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.
In the Gram-positive methylotroph Bacillus methanolicus, methanol oxidation is catalyzed by an NAD-dependent methanol dehydrogenase (Mdh) that belongs to the type III alcohol dehydrogenase (Adh) family. It was previously shown that the in vitro activity of B. methanolicus Mdh is increased by the endogenous activator protein Act, a Nudix hydrolase. Here we show that this feature is not unique, but more widespread among type III Adhs in combination with Act or other Act-like Nudix hydrolases. In addition, we studied the effect of site directed mutations in the predicted active site of Mdh and two other type III Adhs with regard to activity and activation by Act.  相似文献   

2.
Methanol is an attractive substrate for biological production of chemicals and fuels. Engineering methylotrophic Escherichia coli as a platform organism for converting methanol to metabolites is desirable. Prior efforts to engineer methylotrophic E. coli were limited by methanol dehydrogenases (Mdhs) with unfavorable enzyme kinetics. We engineered E. coli to utilize methanol using a superior NAD-dependent Mdh from Bacillus stearothermophilus and ribulose monophosphate (RuMP) pathway enzymes from B. methanolicus. Using 13C-labeling, we demonstrate this E. coli strain converts methanol into biomass components. For example, the key TCA cycle intermediates, succinate and malate, exhibit labeling up to 39%, while the lower glycolytic intermediate, 3-phosphoglycerate, up to 53%. Multiple carbons are labeled for each compound, demonstrating a cycling RuMP pathway for methanol assimilation to support growth. By incorporating the pathway to synthesize the flavanone naringenin, we demonstrate the first example of in vivo conversion of methanol into a specialty chemical in E. coli.  相似文献   

3.
Two groups of metabolically related enzymes, the Group III family of Fe2+-dependent alcohol dehydrogenases (ADHs) and the separate subfamily of nucleoside diphosphates linked to x (nudix) hydrolases that activate Group III ADHs are under-characterized. Here we report the steady-state initial-velocity forward direction (alcohol → aldehyde) reaction of a Group III ADH, namely gamma-hydroxybutyrate dehydrogenase (GHBDH, UniProt: Q59104), cloned from Cupriavidus necator as a fusion protein. We also report the effects of nudix hydrolases on the GHBDH reaction. At optimal pH 9.0, the GHBDH reaction is activated ~2-fold by two different saturating purified nudix hydrolases, namely Bacillus methanolicus activator (ACT, UniProt: I3EA59) and Escherichia coli NudF (UniProt Q93K97) proteins. At physiological pH values of ~7.0, ACT activates by >3.5-fold. Initial-rate characterization at pH 9.0 of the forward direction un-activated and ACT-activated reactions show for both cases competitive inhibition by the product succinic semialdehyde versus GHB, and noncompetitive inhibitions by the three other substrate-product combinations. This pattern is consistent with NAD+ binding first in Mono-Iso Theorell-Chance kinetics. Mutants of some possibly important residues in GHBDH also were characterized. H265, conserved among all Group III ADHs and previously proposed to be a critical general base, is only ~4-fold helpful for GHBDH activity relevant to H265A. The four previously proposed conserved Fe2+ chelators (D193, H197, H261 and H280) each are essential for GHBDH activity. A 2-step explanation for cross-species stimulation by sub-stoichiometric ACT in the forward direction and confirmed lack of ACT stimulation in the reverse direction reaction is proposed.  相似文献   

4.
We have developed a single-step method for the purification of NADP+-dependent alcohol dehydrogenase fromEntamoeba histolyticaand NAD+-dependent alcohol dehydrogenase fromSaccharomyces cerevisiae.It is based on the affinity for zinc of both enzymes. The amebic enzyme was purified almost 800 times with a recovery of 54% and the yeast enzyme was purified 30 times with a recovery of 100%. The kinetic constants of the purified enzymes were similar to those reported for other purification methods. With mammalian alcohol dehydrogenase, we obtained a 40-kDa band suggestive of purified alcohol dehydrogenase, but we failed to retain enzymatic activity in this preparation. Our results suggest that the described method is more applicable to the purification of tetrameric alcohol dehydrogenases.  相似文献   

5.
Malate oxidation in plant mitochondria proceeds through the activities of two enzymes: a malate dehydrogenase and a NAD+-dependent malic enzyme. In cauliflower, mitochondria malate oxidation via malate dehydrogenase is rotenone- and cyanide-sensitive. Addition of exogenous NAD+ stimulates the oxidation of malate via malic enzyme and generates an electron flux that is both rotenone- and cyanide-insensitive. The same effects of exogenous NAD+ are also observed with highly cyanide-sensitive mitochondria from white potato tubers or with mitochondria from spinach leaves. Both enzymes are located in the matrix, but some experimental data also suggest that part of malate dehydrogenase activity is also present outside the matrix compartment (adsorbed cytosolic malate dehydrogenase?). It is concluded that malic enzyme and a specific pool of NAD+/NADH are connected to the cyanide-insensitive alternative pathway by a specific rotenone-insensitive NADH dehydrogenase located on the inner face of the inner membrane. Similarly, malate dehydrogenase and another specific pool of NAD+/NADH are connected to the cyanide- (and antimycin-) sensitive pathway by a rotenone-sensitive NADH dehydrogenase located on the inner face of the inner membrane. A general scheme of electron transport in plant mitochondria for the oxidation of malate and NADH can be given, assuming that different pools of ubiquinone act as a branch point between various dehydrogenases, the cyanide-sensitive cytochrome pathway and the cyanide-insensitive alternative pathway.  相似文献   

6.
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   

7.
The metabolic pathways of one-carbon compounds utilized by colorless sulfur bacterium Beggiatoa leptomitoformis D-402 were revealed based on comprehensive analysis of its genomic organization, together with physiological, biochemical and molecular biological approaches. Strain D-402 was capable of aerobic methylotrophic growth with methanol as a sole source of carbon and energy and was not capable of methanotrophic growth because of the absence of genes of methane monooxygenases. It was established that methanol can be oxidized to CO2 in three consecutive stages. On the first stage methanol was oxidized to formaldehyde by the two PQQ (pyrroloquinolinequinone)-dependent methanol dehydrogenases (MDH): XoxF and Mdh2. Formaldehyde was further oxidized to formate via the tetrahydromethanopterin (H4MPT) pathway. And on the third stage formate was converted to CO2 by NAD+-dependent formate dehydrogenase Fdh2. Finally, it was established that endogenous CO2, formed as a result of methanol oxidation, was subsequently assimilated for anabolism through the Calvin–Benson–Bassham cycle. The similar way of one-carbon compounds utilization also exists in representatives of another freshwater Beggiatoa species—B. alba.  相似文献   

8.
The subcellular distribution of NADP+ and NAD+-dependent glucose-6-phosphate and galactose-6-phosphate dehydrogenases were studied in rat liver, heart, brain, and chick brain. Only liver particulate fractions oxidized glucose-6-phosphate and galactose-6-phosphate with either NADP+ or NAD+ as cofactor. While all of the tissues examined had NADP+-dependent glucose-6-phosphate dehydrogenase activity, only rat liver and rat brain soluble fractions had NADP+-dependent galactose-6-phosphate dehydrogenase activity. Rat liver microsomal and rat brain soluble galactose-6-phosphate dehydrogenase activities were kinetically different (Km's 0.5 mm and 10 mm, respectively, for galactose-6-phosphate), although their reaction products were both 6-phosphogalactonate. Rat brain subcellular fractions did not oxidize 6-phosphogalactonate with either NADP+ or NAD+ cofactors but phosphatase activities hydrolyzing 6-phosphogalactonate, galactose-6-phosphate and galactose-1-phosphate were found in crude brain homogenates. In addition, galactose-6-phosphate and 6-phosphogalactonate were tested as inhibitors of various enzymes, with largely negative results, except that 6-phosphogalactonate was a competitive inhibitor (Ki = 0.5 mM) of rat brain 6-phosphogluconate dehydrogenase.  相似文献   

9.
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   

10.
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   

11.
A novel type of alcohol dehydrogenase enzyme has been characterized from Gram-positive methylotrophic (Bacillus methanolicus, the actinomycetes Amycolatopsis methanolica and Mycobacterium gastri) and non-methylotrophic bacteria (Rhodococcus strains). Its in vivo role is in oxidation of methanol and other primary alcohols. B. methanolicus displays activity of an NAD-dependent methanol dehydrogenase (MDH), which is strongly stimulated by a specific (activator) protein. A. methanolica and M. gastri use an N,N′- dimethyl-4-nitrosoaniline (NDMA)-dependent MDH (methanol: NDMA oxidoreductase; MNO). MDH (43 kDa subunit) and MNO (49 kDa subunit) possess similar decameric structures with five-fold symmetry. Both proteins contain Zn2+- and Mg2+-ions and tightly (but non-covalently) bound NAD(P)(H) cofactors. These nicotinoproteins share a high degree of sequence similarity; they belong to Family III of NAD(P)-dependent alcohol dehydrogenases (ADH). A. methanolica extracts also possess dye (dichlorophenol indophenol, DCPIP and 3-[4,5-dimethylthiazol-2-yl]-2,5-diphenyl-tetrazolium bromide, MTT)-linked ADH activities. These represent the overall activities of multi-enzyme systems. MNO is part of the MTT–ADH complex. The other two proteins and their cofactors most likely participate in transfer of reducing equivalents from the NADPH cofactor in MNO to the respiratory chain.  相似文献   

12.
Two types of 15-hydroxyprostaglandin dehydrogenase (NAD+ and NADP+ dependent) were demonstrated in bovine mesentric arteries and veins. The 15-hydroxyprostaglandin dehydrogenase activity was found in the high-speed supernatant, suggesting that these enzymes are associated with the cytoplasmic fraction of the blood vessels. The levels of activities of both NAD+- and NADP+-dependent dehydrogenases were similar in mesentric blood vessels. Prostaglandin F was preferred to the prostaglandin E2 as subtrate by both NAD+ and NADP+ dependent enzymes. The presence of 15-hydroxyprostaglandin dehydrogenase in blood vessels may play a siginificant role in the regulation of intracellular levels of prostaglandins of the E and F series in blood vessels.  相似文献   

13.
Aldehyde dehydrogenases (ALDHs) have been well established in all three domains of life and were shown to play essential roles, e.g., in intermediary metabolism and detoxification. In the genome of Sulfolobus solfataricus, five paralogs of the aldehyde dehydrogenases superfamily were identified, however, so far only the non-phosphorylating glyceraldehyde-3-phosphate dehydrogenase (GAPN) and α-ketoglutaric semialdehyde dehydrogenase (α-KGSADH) have been characterized. Detailed biochemical analyses of the remaining three ALDHs revealed the presence of two succinic semialdehyde dehydrogenase (SSADH) isoenzymes catalyzing the NAD(P)+-dependent oxidation of succinic semialdehyde. Whereas SSO1629 (SSADH-I) is specific for NAD+, SSO1842 (SSADH-II) exhibits dual cosubstrate specificity (NAD(P)+). Physiological significant activity for both SSO-SSADHs was only detected with succinic semialdehyde and α-ketoglutarate semialdehyde. Bioinformatic reconstructions suggest a major function of both enzymes in γ-aminobutyrate, polyamine as well as nitrogen metabolism and they might additionally also function in pentose metabolism. Phylogenetic studies indicated a close relationship of SSO-SSALDHs to GAPNs and also a convergent evolution with the SSADHs from E. coli. Furthermore, for SSO1218, methylmalonate semialdehyde dehydrogenase (MSDH) activity was demonstrated. The enzyme catalyzes the NAD+- and CoA-dependent oxidation of methylmalonate semialdehyde, malonate semialdehyde as well as propionaldehyde (PA). For MSDH, a major function in the degradation of branched chain amino acids is proposed which is supported by the high sequence homology with characterized MSDHs from bacteria. This is the first report of MSDH as well as SSADH isoenzymes in Archaea.  相似文献   

14.
The interactions of CL4, a biomimetic analogue of NAD+ comprising a nicotinamide functionality coupled via a triazine ring to a dibenzenesulphonic acid unit, and of a series of analogues, with HLADH and other dehydrogenases have been compared to those of the natural coenzymes NAD(P)+. CL4, together with one analogue with one of the sulphonic acid groups shifted by one position and another analogue with a single benzenedisulphonic acid unit, have been shown to be functional mimics of NAD+ in the oxidation of butan-1-ol by horse liver alcohol dehydrogenase (HLADH). A combination of discontinuous HPLC-based assays and continuous fluorescence based assays were used to deduce approximate kinetic constants for this reaction, using the artificial coenzymes, at pH 7.5 and 37°C. HLADH demonstrated a Vmax with the most active analogue which was 4% of that with NAD+. The substrate specificity of HLADH using these coenzymes was found to change relative to that using the natural coenzyme. Activity was sought from a range of other dehydrogenases: Bacillus megaterium glucose dehydrogenase, Leuconostoc mesenteroides glucose-6-phosphate dehydrogenase and sheep liver sorbitol dehydrogenase; all displayed activity using a range of the biomimetic coenzymes.  相似文献   

15.
NAD+-dependent and NADP+-dependent glyceraldehyde-3-phosphate (G-3-P) dehydrogenases were isolated from Euglena gracilis and characterized as to their physical and chemical parameters. NAD+-G-3-P dehydrogenase was found to have a strong resemblance to similar enzymes from muscle tissue. It has a molecular weight of about 140,000, four subunits of identical size and charge, and a single species of NH2-terminal amino acid. Two sulfhydryl groups per subunit are present, one of which is directly involved in the catalytic activity and is rapidly titratable. The enzyme also exhibits the “half the sites reactivity” of sulfhydryl groups as defined by O. P. Malhotra and S. A. Bernhard ((1968) J. Biol. Chem. 243, 1243). The pH and temperature optima are also similar to those of the enzymes from muscle tissue, as are the reaction kinetics and the strict specificity for NAD+.NADP+-dependent G-3-P dehydrogenase is different in many respects. Its molecular weight is slightly lower (~136,000) than that of the NAD+ enzyme, though it also consists of four subunits. It has a higher affinity for the reverse reaction substrates, in line with its probable function in vivo in CO2 fixation. There is only one sulfhydryl group per subunit, and that is not involved in activity, suggesting a difference in reaction mechanisms between the two enzymes. The NADP+-dependent enzyme exhibits activation by ATP, whereas the NAD+-dependent enzyme is competitively inhibited by this nucleotide.The greatest difference observed is in the physical characteristics of the enzymes. NADP+-G-3-P dehydrogenase was highly hydrophobic. Its solubility in a 10% aqueous solution of p-dioxane was approximately four to five times that of the NAD+-enzyme. Isolation of the enzyme was accomplished by fractionation in 1,2-dimethoxyethane, which also stabilized the enzymatic activity, as did aqueous p-dioxane. The high axial ratio of the NADP+-enzyme (~9) coupled with its very low degree of hydration as well as the high degree of amidation of the dicarboxylic amino acids (>90%) indicates that the exterior of the enzyme molecule is probably hydrophobic in nature. This is in agreement with its in vivo hydrophobic environment in the chloroplast membrane and explains the lability of the enzyme once extracted into an aqueous environment as well as its stabilization in solvents.  相似文献   

16.
Drosophila alcohol dehydrogenase (ADH) is an NAD(H)-dependent oxidoreductase that catalyzes the oxidation of alcohols and aldehydes. Structurally and biochemically distinct from all the reported ADHs (typically, the mammalian medium-chain dehydrogenase/reductase–ethanol-metabolizing enzyme), it stands as the only small-alcohol transforming system that has originated from a short-chain dehydrogenase/reductase (SDR) ancestor. The crystal structures of the apo, binary (E·NAD+) and three ternary (E·NAD+·acetone, E·NAD+·3-pentanone and E·NAD+·cyclohexanone) forms of Drosophila lebanonensis ADH have allowed us to infer the structural and kinetic features accounting for the generation of the ADH activity within the SDR lineage.  相似文献   

17.
Two inducible NADP+-dependent glycerol dehydrogenase (GlcDH) activities were identified in Mucor circinelloides strain YR-1. One of these, denoted iGlcDH2, was specifically induced by n-decanol when it was used as sole carbon source in the culture medium, and the second, denoted iGlcDH1, was induced by alcohols and aliphatic or aromatic hydrocarbons when glycerol was used as the only substrate. iGlcDH2 was found to have a much broader substrate specificity than iGlcDH1, with a low activity as an ethanol dehydrogenase with NAD+ or NADP+ as cofactor. Both isozymes showed an optimum pH for activity of 9.0 in Tris-HCl buffer and are subject to carbon catabolite repression. In contrast, the constitutive NADP+-dependent glycerol dehydrogenases (GlcDHI, II, and III) were only present in cell extracts when the fungus was grown in glycolytic carbon sources or glycerol under oxygenation, and their optimum pH was 7.0 in Tris-HCl buffer. In addition to these five NADP+-dependent glycerol dehydrogenases, a NAD+-dependent alcohol dehydrogenase is also present in glycerol or n-decanol medium; this enzyme was found to have weak activity as a glycerol dehydrogenase.  相似文献   

18.
Extracts of Pseudomonas C grown on methanol as sole carbon and energy source contain a methanol dehydrogenase activity which can be coupled to phenazine methosulfate. This enzyme catalyzes two reactions namely the conversion of methanol to formaldehyde (phenazine methosulfate coupled) and the oxidation of formaldehyde to formate (2,6-dichloroindophenol-coupled). Activities of glutathione-dependent formaldehyde dehydrogenase (NAD+) and formate dehydrogenase (NAD+) were also detected in the extracts.The addition of d-ribulose 5-phosphate to the reaction mixtures caused a marked increase in the formaldehyde-dependent reduction of NAD+ or NADP+. In addition, the oxidation of [14C]formaldehyde to CO2, by extracts of Pseudomonas C, increased when d-ribulose 5-phosphate was present in the assay mixtures.The amount of radioactivity found in CO2, was 6.8-times higher when extracts of methanol-grown Pseudomona C were incubated for a short period of time with [1-14C]glucose 6-phosphate than with [U-14C]glucose 6-phosphate.These data, and the presence of high specific activities of hexulose phosphate synthase, phosphoglucoisomerase, glucose-6-phosphate dehydrogenase and 6-phosphogluconate dehydrogenase indicate that in methanol-grown Pseudomonas C, formaldehyde carbon is oxidized to CO2 both via a cyclic pathway which includes the enzymes mentioned and via formate as an oxidation intermediate, with the former predominant.  相似文献   

19.
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   

20.
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   

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