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 375 毫秒
1.
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   

2.
3.
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   

4.
NAD+-dependent formate dehydrogenase (FDH-EC 1.2.1.2) is an important enzyme to regenerate valuable NADH required by NAD+-dependent oxidoreductases in enzyme catalysis. The limitation in the thermostability of FDH enzyme is a crucial problem for development of biotechnological and industrial processes, despite of its advantages. In this study, to investigate the contribution of surface electrostatic interaction to the thermostability of FDH from Candida methylica (cmFDH) N187E, H13E, Q105R, N300E, N147R N300E/N147R, N187E/Q105R, N187E/N147R,Y160R, Y302R, Y160E and Y302E mutants were designed using a homology model of cmFDH based on Candida boidinii (cb) by considering electrostatic interactions on the protein surface. The effects of site-specific engineering on the stability of this molecule was analyzed according to minimal model of folding and assembly reaction and deduced equilibrium properties of the native system with respect to its thermal and denaturant sensitivities. It was observed that mutations did not change the unfolding pattern of native cmFDH and increased numbers of electrostatic interactions can cause either stabilizing or destabilizing effect on the thermostability of this protein. The thermodynamic and kinetic results suggested that except relatively improved mutants, three out of the nine single mutations increased the melting temperature of cmFDH enzyme.  相似文献   

5.
Despite significant influence of secondary bile acids on human health and disease, limited structural and biochemical information is available for the key gut microbial enzymes catalyzing its synthesis. Herein, we report apo‐ and cofactor bound crystal structures of BaiA2, a short chain dehydrogenase/reductase from Clostridium scindens VPI 12708 that represent the first protein structure of this pathway. The structures elucidated the basis of cofactor specificity and mechanism of proton relay. A conformational restriction involving Glu42 located in the cofactor binding site seems crucial in determining cofactor specificity. Limited flexibility of Glu42 results in imminent steric and electrostatic hindrance with 2′‐phosphate group of NADP(H). Consistent with crystal structures, steady state kinetic characterization performed with both BaiA2 and BaiA1, a close homolog with 92% sequence identity, revealed specificity constant (kcat/KM) of NADP+ at least an order of magnitude lower than NAD+. Substitution of Glu42 with Ala improved specificity toward NADP+ by 10‐fold compared to wild type. The cofactor bound structure uncovered a novel nicotinamide‐hydroxyl ion (NAD+‐OH?) adduct contraposing previously reported adducts. The OH? of the adduct in BaiA2 is distal to C4 atom of nicotinamide and proximal to 2′‐hydroxyl group of the ribose moiety. Moreover, it is located at intermediary distances between terminal functional groups of active site residues Tyr157 (2.7 Å) and Lys161 (4.5 Å). Based on these observations, we propose an involvement of NAD+‐OH? adduct in proton relay instead of hydride transfer as noted for previous adducts. Proteins 2014; 82:216–229. © 2013 Wiley Periodicals, Inc.  相似文献   

6.
The ethanologenic bacterium Zymomonas mobilis ZM4 is of special interest because it has a high ethanol yield. This is made possible by the two alcohol dehydrogenases (ADHs) present in Z. mobilis ZM4 (zmADHs), which shift the equilibrium of the reaction toward the synthesis of ethanol. They are metal-dependent enzymes: zinc for zmADH1 and iron for zmADH2. However, zmADH2 is inactivated by oxygen, thus implicating zmADH2 as the component of the cytosolic respiratory system in Z. mobilis. Here, we show crystal structures of zmADH2 in the form of an apo-enzyme and an NAD+-cofactor complex. The overall folding of the monomeric structure is very similar to those of other functionally related ADHs with structural variations around the probable substrate and NAD+ cofactor binding region. A dimeric structure is formed by the limited interactions between the two subunits with the bound NAD+ at the cleft formed along the domain interface. The catalytic iron ion binds near to the nicotinamide ring of NAD+, which is likely to restrict and locate the ethanol to the active site together with the oxidized Cys residue and several nonpolar bulky residues. The structures of the zmADH2 from the proficient ethanologenic bacterium Z. mobilis, with and without NAD+ cofactor, and modeling ethanol in the active site imply that there is a typical metal-dependent catalytic mechanism.  相似文献   

7.
Tuberculosis (TB) remains a worldwide problem and the need for new drugs is increasingly more urgent with the emergence of multidrug- and extensively-drug resistant TB. Inosine 5’-monophosphate dehydrogenase 2 (IMPDH2) from Mycobacterium tuberculosis (Mtb) is an attractive drug target. The enzyme catalyzes the conversion of inosine 5’-monophosphate into xanthosine 5’-monophosphate with the concomitant reduction of NAD+ to NADH. This reaction controls flux into the guanine nucleotide pool. We report seventeen selective IMPDH inhibitors with antitubercular activity. The crystal structures of a deletion mutant of MtbIMPDH2 in the apo form and in complex with the product XMP and substrate NAD+ are determined. We also report the structures of complexes with IMP and three structurally distinct inhibitors, including two with antitubercular activity. These structures will greatly facilitate the development of MtbIMPDH2-targeted antibiotics.  相似文献   

8.
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   

9.
Biphenyl dehydrogenase, a member of short-chain dehydrogenase/reductase enzymes, catalyzes the second step of the biphenyl/polychlorinated biphenyls catabolic pathway in bacteria. To understand the molecular basis for the broad substrate specificity of Pandoraea pnomenusa strain B-356 biphenyl dehydrogenase (BphBB-356), the crystal structures of the apo-enzyme, the binary complex with NAD+, and the ternary complexes with NAD+-2,3-dihydroxybiphenyl and NAD+-4,4′-dihydroxybiphenyl were determined at 2.2-, 2.5-, 2.4-, and 2.1-Å resolutions, respectively. A crystal structure representing an intermediate state of the enzyme was also obtained in which the substrate binding loop was ordered as compared with the apo and binary forms but it was displaced significantly with respect to the ternary structures. These five structures reveal that the substrate binding loop is highly mobile and that its conformation changes during ligand binding, starting from a disorganized loop in the apo state to a well organized loop structure in the ligand-bound form. Conformational changes are induced during ligand binding; forming a well defined cavity to accommodate a wide variety of substrates. This explains the biochemical data that shows BphBB-356 converts the dihydrodiol metabolites of 3,3′-dichlorobiphenyl, 2,4,4′-trichlorobiphenyl, and 2,6-dichlorobiphenyl to their respective dihydroxy metabolites. For the first time, a combination of structural, biochemical, and molecular docking studies of BphBB-356 elucidate the unique ability of the enzyme to transform the cis-dihydrodiols of double meta-, para-, and ortho-substituted chlorobiphenyls.  相似文献   

10.
为了调查5℃低温处理是否改变家蚕Bombyx mori卵滞育NAD代谢, 本研究利用HPLC和分光光度法测定了经25℃和5℃分别处理的滞育卵中NADH 含量、 NAD+含量、 乳酸脱氢酶(LDH)活性和胞质苹果酸脱氢酶(cMDH)活性。结果表明: 5℃处理的NAD(NADH + NAD+)含量和cMDH活性分别增加了106%和53%, 并且显著高于25℃处理(P< 0.01); 但是两种处理的NADH/NAD+比值和LDH活性没有显著差异(P> 0.05)。据此推测, 5℃低温处理加强了家蚕滞育卵NAD+合成和再生能力。  相似文献   

11.
The role of anchorless proteins on the surface of most pathogenic microorganisms has long been studied in context to their interactions with multiple host proteins, facilitating the dissemination of pathogen within the host tissues. In order to gain more insights into anthrax pathogenesis, we hereby report the presence of a prominent moonlighting enzyme, Glyceraldehyde-3-phosphate dehydrogenase (GAPDH) on the surface and in the extracellular medium of Bacillus anthracis. Out of the three heterologously expressed recombinant isoforms, rGapA (334 amino acids in native form; GapA) showed a significant NAD+ mediated GAPDH activity, whereas rGapB (342 amino acids in native form; GapB) showed a slight activity with NADP+. The rGapN (479 amino acids in native form; GapN) was enzymatically inactive with either NAD+ or NADP+. GapA was ascertained to be present in the extracellular medium and on the surface of B. anthracis. On the other hand, GapN was absent from both the surface and extracellular medium, whereas GapB was scarcely present on the surface of B. anthracis. Human plasminogen predominantly interacted with the rGapA isoform at physiological concentrations and the interaction was found to be lysine dependent. Immunization with rGapA resulted in a significant protection upon challenge with Bacillus anthracis in the murine model.  相似文献   

12.
Mark A. Jensen  Philip J. Elving 《BBA》1984,764(3):310-315
The rate constant, kd, for the dimerization of the free radical (NAD·), produced on the initial one-electron reduction of NAD+, was measured by double potential-step chronoamperometry, fast-scan cyclic voltammetry (cathodic-anodic peak current ratio) and slow-scan cyclic voltammetry (peak potential shift) for a medium in which neither NAD+ nor its reduction products are adsorbed at the solution/electrode interface. All three methods give concordant values of kd (approx. 3·107 M?1·s?1), which are in reasonable accord with the values determined by pulse radiolysis but are considerably greater than values previously determined electrochemically. For the NAD+/NAD· couple, the heterogeneous rate constant (ks,h) exceeds 1 cm·s?1 at 25°C and the formal potential (E0c) vs. sce is ? 1.155 V at 25°C and ? 1.149 V at 1°C at pH 9.1, with an uncertainty of about ±0.005 V.  相似文献   

13.
Crystal structures of short chain dehydrogenase-like l-threonine dehydrogenase from Cupriavidus necator (CnThrDH) in the apo and holo forms were determined at 2.25 and 2.5 Å, respectively. Structural comparison between the apo and holo forms revealed that four regions of CnThrDH adopted flexible conformations when neither NAD+ nor l-Thr were bound: residues 38–59, residues 77–87, residues 180–186, and the catalytic domain. Molecular dynamics simulations performed at the 50-ns time scale revealed that three of these regions remained flexible when NAD+ was bound to CnThrDH: residues 80–87, residues 180–186, and the catalytic domain. Molecular dynamics simulations also indicated that the structure of CnThrDH changed from a closed form to an open form upon NAD+ binding. The newly formed cleft in the open form may function as a conduit for substrate entry and product exit. These computational results led us to hypothesize that the CnThrDH reaction progresses by switching between the closed and open forms. Enzyme kinetics parameters of the L80G, G184A, and T186N variants also supported this prediction: the kcat/Km, l-Thr value of the variants was >330-fold lower than that of the wild type; this decrease suggested that the variants mostly adopt the open form when l-Thr is bound to the active site. These results are summarized in a schematic model of the stepwise changes in flexibility and structure that occur in CnThrDH upon binding of NAD+ and l-Thr. This demonstrates that the dynamical structural changes of short chain dehydrogenase-like l-threonine dehydrogenase are important for the reactivity and specificity of the enzyme.  相似文献   

14.
The gene encoding isocitrate dehydrogenase (IDH) of Methylococcus capsulatus (McIDH) was cloned and overexpressed in Escherichia coli. The purified enzyme was NAD+-dependent with a thermal optimum for activity at 55–60°C and an apparent midpoint melting temperature (T m) of 70°C. Analytical ultracentrifugation (AUC) revealed a homotetrameric state, and McIDH thus represents the first homotetrameric NAD+-dependent IDH that has been characterized. Based on a structural alignment of McIDH and homotetrameric homoisocitrate dehydrogenase (HDH) from Thermus thermophilus (TtHDH), we identified the clasp-like domain of McIDH as a likely site for tetramerization. McIDH showed moreover, higher sequence identity (48%) to TtHDH than to previously characterized IDHs. Putative NAD+-IDHs with high sequence identity (48–57%) to McIDH were however identified in a variety of bacteria showing that NAD+-dependent IDHs are indeed widespread within the domain, Bacteria. Phylogenetic analysis including these new sequences revealed a close relationship with eukaryal allosterically regulated NAD+-IDH and the subfamily III of IDH was redefined to include bacterial NAD+- and NADP+-dependent IDHs. This apparent relationship suggests that the mitochondrial genes encoding NAD+-IDH are derived from the McIDH-like IDHs.  相似文献   

15.
A 50-ns molecular dynamics simulation has been used to study the homotetramer of the enzyme glycosomal glyceraldehyde 3-phosphate dehydrogenase (gGAPDH) complexes, from Trypanosoma cruzi, with nicotinamide adenine dinucleotide (NAD+) cofactors in aqueous solution. The root mean square deviation indicates that the overall structure of the homotetramer does not undergo significant change. The largest structural change observed was in the NAD+ binding domain of subunit (chain) D; as a consequence, the NAD+ cofactor was dislocated from its initial position. However, the other subunits were not affected, suggesting that the gGAPDH enzyme exhibits non-cooperative behaviour. Our simulation estimates that the NAD+ binding domain rotates about 4.8° relative to the catalytic domain in the apo–holo form transition. The hydrogen bond analysis reveals that the residues R12, I13, D38 and M39 are essential for gGAPDH–NAD+ interaction. Furthermore, two promising cavities to be explored in drug design were found: one formed by residues I13, R12, T197, T199, E336 and Y339, and the other by residues C166, H194, R249, I13, R12, T197, T199, E336 and Y339. The results presented in this paper offer new insight into the search for inhibitors of the gGAPDH enzyme of T. cruzi protozoan.  相似文献   

16.
Aldehyde dehydrogenases are found in all organisms and play an important role in the metabolic conversion and detoxification of endogenous and exogenous aldehydes. Genomes of many organisms including Escherichia coli and Salmonella typhimurium encode two succinate semialdehyde dehydrogenases with low sequence similarity and different cofactor preference (YneI and GabD). Here, we present the crystal structure and biochemical characterization of the NAD(P)+‐dependent succinate semialdehyde dehydrogenase YneI from S. typhimurium. This enzyme shows high activity and affinity toward succinate semialdehyde and exhibits substrate inhibition at concentrations of SSA higher than 0.1 mM. YneI can use both NAD+ and NADP+ as cofactors, although affinity to NAD+ is 10 times higher. High resolution crystal structures of YneI were solved in a free state (1.85 Å) and in complex with NAD+ (1.90 Å) revealing a two domain protein with the active site located in the interdomain interface. The NAD+ molecule is bound in the long channel with its nicotinamide ring positioned close to the side chain of the catalytic Cys268. Site‐directed mutagenesis demonstrated that this residue, as well as the conserved Trp136, Glu365, and Asp426 are important for activity of YneI, and that the conserved Lys160 contributes to the enzyme preference to NAD+. Our work has provided further insight into the molecular mechanisms of substrate selectivity and activity of succinate semialdehyde dehydrogenases. © 2012 Wiley Periodicals, Inc.  相似文献   

17.
Two commercially important enzymes, L-lactate dehydrogenase (LDH) and L-malate dehydrogenase (MDH) were purified simultaneously from bovine heart, on an agarose affinity adsorbent. This adsorbent bears a dye-ligand composed of an anthraquinone chlorotriazine chromophore linked to a biomimetic terminal 4-aminophenyloxanylic acid moiety. The purification protocol exploited the biomimetic affinity adsorbent, in combination with a cross-linked agarose DEAE anion-exchanger. The procedure comprised a preliminary anion-exchange first step, for the separation of the three enzyme activities, mMDH, cMDH and LDH. In the second step, that of affinity chromatography, the unbound mMDH obtained from the first step, was purified by specific elution with NAD+/sulphite (22.5-fold purification, 55% step-yield). The procedure afforded mMDH preparation of specific activity approx. 1,300?u/mg (25?°C) at 45% overall yield, free of cytoplasmic MDH, glutamic-oxaloacetic transaminase (GOT) and fumarase. The LDH activity, which, bound to the anion-exchanger during the first step, was recovered from the adsorbent in 200?mM KCl, and finally purified by biomimetic-dye affinity chromatography (NAD+/sulphite elution) and a second ion-exchange chromatography step (elution with 200?mM KCl). The LDH preparation exhibited specific activity approx. 500?u/mg at 25?°C (content of impurities: pyruvate kinase and GOT were not detected; MDH, 0.01%).  相似文献   

18.

Objectives

To investigate the efficiency of a cofactor regeneration enzyme co-expressed with a glycerol dehydrogenase for the production of 1,3-dihydroxyacetone (DHA).

Results

In vitro biotransformation of glycerol was achieved with the cell-free extracts containing recombinant GlyDH (glycerol dehydrogenase from Escherichia coli), LDH (lactate dehydrogenase form Bacillus subtilis) or LpNox1 (NADH oxidase from Lactobacillus pentosus), giving DHA at 1.3 g l?1 (GlyDH/LDH) and 2.2 g l?1 (GlyDH/LpNox1) with total turnover number (TTN) of NAD+ recycling of 6039 and 11100, respectively. Whole cells of E. coli (GlyDH–LpNox1) co-expressing both GlyDH and LpNox1 were constructed and converted 10 g glycerol l?1 to DHA at 0.2–0.5 g l?1 in the presence of zero to 2 mM exogenous NAD+. The cell free extract of E. coli (GlyDH–LpNox) converted glycerol (2–50 g l?1) to DHA from 0.5 to 4.0 g l?1 (8–25 % conversion) without exogenous NAD+.

Conclusions

The disadvantage of the expensive consumption of NAD+ for the production of DHA has been overcome.
  相似文献   

19.
20.
Three isoenzymes of malate dehydrogenase have been isolated from 9-day-old wheat shoots. The microbody (peroxisome) and chloroplast MDH are similar in their electrophoretic behaviour. The mitochondrial MDH, soluble MDH and chloroplast MDH differ in Km values for malate and NAD. The activity of MDH isoenzymes with NAD+-analogues as substrate was in the order 3-AP-NAD+ > 3-AP-deam NAD+ > NAD+ > TN-NAD+ and deam NAD+. The thermal stabilities of the isoenzymes were significantly different: C-MDH > m-MDH > S-MDH.  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