hydrogen from NADPH to thio-NAD(+) was only slightly influenced by 2'-AMP, ADP, ATP, and other nucleotides. The kinetics of the transhydrogenase reactions which utilized thio-NAD(+) as hydrogen acceptor and NADH or NADPH as hydrogen donor were studied in some detail. The results suggest that there are distinct binding sites for NADH and NAD(+) and perhaps a third regulator site for NADP(+) or 2'-AMP. The heats of activation for the transhydrogenase reactions were determined. The properties of this enzyme are compared with those of other partially purified transhydrogenases with respect to the regulatory functions of 2'-AMP and other nucleotides on the direction of flow of hydrogen between NAD(+) and NADP(+).  相似文献   

15.
Transhydrogenase couples the redox reaction between NAD(H) and NADP(H) to proton translocation across a membrane. Coupling is achieved through changes in protein conformation. Upon mixing, the isolated nucleotide-binding components of transhydrogenase (dI, which binds NAD(H), and dIII, which binds NADP(H)) form a catalytic dI(2).dIII(1) complex, the structure of which was recently solved by x-ray crystallography. The fluorescence from an engineered Trp in dIII changes when bound NADP(+) is reduced. Using a continuous flow device, we have measured the Trp fluorescence change when dI(2).dIII(1) complexes catalyze reduction of NADP(+) by NADH on a sub-millisecond scale. At elevated NADH concentrations, the first-order rate constant of the reaction approaches 21,200 s(-1), which is larger than that measured for redox reactions of nicotinamide nucleotides in other, soluble enzymes. Rather high concentrations of NADH are required to saturate the reaction. The deuterium isotope effect is small. Comparison with the rate of the reverse reaction (oxidation of NADPH by NAD(+)) reveals that the equilibrium constant for the redox reaction on the complex is >36. This high value might be important in ensuring high turnover rates in the intact enzyme.  相似文献   

16.
Sir2 (silent information regulator 2) enzymes catalyze a unique protein deacetylation reaction that requires the coenzyme NAD(+) and produces nicotinamide and a newly discovered metabolite, O-acetyl-ADP-ribose (OAADPr). Conserved from bacteria to humans, these proteins are implicated in the control of gene silencing, metabolism, apoptosis, and aging. Here we examine the role of NAD(+) metabolites/derivatives and salvage pathway intermediates as activators, inhibitors, or coenzyme substrates of Sir2 enzymes in vitro. Also, we probe the coenzyme binding site using inhibitor binding studies and alternative coenzyme derivatives as substrates. Sir2 enzymes showed an exquisite selectivity for the nicotinamide base coenzyme, with the most dramatic losses in binding affinity/reactivity resulting from relatively minor changes in the nicotinamide ring, either by reduction, as in NADH, or by converting the amide to its acid analogue. Both ends of the dinucleotide NAD(+) are shown to be critical for high selectivity and high affinity. Among the NAD(+) metabolites tested none were able to allosterically activate, although all led to various extents of inhibition, consistent with competition at the coenzyme binding site. Nicotinamide was the most potent inhibitor examined, suggesting that cellular nicotinamide levels would provide an effective small molecule regulator of protein deacetylation and generation of OAADPr. The presented findings also suggest that changes in the physiological NAD(+):NADH ratio, without a change in NAD(+), would yield little alteration in Sir2 activity. That is, NADH is an extremely ineffective inhibitor of Sir2 enzymes (average IC(50) of 17 mm). We propose that changes in both free nicotinamide and free NAD(+) afford the greatest contribution to cellular activity of Sir2 enzymes but with nicotinamide having a more dramatic effect during smaller fluctuations in concentration.  相似文献   

17.
Kinetic studies of formate dehydrogenase   总被引:4,自引:1,他引:3       下载免费PDF全文
1. The kinetic mechanism of formate dehydrogenase is a sequential pathway. 2. The binding of the substrates proceeds in an obligatory order, NAD(+) binding first, followed by formate. 3. It seems most likely that the interconversion of the central ternary complex is extremely rapid, and that the rate-limiting step is the formation or possible isomerization of the enzyme-coenzyme complexes. 4. The secondary plots of the inhibitions with HCO(3) (-) and NO(3) (-) are non-linear, which suggests that more than one molecule of each species is able to bind to the same enzyme form. 5. The rate of the reverse reaction with carbon dioxide at pH6.0 is 20 times that with bicarbonate at pH8.0, although no product inhibition could be detected with carbon dioxide. The low rate of the reverse reaction precluded any steady-state analysis as the enzyme concentrations needed to obtain a measurable rate are of the same order as the K(m) values for NAD(+) and NADH.  相似文献   

18.
Inosine 5'-monophosphate dehydrogenase (IMPDH) catalyzes the oxidation of inosine 5'-monophosphate (IMP) to xanthosine 5'-monophosphate with the concomitant reduction of NAD to NADH. Escherichia coli IMPDH is activated by K(+), Rb(+), NH(+)(4), and Cs(+). K(+) activation is inhibited by Li(+), Na(+), Ca(2+), and Mg(2+). This inhibition is competitive versus K(+) at high K(+) concentrations, noncompetitive versus IMP, and competitive versus NAD. Thus monovalent cation activation is linked to the NAD site. K(+) increases the rate constant for the pre-steady-state burst of NADH production, possibly by increasing the affinity of NAD. Three mutant IMPDHs have been identified which increase the value of K(m) for K(+): Asp13Ala, Asp50Ala, and Glu469Ala. In contrast to wild type, both Asp13Ala and Glu469Ala are activated by all cations tested. Thus these mutations eliminate cation selectivity. Both Asp13 and Glu469 appear to interact with the K(+) binding site identified in Chinese hamster IMPDH. Like wild-type IMPDH, K(+) activation of Asp50Ala is inhibited by Li(+), Na(+), Ca(2+), and Mg(2+). However, this inhibition is noncompetitive with respect to K(+) and competitive with respect to both IMP and NAD. Asp50 interacts with residues that form a rigid wall in the IMP site; disruption of this wall would be expected to decrease IMP binding, and the defect could propagate to the proposed K(+) site. Alternatively, this mutation could uncover a second monovalent cation binding site.  相似文献   

19.
The mechanisms by which ethanol (EtOH) inhibits the human chorionic gonadotropin (hCG)-stimulated testosterone synthesis was studied in isolated rat Leydig cells in vitro. EtOH inhibited steroidogenesis, but this inhibition was reversed by L-glutamate (Glu) and an uncoupler of the oxidative phosphorylation, 2,4-dinitrophenol (DNP). The mechanism of EtOH-induced inhibition was studied by measuring steroidogenic precursors and comparing them with the cytosolic and mitochondrial NADH redox states during uncoupling or in the presence of Glu. DNP had a dual effect. Low concentrations abolished the EtOH-induced inhibition of progesterone to testosterone formation suggesting that the inhibitory step was at or before progesterone formation. A large concentration led to an overall decrease in steroidogenesis indicating toxic effects on steroidogenesis. The mitochondrial NADH/NAD+ ratio, measured as the 3-hydroxybutyrate/acetoacetate ratio, decreased simultaneously when steroidogenesis was stimulated, either during uncoupling or in the presence of Glu, whereas cytosolic NADH/NAD+ ratio, measured as lactate/pyruvate ratio showed no response. These results demonstrate that the rise in the mitochondrial NADH/NAD+ ratio rather than in the cytosolic one is connected with the inhibition of testosterone synthesis by EtOH in isolated Leydig cells. The EtOH-induced high mitochondrial NADH/NAD+ ratio may deplete mitochondrial oxalacetate concentrations. This can decrease the activity of several transport shuttles and interrupt the flow of mitochondrial citrate into the smooth endoplasmic reticulum, which then reflects to decreased rate of steroidogenesis in the presence of ethanol.  相似文献   

20.
Acholeplasma laidlawii A possesses a nicotinamide adenine dinucleotide (NAD)-dependent l(+)-lactate dehydrogenase (LDH) which is activated specifically by low concentrations of fructose-1, 6-diphosphate (FDP). Studies with partially purified enzyme show that the kinetic response to FDP is hyperbolic. The enzyme is inhibited by inorganic phosphate, adenosine triphosphate, and high concentrations of reduced NAD (NADH). Low activity is demonstrable in the absence of FDP at pH 6.0 to 7.2, but FDP is absolutely required in the region of pH 8. FDP causes an upward shift in the optimum pH of the enzyme, which is near 7.2 in tris (hydroxymethyl)aminomethane buffer. Activation of the enzyme by FDP is markedly affected by substrate concentration; FDP lowers the apparent K(m) for pyruvate and NADH. The affinity of the enzyme for pyruvate is also influenced by H(+) concentration. The pyruvate analogue alpha-ketobutyrate serves as an effective substrate for the enzyme; when it is utilized, the enzyme is still activated by FDP. Reversal of the pyruvate reduction reaction catalyzed by the enzyme can be demonstrated with the 3-acetylpyridine analogue of NAD. The catalytic properties of the A. laidlawii enzyme and the known FDP-activated LDHs which occur among lactic acid bacteria are discussed.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号