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.
We have previously shown that a range of nicotinamide containing ‘biomimetic coenzymes’ function as active analogues of NAD+ in the oxidation of alcohols by horse liver alcohol dehydrogenase (HLADH), despite their apparently astonishing lack of structural similarity to the natural coenzyme. The simplest structure as yet shown to exhibit activity is the biomimetic coenzyme CL4. To investigate the effect of the structure of this truncated artificial coenzyme on its activity, a range of close structural analogues of CL4 were designed, synthesized and characterized. The electrochemical reduction potentials of the analogues were strongly influenced by the nature of the groups attached to the pyridine ring. All of the analogues could be chemically reduced using sodium borohydride, to give compounds with altered UV‐visible absorption and fluorescence properties. An HPLC‐based assay suggested that two of the new analogues were coenzymically active in the oxidation of butan‐1‐ol by HLADH, with one displaying a significantly higher activity than CL4. The results demonstrate which features of the structures of the coenzymes lead to desirable electrochemical and spectroscopic properties, but suggest that the structural requirements for a functional coenzyme are quite stringent. These observations may be used to design an artificial coenzyme which combines the best features of those studied so far. Copyright Copyright © 1999 John Wiley & Sons, Ltd.  相似文献   

2.
A range of biomimetic analogues of the nicotinamide nucleotide coenzymes NAD(P)(H) have been developed based on the structure of a triazine dye template. These biomimetic redox coenzymes are relatively straightforward and inexpensive to synthesise and display NAD+-like activity with different dehydrogenases, despite their apparently minimal structural similarity to the native coenzyme NAD+. Horse liver alcohol dehydrogenase oxidises butan-1-ol, using the most active biomimetic coenzyme (Nap 1), with a k cat value an order of magnitude lower and a K m for the coenzyme two orders of magnitude higher than those using native NAD+. The enzymatically reduced biomimetic coenzymes may be reoxidised by phenazine methosulfate. We believe that these coenzymes may find applications in biotransformations and biosensors, and in the development of biomimetic catalysts where the redox enzyme itself is replaced by a synthetic binding site. Received: 26 October 1998 / Received revision: 25 January 1999 / Accepted: 31 January 1999  相似文献   

3.
The steady-state kinetics of alcohol dehydrogenases (alcohol:NAD+ oxidoreductase, EC 1.1.1.1 and alcohol:NADP+ oxidoreductase, EC 1.1.1.2), lactate dehydrogenases (l-lactate:NAD+ oxidoreductase, EC 1.1.1.27 and d-lactate:NAD+ oxidoreductase, EC 1.1.1.28), malate dehydrogenase (l-malate:NAD+ oxidoreductase, EC 1.1.1.37), and glyceraldehyde-3-phosphate dehydrogenases [d-glyceraldehyde-3-phosphate:NAD+ oxidoreductase (phosphorylating), EC 1.2.1.12] from different sources (prokaryote and eukaryote, mesophilic and thermophilic organisms) have been studied using NAD(H), N6-(2-carboxyethyl)-NAD(H), and poly(ethylene glycol)-bound NAD(H) as coenzymes. The kinetic constants for NAD(H) were changed by carboxyethylation of the 6-amino group of the adenine ring and by conversion to macromolecular form. Enzymes from thermophilic bacteria showed especially high activities for the derivatives. The relative values of the maximum velocity (NAD = 1) of Thermus thermophilus malate dehydrogenase for N6-(2-carboxyethyl)-NAD and poly(ethylene glycol)-bound NAD were 5.7 and 1.9, respectively, and that of Bacillus stearothermophilus glyceraldehyde-3-phosphate dehydrogenase for poly(ethylene glycol)-bound NAD was 1.9.  相似文献   

4.
Dehydrogenation of 2-trimethylsilyl-1-propanol (1) was carried out with horse liver alcohol dehydrogenase (HLADH, EC 1.1.1.1). It was found that the hydrogenation of 1 proceeded enantioselectively with only HLADH and a catalytic amount of NAD+ due to in-situ NAD+ regeneration based on a specific property of -carbonylsilanes. That is, (+)-1 was enantioselectively dehydrogenated by HLADH to 2-trimethylsilyl-1-propanal, which was spontaneously degraded by addition of water into trimethylsilanol and n-propanal. Then, NAD+ was regenerated through HLADH-catalyzed reduction of n-propanal to n-propanol. On the other hand, dehydrogenation of the carbon analogue of 1 was negligible with a catalytic amount of NAD+, indicating that the in-situ NAD+ regeneration was not available without the specific property of organosilicon compounds. Other primary -hydroxysilanes having different substituents on the chiral center or on the silicon atom were also found to serve as substrates in enantioselective dehydrogenation by HLADH with this novel NAD+ regeneration system. Chiral recognition of HLADH toward primary alcohols is also discussed.  相似文献   

5.
The mechanism by which Helminthosporium maydis race T toxin inhibits respiration dependent on NAD+-linked substrates in T cytoplasm corn mitochondria was investigated. The toxin did not cause leakage of the soluble matrix enzyme malate dehydrogenase from the mitochondria or inhibit malate dehydrogenase or isocitrate dehydrogenase directly. The toxin did increase the permeability of the inner membranes of T cytoplasm, but not N cytoplasm, mitochondria to NAD+. Added NAD+ partially or fully restored toxin-inhibited electron transport in T cytoplasm mitochondria. Thiamin pyrophosphate had a similar effect when malate was the substrate. It was concluded that the inhibition of respiration of NAD+-linked substrates by the toxin is due to depletion of the intramitochondrial pool of NAD+ and other coenzymes.  相似文献   

6.
In this article we compare the kinetic behavior toward pyridine nucleotides (NAD+, NADH) of NAD+-malic enzyme, pyruvate dehydrogenase, isocitrate dehydrogenase, α-ketoglutarate dehydrogenase, and glycine decarboxylase extracted from pea (Pisum sativum) leaf and potato (Solanum tuberosum) tuber mitochondria. NADH competitively inhibited all the studied dehydrogenases when NAD+ was the varied substrate. However, the NAD+-linked malic enzyme exhibited the weakest affinity for NAD+ and the lowest sensitivity for NADH. It is suggested that NAD+-linked malic enzyme, when fully activated, is able to raise the matricial NADH level up to the required concentration to fully engage the rotenone-resistant internal NADH-dehydrogenase, whose affinity for NADH is weaker than complex I.  相似文献   

7.
Metabolic enzymes are usually characterized to have one specific function, and this is the case of UDP-glucose dehydrogenase that catalyzes the twofold NAD+-dependent oxidation of UDP-glucose into UDP-glucuronic acid. We have determined that this enzyme is also capable of participating in other cellular processes. Here, we report that the bacterial UDP-glucose dehydrogenase (UgdG) from Sphingomonas elodea ATCC 31461, which provides UDP-glucuronic acid for the synthesis of the exopolysaccharide gellan, is not only able to bind RNA but also acts as a ribonuclease. The ribonucleolytic activity occurs independently of the presence of NAD+ and the RNA binding site does not coincide with the NAD+ binding region. We have also performed the kinetics of interaction between UgdG and RNA. Moreover, computer analysis reveals that the N- and C-terminal domains of UgdG share structural features with ancient mitochondrial ribonucleases named MAR. MARs are present in lower eukaryotic microorganisms, have a Rossmannoid-fold and belong to the isochorismatase superfamily. This observation reinforces that the Rossmann structural motifs found in NAD+-dependent dehydrogenases can have a dual function working as a nucleotide cofactor binding domain and as a ribonuclease.  相似文献   

8.
Methanobacterium thermoautotrophicum (strain Marburg) was found to contain two malate dehydrogenases, which were partially purified and characterized. One was specific for NAD+ and catalyzed the dehydrogenation of malate at approximately one-third of the rate of oxalacetate reduction, and the other could equally well use NAD+ and NADP+ as coenzyme and catalyzed essentially only the reduction of oxalacetate. Via the N-terminal amino acid sequences, the encoding genes were identified in the genome of M. thermoautotrophicum (strain ΔH). Comparison of the deduced amino acid sequences revealed that the two malate dehydrogenases are phylogenetically only distantly related. The NAD+-specific malate dehydrogenase showed high sequence similarity to l-malate dehydrogenase from Methanothermus fervidus, and the NAD(P)+-using malate dehyrogenase showed high sequence similarity to l-lactate dehydrogenase from Thermotoga maritima and l-malate dehydrogenase from Bacillus subtilis. A function of the two malate dehydrogenases in NADPH:NAD+ transhydrogenation is discussed. Received: 29 December 1997 / Accepted: 4 March 1998  相似文献   

9.
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   

10.
A new, acyclic NAD-analog, acycloNAD+ has been synthesized where the nicotinamide ribosyl moiety has been replaced by the nicotinamide (2-hydroxyethoxy)methyl moiety. The chemical properties of this analog are comparable to those of β-NAD+ with a redox potential of −324 mV and a 341 nm λmax for the reduced form. Both yeast alcohol dehydrogenase (YADH) and horse liver alcohol dehydrogenase (HLADH) catalyze the reduction of acycloNAD+ by primary alcohols. With HLADH 1-butanol has the highest Vmax at 49% that of β-NAD+. The primary deuterium kinetic isotope effect is greater than 3 indicating a significant contribution to the rate limiting step from cleavage of the carbon–hydrogen bond. The stereochemistry of the hydride transfer in the oxidation of stereospecifically deuterium labeled n-butanol is identical to that for the reaction with β-NAD+. In contrast to the activity toward primary alcohols there is no detectable reduction of acycloNAD+ by secondary alcohols with HLADH although these alcohols serve as competitive inhibitors. The net effect is that acycloNAD+ has converted horse liver ADH from a broad spectrum alcohol dehydrogenase, capable of utilizing either primary or secondary alcohols, into an exclusively primary alcohol dehydrogenase. This is the first example of an NAD analog that alters the substrate specificity of a dehydrogenase and, like site-directed mutagenesis of proteins, establishes that modifications of the coenzyme distance from the active site can be used to alter enzyme function and substrate specificity. These and other results, including the activity with α-NADH, clearly demonstrate the promiscuity of the binding interactions between dehydrogenases and the riboside phosphate of the nicotinamide moiety, thus greatly expanding the possibilities for the design of analogs and inhibitors of specific dehydrogenases.  相似文献   

11.
NAD+ and NADP+, chemically similar and with almost identical standard oxidation–reduction potentials, nevertheless have distinct roles, NAD+ serving catabolism and ATP generation whereas NADPH is the biosynthetic reductant. Separating these roles requires strict specificity for one or the other coenzyme for most dehydrogenases. In many organisms this holds also for glutamate dehydrogenases (GDH), NAD+-dependent for glutamate oxidation, NADP+-dependent for fixing ammonia. In higher animals, however, GDH has dual specificity. It has been suggested that GDH in mitochondria reacts only with NADP(H), the NAD+ reaction being an in vitro artefact. However, contrary evidence suggests mitochondrial GDH not only reacts with NAD+ but maintains equilibrium using the same pool as accessed by β-hydroxybutyrate dehydrogenase. Another complication is the presence of an energy-linked dehydrogenase driving NADP+ reduction by NADH, maintaining the coenzyme pools at different oxidation–reduction potentials. Its coexistence with GDH makes possible a futile cycle, control of which is not yet properly explained. Structural studies show NAD+-dependent, NADP+-dependent and dual-specificity GDHs are closely related and a few site-directed mutations can reverse specificity. Specificity for NAD+ or for NADP+ has probably emerged repeatedly during evolution, using different structural solutions on different occasions. In various GDHs the P7 position in the coenzyme-binding domain plays a key role. However, whereas in other dehydrogenases an acidic P7 residue usually hydrogen bonds to the 2′- and 3′-hydroxyls, dictating NAD+ specificity, among GDHs, depending on detailed conformation of surrounding residues, an acidic P7 may permit binding of NAD+ only, NADP+ only, or in higher animals both.  相似文献   

12.
The kinetic locking-on strategy utilizes soluble analogues of the target enzymes' specific substrate to promote selective adsorption of individual NAD+-dependent dehydrogenases on their complementary immobilized cofactor derivative. Application of this strategy to the purification of NAD+-dependent dehydrogenases from crude extracts has proven that it can yield bioaffinity systems capable of producing one-chromatographic-step purifications with yields approaching 100%. However, in some cases the purified enzyme preparation was found to be contaminated with other proteins weakly bound to the immobilized cofactor derivative through binary complex formation and/or nonspecific interactions, which continuously “dribbled” off the matrix during the chromatographic procedure. The fact that this problem can be overcome by including a short pulse of 5′-AMP (stripping ligand) in the irrigant a couple of column volumes prior to the discontinuation of the specific substrate analogue (locking-on ligand) is clear from the results presented in this report. The general effectiveness of this auxiliary tactic has been assessed using model studies and through incorporation into an actual purification from a crude cellular extract. The results confirm the usefulness of the stripping-ligand tactic for the resolution and purification of NAD+-dependent dehydrogenases when using the locking-on strategy. These studies have been carried out using bovine liver glutamate dehydrogenase (GDH, EC 1.4.1.3), yeast alcohol dehydrogenase (YADH, EC 1.1.1.1), porcine heart mitochondrial malate dehydrogenase (mMDH, EC 1.1.1.37), and bovine heart -lactate dehydrogenase ( -LDH, EC 1.1.1.27).  相似文献   

13.
Anaerobically grown cells of Saccharomyces cerevisiae entrapped in polyacrylamide gel have been shown to provide a stable source of alcohol dehydrogenase [(ADH) alcohol:NAD+ oxidoreductase, EC 1.1.1.1] for effective regeneration of NAD(H). This system was able to provide the coenzyme required for the operation of other dehydrogenases, such as lactate dehydrogenase [(LDH) l-lactate: NAD+ oxidoreductase, EC 1.1.1.27] and malate dehydrogenase [(MDH) l-malate:NAD+ oxidoreductase, EC 1.1.1.37]. Yeast cells coimmobilized with a dehydrogenase are capable of the reversible regeneration of the reduced or oxidized coenzyme, depending on the additions made. A two-cell system can also be constituted using the same strain of yeast, adapted differently. Cells grown anaerobically and aerobically as sources of ADH and MDH, respectively, can operate efficiently on coimmobilization. The system can be used repeatedly without measurable loss of efficiency.  相似文献   

14.
1. Two different gels have been prepared suitable for the separation of a number of enzymes, in particular NAD+-dependent dehydrogenases, by affinity chromatography. For both the matrix used was Sepharose 4B. For preparation (a), NAD+–Sepharose, 6-aminohexanoic acid has been coupled to the gel by the cyanogen bromide method and then NAD+ was attached by using dicyclohexylcarbodi-imide; for preparation (b), AMP–Sepharose, N6-(6-aminohexyl)-AMP has been coupled directly to cyanogen bromide-activated gel. 2. Affinity columns of both gels retain only the two enzymes when a mixture of bovine serum albumin, lactate dehydrogenase and glyceraldehyde 3-phosphate dehydrogenase is applied. Subsequent elution with the cofactor NAD+ yields glyceraldehyde 3-phosphate dehydrogenase whereas lactate dehydrogenase is eluted by applying the same molarity of the reduced cofactor. 3. The binding of both glyceraldehyde 3-phosphate dehydrogenase and lactate dehydrogenase to the gel tested, AMP–Sepharose, is strong enough to resist elution by gradients of KCl of up to at least 0.5m. A 0.0–0.15m gradient of the competitive inhibitor salicylate, however, elutes both enzymes efficiently and separately. 4. The elution efficiency of lactate dehydrogenase from AMP–Sepharose has been examined by using a series of eluents under comparable conditions of concentration etc. The approximate relative efficiencies are: 0 (lactate); 0 (lactate+semicarbazide); 0 (0.5mm-NAD+); 80 (lactate+NAD+); 95 (lactate+semicarbazide+NAD+); 100 (0.5mm-NADH). 5. All contaminating lactate dehydrogenase activity can be removed from commercially available crude pyruvate kinase in a single-step procedure by using AMP–Sepharose.  相似文献   

15.
Several denitrifying Pseudomonas strains contained an NADP+-specific 2-oxoglutarate dehydrogenase, in contrast to an NAD+-specific pyruvate dehydrogenase, if the cells were grown anaerobically with aromatic compounds. With non-aromatic substrates or after aerobic growth the coenzyme specificity of 2-oxoglutarate dehydrogenase changed to NAD+-specificity. The reaction stoichiometry and the apparent K m-values of the enriched enzymes were determined: pyruvate 0.5 mM, coenzyme A 0.05 mM, NAD+ 0.25 mM; 2-oxoglutarate 0.6 mM, coenzyme A 0.05 mM, NADP+ 0.03 mM. Isocitrate dehydrogenase was NADP+-specific. The findings suggest that these strains contained at least two lipoamide dehydrogenases, one NAD+-specific, the other NADP+-specific.  相似文献   

16.
Activity of isocitrate dehydrogenase, α-ketoglutarate dehydrogenase, malate dehydrogenase, and the NAD+/NADН ratio were studied in the liver mitochondrial fraction of rats with toxic hepatitis induced by acetaminophen under conditions of alimentary protein deficiency. Acetaminophen-induced hepatitis was characterized by a decrease of isocitrate dehydrogenase, α-ketoglutarate dehydrogenase and malate dehydrogenase activities, while the mitochondrial NAD+/NADН ratio remained at the control level. Modeling of acetaminophen-induced hepatitis in rats with alimentary protein deficiency caused a more pronounced decrease in the activity of studied Krebs cycle NAD+-dependent dehydrogenases and a 2.2-fold increase of the mitochondrial NAD+/NADН ratio.  相似文献   

17.
Glucose-6-phosphate dehydrogenase from Leuconostoc mesenteroides utilizes either NAD+ or NADP+ as coenzyme. Kinetic studies showed that NAD+ and NADP+ interact with different enzyme forms (Olive, C., Geroch, M. E., and Levy, H. R. (1971) J. Biol. Chem.246, 2047–2057). In the present study the techniques of fluorescence quenching and fluorescence enhancement were used to investigate the interaction between Leuconostoc mesenteroides glucose-6-phosphate dehydrogenase and coenzymes. In addition, kinetic studies were performed to examine interaction between the enzyme and various coenzyme analogs. The maximum quenching of protein fluorescence is 5% for NADP+ and 50% for NAD+. The dissociation constant for NADP+, determined from fluorescence quenching measurements, is 3 μm, which is similar to the previously determined Km of 5.7 μm and Ki of 5 μm. The dissociation constant for NAD+ is 2.5 mm, which is 24 times larger than the previously determined Km of 0.106 mm. Glucose 1-phosphate, a substrate-competitive inhibitor, lowers the dissociation constant and maximum fluorescence quenching for NAD+ but not for NADP+. This suggests that glucose 6-phosphate may act similarly and thus play a role in enabling the enzyme to utilize NAD+ under physiological conditions. When NADPH binds to the enzyme its fluorescence is enhanced 2.3-fold. The enzyme was titrated with NADPH in the absence and presence of NAD+; binding of these two coenzymes is competitive. The dissociation constant for NADPH from these measurements is 24 μm; the previously determined Ki is 37.6 μm. The dissociation constant for NAD′ is 2.8 mm, in satisfactory agreement with the value obtained from protein fluorescence quenching measurements. Various compounds which resemble either the adenosine or the nicotinamide portion of the coenzyme structure are coenzyme-competitive inhibitors; 2′,5′-ADP, the most inhibitory analog tested, gives NADP+-competitive and NAD+-noncompetitive inhibition, consistent with the kinetic mechanism previously proposed. By using pairs of coenzyme-competitive inhibitors it was shown in kinetic studies that the two portions of the NAD+ structure cannot be accommodated on the enzyme simultaneously unies they are covalently linked. Fluorescence studies showed that there are both “buried” and “exposed” tryptophan residues in the enzyme structure.  相似文献   

18.
The mechanisms and accurate control of citrate oxidation by Percoll-purified potato (Solanum tuberosum) tuber mitochondria were characterized in various metabolic conditions by recording time course evolution of the citric acid cycle related intermediates and O2 consumption. Intact potato tuber mitochondria showed good rates of citrate oxidation, provided that nonlimiting amounts of NAD+ and thiamine pyrophosphate were present in the matrix space. Addition of ATP increased initial oxidation rates, by activation of the energy-dependent net citrate uptake, and stimulated succinate and malate formation. When the intramitochondrial NADH to NAD+ ratio was high, α-ketoglutarate only was excreted from the matrix space. After addition of ADP, aspartate, or oxaloacetate, which decreased the NADH to NAD+ ratio, flux rates through the Krebs cycle dehydrogenases were strongly increased and α-ketoglutarate, succinate, and malate accumulated up to steady-state concentrations in the reaction medium. It was concluded that NADH to NAD+ ratio could be the primary signal for coordination of fluxes through electron transport chain or malate dehydrogenase and NAD+-linked Krebs cycle dehydrogenases. In addition, these results clearly showed that the tricarboxylic acid cycle could serve as an important source of carbon skeletons for extra-mitochondrial synthetic processes, according to supply and demand of metabolites.  相似文献   

19.
This study is concerned with further development of the kinetic locking-on strategy for bioaffinity purification of NAD+-dependent dehydrogenases. Specifically, the synthesis of highly substituted N6-linked immobilized NAD+ derivatives is described using a rapid solid-phase modular approach. Other modifications of the N6-linked immobilized NAD+ derivative include substitution of the hydrophobic diaminohexane spacer arm with polar spacer arms (9 and 19.5 Å) in an attempt to minimize nonbiospecific interactions. Analysis of the N6-linked NAD+ derivatives confirm (i) retention of cofactor activity upon immobilization (up to 97%); (ii) high total substitution levels and high percentage accessibility levels when compared to S6-linked immobilized NAD+ derivatives (also synthesized with polar spacer arms); (iii) short production times when compared to the preassembly approach to synthesis. Model locking-on bioaffinity chromatographic studies were carried out with bovine heart -lactate dehydrogenase ( -LDH, EC 1.1.1.27), bakers yeast alcohol dehydrogenase (YADH, EC 1.1.1.1) and Sporosarcinia sp. -phenylalanine dehydrogenase ( -PheDH, EC 1.4.1.20), using oxalate, hydroxylamine, and -phenylalanine, respectively, as locking-on ligands. Surprisingly, two of these test NAD+-dependent dehydrogenases (lactate and alcohol dehydrogenase) were found to have a greater affinity for the more lowly substituted S6-linked immobilized cofactor derivatives than for the new N6-linked derivatives. In contrast, the NAD+-dependent phenylalanine dehydrogenase showed no affinity for the S6-linked immobilized NAD+ derivative, but was locked-on strongly to the N6-linked immobilized derivative. That this locking-on is biospecific is confirmed by the observation that the enzyme failed to lock-on to an analogous N6-linked immobilized NADP+ derivative in the presence of -phenylalanine. This differential locking-on of NAD+-dependent dehydrogenases to N6-linked and S6-linked immobilized NAD+ derivatives cannot be explained in terms of final accessible substitutions levels, but suggests fundamental differences in affinity of the three test enzymes for NAD+ immobilized via N6-linkage as compared to thiol-linkage.  相似文献   

20.
The catalytic properties of the purified horseshoe crab and seaworm d-lactate dehydrogenases were determined and compared with those of several l-lactate dehydrogenases. Apparent Km's and degrees of substrate inhibition have been determined for both enzymes for pyruvate, d-lactate, NAD+ and NADH. They are similar to those found for l-lactate dehydrogenases. The Limulus “muscle”-type lactate dehydrogenase is notably different from the “heart”-type lactate dehydrogenase of this organism in a number of properties.The Limulus heart and muscle enzymes have been shown by several criteria to be stereospecific for d-lactate. They also stereospecifically transfer the 4-α hydrogen of NADH to pyruvate. The turnover number for purified Limulus muscle lactate dehydrogenase is 38,000 moles NADH oxidized per mole of enzyme, per minute. Limulus and Nereis lactate dehydrogenases are inhibited by oxamate and the reduced NAD-pyruvate adduct.Limulus muscle lactate dehydrogenase is stoichiometrically inhibited by para-hydroxymercuribenzoate. Extrapolation to two moles parahydroxymercuribenzoate bound to one mole of enzyme yields 100% inhibition. Alkylation by iodoacetamide or iodoacetate occurs even in the absence of urea or guanidine-HCl. Evidence suggests that the reactive sulfhydryl group may not be located at the coenzyme binding site.Reduced coenzyme (NADH or the 3-acetyl-pyridine analogue of NADH) stoichiometrically binds to Limulus muscle lactate dehydrogenase (two moles per mole of enzyme).Several pieces of physical and catalytic evidence suggest that the d- and l-lactate dehydrogenase are products of homologous genes. A consideration of a possible “active site” shows that as few as one or two key conservative amino acid changes could lead to a reversal of the lactate stereospecificity.  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