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 consider the controversial conclusions of the charge at the migrating hydrogen in the transition state of hydride‐transfer reactions from CH‐groups to hydride acceptors. Quantum chemical calculations were performed on elementary organic reactions involving carbenium ions, which can be considered as hydride acceptors. We also discuss the biochemical hydride‐transfer reactions in which the coenzyme NADH‐NAD+ plays an important role. With the calculations and the experimental model systems, an answer is given for the stereospecificity of the hydride transfer. Generally, the hydride transfer occurs via a trigonal pyramidal geometry in which the transferred hydride of the CH‐group is located in the axial position. In the case of the coenzyme NADH‐NAD+, the hydride transfer is coupled with an out‐of‐plane orientation of the carboxamide group of the pyridinium moiety, resulting in an increased stereospecificity. © 2004 Wiley Periodicals, Inc. Int J Quantum Chem, 2005  相似文献   

2.
For the first time, employment of nicotinamide coenzyme NAD analogs has overcome the limitations of NAD in electrochemical regeneration. It has been shown that NAD analogs, APAD and PAAD, were electrochemically reduced more efficiently than original NAD and that the stability of their reduced products was also much higher than NADH.  相似文献   

3.
A methyl substituent at C-5 of the nicotinamide ring is found to confer increased acid stability in a reduced nicotinamide model (5-20 fold) and in a reduced dinucleotide coenzyme (2-3 fold), while retaining reactivity towards hydride transfer.  相似文献   

4.
We have identified multiple reactive configurations (MRCs) of an enzyme-coenzyme complex that have measurably different kinetic properties. In the complex formed between morphinone reductase (MR) and the NADH analogue 1,4,5,6-tetrahydro-NADH (NADH4) the nicotinamide moiety is restrained close to the FMN isoalloxazine ring by hydrogen bonds from Asn-189 and His-186 as determined from the X-ray crystal structure. Molecular dynamic simulations indicate that removal of one of these hydrogen bonds in the N189A MR mutant allows the nicotinamide moiety to occupy a region of configurational space not accessible in wild-type enzyme. Using stopped-flow spectroscopy, we show that reduction of the FMN cofactor by NADH in N189A MR is multiphasic, identifying at least four different reactive configurations of the MR-NADH complex. This contrasts with wild-type MR in which hydride transfer occurs by environmentally coupled tunneling in a single kinetic phase [Pudney et al. J. Am. Chem. Soc. 2006, 128, 14053-14058]. Values for primary and alpha-secondary kinetic isotope effects, and their temperature dependence, for three of the kinetic phases in the N189A MR are consistent with hydride transfer by tunneling. Our analysis enables derivation of mechanistic information concerning different reactive configurations of the same enzyme-coenzyme complex using ensemble stopped-flow methods. Implications for the interpretation from kinetic data of tunneling mechanisms in enzymes are discussed.  相似文献   

5.
Formate dehydrogenase (FDH) has been widely used for the regeneration of the reduced nicotinamide adenine dinucleotide (NADH). To utilize nicotinamide cytosine dinucleotide (NCD) as a non-natural redox cofactor, it remains challenging as NCDH, the reduced form of NCD, has to be efficiently regenerated. Here we demonstrate successful engineering of FDH for NCDH regeneration. Guided by the structural information of FDH from Pseudomonas sp. 101 (pseFDH) and the NAD–pseFDH complex, semi-rational strategies were applied to design mutant libraries and screen for NCD-linked activity. The most active mutant reached a cofactor preference switch from NAD to NCD by 3700-fold. Homology modeling analysis showed that these mutants had reduced cofactor binding pockets and dedicated hydrophobic interactions for NCD. Efficient regeneration of NCDH was implemented by powering an NCD-dependent D -lactate dehydrogenase for stoichiometric and stereospecific reduction of pyruvate to D -lactate at the expense of formate.  相似文献   

6.
Rhodium complexes of the imidazolylidene (C-im) N-heterocyclic carbene (NHC) ligand, C-im-pyH(+), bearing a nicotinamide cation substituent (pyH(+)) have been targeted for ligand-centered uptake and delivery of hydride ion. This work reveals that rhodium(I) complexes such as [Rh(C-im-pyH(+))(COD)X][PF(6)] (1, a: X = Cl, b: X = I) undergo facile C-metalation of the nicotinamide ring to afford rhodium complexes of a novel chelate ligand, C,C'-im-py, with coordinated imidazolylidene (C(im)) and pyridylidene (C(py)) NHC-donors. Seven examples were characterized and include rhodium(III) monomers of the general formula [Rh(C,C'-im-py)L(x)I(2)](z+) (2: z = 1, L = H(2)O or solvent, x = 2; 3, 5, 7: z = 0, L = carboxylate, x = 1) and novel rhodium(II) dimers, the anti/syn-isomers of [Rh(2)(C,C'-im-py)(2)(μOAc)(2)I(2)] (4-anti/syn). The NMR data, backed by DFT calculations, is consistent with attribution of the C,C'-im-py ligand as a bis(carbene) donor. Single crystal X-ray diffraction studies are reported for 2, 3, 4-anti, 4-syn and 7. Consistently, within the each complex, the Rh-C(im) bond length is shorter than the Rh-C(py) bond length, which is the opposite trend to that expected based on simple electronic considerations. It is proposed that intramolecular steric interactions imposed by different rings in the rigid C,C'-im-py chelate ligand dictate the observed Rh-C(NHC) bond lengths. Attempts to add hydride to the C-metalated nicotinamide ring in 3 were unsuccessful. The redox behavior of 3 and 4 and, for comparison, an analogous bis(imidazolylidene)rhodium(III) monomer (8), were characterized by cyclic voltammetry, electron paramagnetic resonance (EPR), and UV-vis spectroelectrochemistry. In 3 and 4, the C-metalated nicotinamide ring is found to exhibit a one-electron reduction process at far lower potential (-2.34 V vs. Fc(+)/Fc in acetonitrile) than the two-electron nicotinamide cation-dihydronicotinamide couple found for the corresponding nonmetalated ring (-1.24 V). The C,C'-ligand is electrochemically silent over a large potential range (from -2.3 V to the anodic solvent limit), thus for both 3 and 4 the first reduction processes are metal-centered. For 4-anti, the cyclic voltammetry and UV-vis spectrochemical results are consistent with a diamagnetic [Rh(I)Rh(II)](2) tetrameric reduction product. Density functional theory (DFT) calculations were used to further probe the uptake of hydride ion by the nicotinamide ring, both before and after C-metalation. It is found that C-metalation significantly decreases the ability of the nicotinamide ring to take up hydride ion, which is attributed to the "carbene-like" character of a C-metalated pyridylidene ring.  相似文献   

7.
Regioselective hydrogenation of the oxidized form of β-nicotinamide adenine dinucleotide (NAD(+)) to the reduced form (NADH) with hydrogen (H(2)) has successfully been achieved in the presence of a catalytic amount of a [C,N] cyclometalated organoiridium complex [Ir(III)(Cp*)(4-(1H-pyrazol-1-yl-κN(2))benzoic acid-κC(3))(H(2)O)](2) SO(4) [1](2)·SO(4) under an atmospheric pressure of H(2) at room temperature in weakly basic water. The structure of the corresponding benzoate complex Ir(III)(Cp*)(4-(1H-pyrazol-1-yl-κN(2))-benzoate-κC(3))(H(2)O) 2 has been revealed by X-ray single-crystal structure analysis. The corresponding iridium hydride complex formed under an atmospheric pressure of H(2) undergoes the 1,4-selective hydrogenation of NAD(+) to form 1,4-NADH. On the other hand, in weakly acidic water the complex 1 was found to catalyze the hydrogen evolution from NADH to produce NAD(+) without photoirradiation at room temperature. NAD(+) exhibited an inhibitory behavior in both catalytic hydrogenation of NAD(+) with H(2) and H(2) evolution from NADH due to the binding of NAD(+) to the catalyst. The overall catalytic mechanism of interconversion between NADH and NAD(+) accompanied by generation and consumption of H(2) was revealed on the basis of the kinetic analysis and detection of the catalytic intermediates.  相似文献   

8.
Hydride transfer from dihydronicotinamide adenine dinucleotide (NADH) analogues, such as 10-methyl-9,10-dihydroacridine (AcrH 2) and its derivatives, 1-benzyl-1,4-dihydronicotinamide (BNAH), and their deuterated compounds, to non-heme oxoiron(IV) complexes such as [(L)Fe (IV)(O)] (2+) (L = N4Py, Bn-TPEN, and TMC) occurs to yield the corresponding NAD (+) analogues and non-heme iron(II) complexes in acetonitrile. Hydride transfer from the NADH analogues to p-chloranil (Cl 4Q) also occurs to produce the corresponding NAD (+) analogues and the hydroquinone anion (Cl 4QH (-)). The logarithms of the observed second-order rate constants (log k H) of hydride transfer from NADH analogues to non-heme oxoiron(IV) complexes are linearly correlated with those of hydride transfer from the same series of NADH analogues to Cl 4Q, including similar kinetic deuterium isotope effects. The log k H values of hydride transfer from NADH analogues to non-heme oxoiron(IV) complexes are also linearly correlated with those of deprotonation of the radical cations of NADH analogues. Such linear correlations indicate that overall hydride-transfer reactions of NADH analogues to both non-heme oxoiron(IV) complexes and Cl 4Q occur via electron transfer from NADH analogues to the oxoiron(IV) complexes, followed by rate-limiting deprotonation from the radical cations of NADH analogues and subsequent rapid electron transfer from the deprotonated radicals to the Fe(III) complexes to yield the corresponding NAD (+) analogues and the Fe(II) complexes. The electron-transfer pathway was accelerated by the presence of perchloric acid, and the resulting radical cations of NADH analogues were detected by electron spin resonance spectroscopy and UV-vis spectrophotometry in the acid-promoted hydride-transfer reactions from NADH analogues to non-heme oxoiron(IV) complexes. This result provides the first direct evidence that a hydride transfer from NADH analogues to non-heme oxoiron(IV) complexes proceeds via an electron-transfer pathway.  相似文献   

9.
脱氢酶电化学生物传感器的研究进展   总被引:1,自引:0,他引:1  
自然界中超过400种脱氢酶使用辅酶-烟酰胺腺嘌呤二核苷酸(NAD+)或烟酰胺腺嘌呤二核苷酸磷酸(NADP+)作为生物催化反应中氢和电子的传递体,因此烟酰胺型辅酶的电化学氧化对构筑此类脱氢酶电化学生物传感器具有重要的意义.本文介绍了还原型辅酶在人工电子媒介体存在下的电化学氧化,以及脱氢酶电化学生物传感器的设计和应用.  相似文献   

10.
The temperature dependence of the primary kinetic isotope effect (KIE), combined temperature-pressure studies of the primary KIE, and studies of the alpha-secondary KIE previously led us to infer that hydride transfer from nicotinamide adenine dinucleotide to flavin mononucleotide in morphinone reductase proceeds via environmentally coupled hydride tunneling. We present here a computational analysis of this hydride transfer reaction using QM/MM molecular dynamics simulations and variational transition-state theory calculations. Our calculated primary and secondary KIEs are in good agreement with the corresponding experimental values. Although the experimentally observed KIE lies below the semiclassical limit, our calculations suggest that approximately 99% of the reaction proceeds via tunneling: this is the first "deep tunneling" reaction observed for hydride transfer. We also show that the dominant tunneling mechanism is controlled by the isotope at the primary rather than the secondary position: with protium in the primary position, large-curvature tunneling dominates, whereas with deuterium in this position, small-curvature tunneling dominates. Also, our study is consistent with tunneling being preceded by reorganization: in the reactant, the rings of the nicotinamide and isoalloxazine moieties are stacked roughly parallel to each other, and as the system moves toward a "tunneling-ready" configuration, the nicotinamide ring rotates to become almost perpendicular to the isoalloxazine ring.  相似文献   

11.
The reduction of phenylacetaldoxime was catalysed by alcohol dehydrogenases in the presence of NAD(P)H yielding finally the primary alcohol via the imine and aldehyde intermediates. This suggests that the hydride of the cofactor NAD(P)H is transferred to the N-atom of the oxime moiety and not to the carbon atom, as usual stated. This reaction represents the first example of a catalytic chemo-promiscuity of alcohol dehydrogenases.  相似文献   

12.
InhA, the enoyl reductase from Mycobacterium tuberculosis, catalyzes the NADH-dependent reduction of trans-2-enoyl-ACPs. In the present work, Raman spectroscopy has been used to identify catalytically relevant changes in the conformation of the nicotinamide ring that occur when NADH binds to InhA. For 4(S)-NADD, there is an 11 cm-1 decrease in the wavenumber of the C4-D stretching band (nuC-D) and a 50% decrease in the width of this band upon binding to InhA. While a similar reduction in line width is observed for the corresponding band arising from 4(R)-NADD, nuC-D for this isomer increases 34 cm-1 upon binding to InhA. These changes in nuC-D indicate that the nicotinamide ring adopts a bound conformation in which the 4(S)C-D bond is in a pseudoaxial orientation. Mutagenesis of F149, a conserved active site residue close to the cofactor, demonstrates that this enzyme-induced modulation in cofactor structure is directly linked to catalysis. In contrast to the wild-type enzyme, Raman spectra of NADD bound to F149A InhA resemble those of NADD in solution. Consequently, F149A is no longer able to optimally position the cofactor for hydride transfer, which correlates with the 30-fold decrease in kcat and 2-fold increase in D(V/KNADH) caused by this mutation. These studies thus substantiate the proposal that hydride transfer is promoted by pseudoaxial positioning of the NADH pro-4S bond, and indicate that catalysis of substrate reduction by InhA results, in part, from correct orientation of the cofactor in the ground state.  相似文献   

13.
A robust redox extraction protocol for quantitative and reproducible metabolite isolation and recovery has been developed for simultaneous measurement of nicotinamide adenine dinucleotide (NAD) and its reduced form, NADH, from Saccharomyces cerevisiae. Following culture in liquid media, yeast cells were harvested by centrifugation and then lysed under nonoxidizing conditions by bead blasting in ice-cold, nitrogen-saturated 50 mM ammonium acetate. To enable protein denaturation, ice cold nitrogen-saturated CH(3)CN/50 mM ammonium acetate (3:1 v/v) was added to the cell lysates. Chloroform extractions were performed on supernatants to remove organic solvent. Samples were lyophilized and resuspended in 50 mM ammonium acetate. NAD and NADH were separated by HPLC and quantified using UV-Vis absorbance detection. NAD and NADH levels were evaluated in yeast grown under normal (2% glucose) and calorie restricted (0.5% glucose) conditions. Results demonstrate that it is possible to perform a single preparation to reliably and robustly quantitate both NAD and NADH contents in the same sample. Robustness of the protocol suggests it will be (i) applicable to quantification of these metabolites in other cell cultures; and (ii) amenable to isotope labeling strategies to determine the relative contribution of specific metabolic pathways to total NAD and NADH levels in cell cultures.  相似文献   

14.
A series of arylthiolated 2,3-ethylenedioxy-1,4-benzoquinones as a coenzyme Q (CoQ) antagonist was tested for inhibition of succinate oxidase and reduced nicotinamide adenine dinucleotide (NADH) oxidase systems in the mitochondrial respiratory chain. The following characteristics were revealed: (1) 2,3-ethylenedioxy, 5-arylthio and 5,6-diarylthio groups were confirmed to be favorable for inhibition of both systems; (2) these analogs were more effective in the succinate oxidase system than in the NADH oxidase system; (3) 4' substituents on the benzene side ring had little effect on inhibitory activity; (4) the acting sites of these analogs had no strict stereospecificity. The reduced minus oxidized difference spectra revealed that these analogs inhibited the succinate oxidase system at the site between succinate and CoQ, and the NADH oxidase system at the site after cytochrome a + a3, suggesting these analogs might act as antagonists of CoQ in the succinate oxidase system. However, 5-(4'-nitrophenylthio)-2,3-ethylenedioxy-1,4-benzoquinone (If) strongly inhibited only the succinate oxidase system at the site after cytochrome a + a3.  相似文献   

15.
Molecular dynamics simulations of the oxidation of benzyl alcohol by horse liver alcohol dehydrogenase (HLADH) have been carried out. The following three states have been studied: HLADH.PhCH(2)OH.NAD(+) (MD1), HLADH.PhCH(2)O(-).NAD(+) (MD2), and HLADH.PhCHO.NADH (MD3). MD1, MD2, and MD3 simulations were carried out on one of the subunits of the dimeric enzyme covered in a 32-A-radius sphere of TIP3P water centered on the active site. The proton produced on ionization of the alcohol when HLADH.PhCH(2)OH.NAD(+) --> HLADH.PhCH(2)O(-).NAD(+) is transferred from the active site to solvent water via a hydrogen bonding network consisting of serine48 hydroxyl, ribose 2'- and 3'-hydroxyl groups, and Hist51. Hydrogen bonding of the 3'OH of ribose to Ile269 carbonyl maintains this proton in position to be transferred to water. Molecular dynamic simulations have been employed to track water1287 from the TIP3 water pool to the active site, thus exhibiting the mode of entrance of water to the active site. With time the water1287 accumulates in two different positions in order to accept the proton from the ribose 3'-OH and from His51. There can be identified two structural substates for proton passage. In the first substate the imidazole Ne2 of His51 is adjacent to the nicotinamide ribose C2'-OH and hydrogen bonding distances for proton transfer through the hydrogen bonded relay series PhCH(2)OH...Ser48-OH...Ribose2'-OH...His51...OH(2) (path 1) average 2.0, 2.0, and 2.1 A and (for His51...OH(2)) minimal distances less or equal to 2.5 A. The structure for path 1 is present 20% of the time span. And in the second substate, there are two possible proton passages: path 1 as before and path 2. Path 2 involves the hydrogen-bonded relay series PhCH(2)OH...Ser48-OH...Ribose2'-OH...Ribose3'-OH...His51.OH(2) with the average bonding distances being 2.0, 2.0, 2.1, and 2.0 A and (for His51...OH(2)) minimal distances less or equal to 2.5 A (20% probability of the time span), respectively. During the molecular dynamics simulation the NAD(+) ribose conformations have stabilized at the C2'-endo-C3'-exo or the C2'-endo conformations. With the C2'-endo conformation the first and second substates are able to persist for different time spans, while with the C2'-endo-C3'-exo conformation the only possible pathway involves the first substate. For both first and second substates the fluctuation of the distances between the ribose-OH protons and N epsilon 2 of His51 imidazole ring is partially contributed by the "windshield wiper" motion of the His51 imidazole ring. Since the imidazole of His-51 contributes only about 10-fold to activity, as estimated from the decrease in activity upon substitution with a Gln, there must be an alternate route for the proton to pass to solvent without going through this histidine. A third pathway involves ribose C3'-OH and Ile-269. In MD2, near attack conformers (NACs) for hydride transfer from PhCH(2)O(-) to NAD(+) represent approximately 60% of E.S conformers. The molecular dynamic study of MD3 at mildly basic pH reveals that reactive ground state conformers (NACs) for hydride transfer from NADH to PhCHO amount to 12 mol % of conformers. In MD3, anisotropic bending of the dihydronicotinamide ring of NADH (average value of alpha(c) = 4.0 degrees and alpha(n) = 0.5 degrees, respectively) is observed.  相似文献   

16.
The reaction rates of 1-(p-substituted benzyl)-1,4-dihydronicotinamide (G-BNAH) with N-benzylphenothiazine radical cation (PTZ(*+)) in acetonitrile were determined. The results show that the reaction rates (k(obs)) decreased from 2.80 x 10(7) to 2.16 x 10(7) M(-1) s(-1) for G = H as the reaction temperature increased from 298 to 318 K. The activation enthalpies of the reactions were estimated according to Eyring equation to give negative values (-3.4 to -2.9 kcal/mol). Investigation of the reaction intermediate shows that the charge-transfer complex (CT-complex) between G-BNAH and PTZ(*+) was formed in front of the hydride transfer from G-BNAH to PTZ(*+). The formation enthalpy of the CT-complex was estimated by using the Benesi-Hildebrand equation to give the values from -6.4 to -6.0 kcal/mol when the substituent G in G-BNAH changes from CH(3)O to Br. Detailed thermodynamic analyses on each elementary step in the possible reaction pathways suggest that the hydride transfer from G-BNAH to PTZ(*+) occurs by a concerted hydride transfer via a CT-complex. The effective charge distribution on the pyridine ring in G-BNAH at the various stages-the reactant G-BNAH, the charge-transfer complex, the transition-state, and the product G-BNA(+)-was estimated by using the method of Hammett-type linear free energy analysis, and the results show that the pyridine ring carries relative effective positive charges of 0.35 in the CT-complex and 0.45 in the transition state, respectively, which indicates that the concerted hydride transfer from G-BNAH to PTZ(*+) was practically performed by the initial charge (-0.35) transfer from G-BNAH to PTZ(*+) and then followed by the transfer of hydrogen atom with partial negative charge (-0.65). It is evident that the present work would be helpful in understanding the nature of the negative temperature effect, especially on the reaction of NADH coenzyme with the drug phenothiazine in vivo.  相似文献   

17.
利用含人工氧还酶体系的粗酶液代替纯酶催化反应,以省去酶分离纯化过程.由苹果酸酶突变体ME-t(MEL310R/Q401C)和非天然辅酶烟酰胺5-氟胞嘧啶二核苷酸(NFCD+)组成的人工氧还酶体系可以催化氧化L-苹果酸生成丙酮酸,并得到非天然辅酶的还原态(NFCDH).利用含人工氧还酶体系的粗酶液催化反应,只得到单一产物丙酮酸,其选择性与纯酶催化的相同.来自粪肠球菌Enterococcus faecalis的NADH氧化酶(NOX)可再生NFCD+.与含NAD+,ME粗酶液和NOX粗酶液的偶联反应体系相比,含NFCD+,ME-t粗酶液和NOX粗酶液的体系获得的丙酮酸产率高9%,而副产物乳酸明显减少.可见人工氧还酶体系使用更方便,且产物选择性更高,有望代替纯酶催化反应.这为降低生物催化剂的成本,扩大生物催化反应的应用提供了一种新的策略.  相似文献   

18.
《Tetrahedron》1986,42(4):941-959
The present work deals with the mechanistic investigations of the hydride transfer reactions concerning the redox couple NAD+/NADH. Based on the theoretical and experimental investigatins of NAD(H) model compounds as 3-carbamoyl pyridinium cations (3-carbamoyl-1,4-dihydropyridine) it was found that the out-of-plane rotation of the carbonyl function controls the stereo-and regiospecificity of the introduced hydride anion. It was found that the hydride anion transfered in the reaction, is always syn-positioned with respect to the carbonyl group. The unique stereoselectivity exhibits a strong coherence with the recent crystallographic 3D-data for the ternary complex of NAD bonded horse liver alcohol dehydrogenase. The results show that the amide group is 30° out of the plane with the carbonyl directed toward the A side. There are observations that the absolute configuration of the introduced chirality in the 3-carbamoyl pyridinium cations selects between the hydride uptake coresponding with the enzymatic A or B specificity.  相似文献   

19.
The interactive behavior of flavin adenine dinucleotide (FAD) with a bare glassy carbon electrode (GCE) and a Ru-modified GCE was investigated. The reduction of FAD at a GCE/ruthenium-modified GCE surface is a quasi-reversible, surface-controlled process, and our data implied that the attachment of FAD onto the surface is caused by nonspecific adsorption instead of covalent linkage, in which the adenine ring of FAD adopts a flat orientation on the GCE surface in neutral and dilute solutions in order to maximize the pi-pi stacking with the carbon surface and reorients to a perpendicular orientation as the surface gets more crowded. FAD desorption during the exchange with nicotinamide adenine dinucleotide (NAD+) is one order of magnitude slower than desorption in the absence of NAD+, which indicates a strong interaction between FAD and NAD+. General knowledge of the interactive behavior of NAD+ on a FAD-adsorbed GCE provides useful information for the design of a modified electrode surface for the generation of NADH from NAD+.  相似文献   

20.
Asymmetric reduction of methyl benzoylformate by a new NADH mimic is reported; depending on the hydride source used to reduce the NAD+ precursor, NADH mimics so obtained lead to an inversion of enantioselectivity, affording either (R)-methyl mandelate in 88% ee or (S)-methyl mandelate in 78% ee.  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