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.
Ultraviolet resonance Raman (UVRR) spectra, with 260-nm excitation, are reported for oxidized and reduced nicotinamide adenine dinucleotides (NAD+ and NADH, respectively). Corresponding spectra are reported for these coenzymes when bound to the enzymes glyceraldehyde-3-phosphate dehydrogenase (GAPDH) and liver and yeast alcohol dehydrogenases (LADH and YADH). The observed differences between the coenzyme spectra are interpreted in terms of conformation, hydrogen bonding, and general environment polarity differences between bound and free coenzymes and between coenzymes bound to different enzymes. The possibility of adenine protonation is discussed. UVRR spectra with 220-nm excitation also are reported for holo- and apo-GAPDH (GAPDH-NAD+ and GAPDH alone, respectively). In contrast with the 260-nm spectra, these show only bands due to vibrations of aromatic amino acid residues of the protein. The binding of coenzyme to GAPDH has no significant effect on the aromatic amino acid bands observed. This result is discussed in the light of the known structural change of GAPDH on binding coenzyme. Finally, UVRR spectra with 240-nm excitation are reported for GAPDH and an enzyme-substrate intermediate of GAPDH. Perturbations are reported for tyrosine and tryptophan bands on forming the acyl enzyme.  相似文献   

2.
H Deng  J Zheng  D Sloan  J Burgner  R Callender 《Biochemistry》1989,28(4):1525-1533
The binding of the coenzymes NAD+ and NADH to lactate dehydrogenase causes significant changes in the Raman spectra of both of these molecules relative to spectra obtained in the absence of enzyme. The molecular motions of the bound adenine moiety of both NAD+ and NADH as well as adenine containing analogues of these coenzymes produce Raman bands that are essentially identical, suggesting that the binding of adenine to the enzyme is the same regardless of the nicotinamide head-group nature. We also have observed that the molecular motions of the bound adenine moiety are different from both those obtained when it is in either water, various hydrophobic solvents, or various other solvent compositions. Protonation of the bound adenine ring at the 3-position is offered as a possible explanation. Significant shifts are observed in both the stretching frequency of the carboxamide carbonyl of NAD+ and the rocking motion of the carboxamide NH2 group of NADH. These shifts are probably caused by hydrogen bonding with the enzyme. The interaction energies of these hydrogen-bonding patterns are discussed. The aromatic nature of the nicotinamide moiety of NAD+ appears to be unchanged upon binding. Pronounced changes in the Raman spectrum of the nicotinamide moiety of NADH are observed upon binding; some of these changes are understood and discussed. Finally, these results are compared to analogous results that were recently reported for liver alcohol dehydrogenase [Chen et al. (1987) Biochemistry 26, 4776-4784]. In general, the coenzyme binding properties are found to be quite similar, but not identical, for the two enzymes.  相似文献   

3.
The synergism between coenzyme and anion binding to liver alcohol dehydrogenase has been examined by equilibrium measurements and transient-state kinetic methods to characterize electrostatic interactions of coenzymes with ligands which are bound to the catalytic zinc ion of the enzyme subunit. Inorganic anions typically exhibit an at least 200-fold higher affinity for the general anion-binding site than for catalytic zinc on complex formation with free enzyme. Acetate and SCN- interact more strongly with catalytic zinc in the enzyme X NAD+ complex than with the general anion-binding site in free enzyme. CN- shows no significant affinity for the general anion-binding site, but combines to catalytic zinc in the absence as well as the presence of coenzymes. Coordination of CN- to catalytic zinc weakens the binding of NADH by a factor of 50, and tightens the binding of NAD+ to approximately the same extent through interactions which do not include any contributions from covalent adduct formation between CN- and NAD+. These observations provide unambiguous information about the magnitude of electrostatic field effects of coenzymes on anion (e.g. hydroxyl ion) binding to catalytic zinc. They lead to the important inference that coenzyme binding must be strongly affected by ionization of zinc-bound water irrespective of the actual acidity of the latter group. It is concluded on such grounds that the much debated pH dependence of coenzyme binding to liver alcohol dehydrogenase must derive from ionization of zinc-bound water. The assumption that such is not the case leads to the inference that there is no detectable effect of ionization of zinc-bound water on coenzyme binding over the pH range 6-12, a possibility which is definitely excluded by the present results.  相似文献   

4.
Three crystal structures have been determined of active site specific substituted Cd(II) horse liver alcohol dehydrogenase and its complexes. Intensities were collected for the free, orthorhombic enzyme to 2.4-A resolution and for a triclinic binary complex with NADH to 2.7-A resolution. A ternary complex was crystallized from an equilibrium mixture of NAD+ and p-bromobenzyl alcohol. The microspectrophotometric analysis of these single crystals showed the protein-bound coenzyme to be largely NADH, which proves the complex to consist of CdII-LADH, NADH, and p-bromobenzyl alcohol. Intensity data for this abortive ternary complex were collected to 2.9-A resolution. The coordination geometry in the free Cd(II)-substituted enzyme is highly similar to that of the native enzyme. Cd(II) is bound to Cys-46, Cys-174, His-67, and a water molecule in a distorted tetrahedral geometry. Binding of coenzymes induces a conformational change similar to that in the native enzyme. The interactions between the coenzyme and the protein in the binary and ternary complexes are highly similar to those in the native ternary complexes. The substrate binds directly to the cadmium ion in a distorted tetrahedral geometry. No large, significant structural changes compared to the native ternary complex with coenzyme and p-bromobenzyl alcohol were found. The implications of these results for the use of active site specific Cd(II)-substituted horse liver alcohol dehydrogenase as a model system for the native enzyme are discussed.  相似文献   

5.
The binding of oxidized and reduced coenzyme (NAD+ and NADH) to 3-phosphoglyceroyl-glyceraldehyde-3-phosphate dehydrogenase has been studied spectrophotometrically and fluorimetrically. The binding of NAD+ to the acylated sturgeon enzyme is characterized by a significant quenching of the enzyme fluorescence (about 25%) and the induction of a difference spectrum in the ultraviolet absorbance region of the enzyme. Both of these spectroscopic properties are quantitatively distinguishable from those of the corresponding binary enzyme-NAD+ complex. Binding isotherms estimated by gel filtration of the acylated enzyme are in close agreement to those obtained by spectrophotometric and fluorimetric titrations. Up to four NAD+ molecules are bound to the enzyme tetramer. No anticooperativity can be detected in the binding of oxidized coenzyme, which is well described on the basis of a single class of four binding sites with a dissociation constant of 25 muM at 10 degrees C, pH 7.0. The binding of NADH to the acylenzyme has been characterized spectrophotometrically. The absorption band of the dihydronicotinamide moiety of the coenzyme is blue-shifted to 335 nm with respect to free NADH. In addition, a large hypochromicity (23%) is observed together with a significant increase of the bandwidth at half height of this absorption band. This last property is specific to the acylenzyme-DADH complex, since it disappears upon arsenolysis of the acylenzyme. The binding affinity of NADH to the acylated enzyme has been estimated by performing simultaneous spectrophotometric and fluorimetric titrations of the NADH appearance upon addition of NAD+ to a mixture of enzyme and excess glyceraldehyde 3-phosphate. In contrast to NAD+, the reduced coenzyme NADH appears to be relatively strongly bound to the acylated enzyme, the dissociation constant of the acylenzyme-NADH complex being estimated as 2.0 muM at 25 degrees C. In addition a large quenching of the NADH fluorescence (about 83%) is observed. The comparison of the dissociation constants of the coenzyme-acylenzyme complexes and the corresponding Michaelis constants suggests a reaction mechanism of the enzyme in which significant formation and dissociation of NAD+-acylenzyme and NADH-acylenzyme complexes occur. Under physiological conditions the activity of the enzyme can be regulated by the ratio of oxidized and reduced coenzymes. Possible reasons for the lack of anticooperativity in coenzyme binding to the acylated form of the enzyme are discussed.  相似文献   

6.
Yeast alcohol dehydrogenase (YADH) with its cofactor nicotinamide adenine dinucleotide (NAD+) could be stably encapsulated in liposomes composed of POPC (1-palmitoyl-2-oleoyl-sn-glycero-3- phosphocholine). The YADH- and NAD+-containing liposomes (YADH-NADL) were 100 nm in mean diameter. The liposomal YADH and NAD+ concentrations were 2.3 mg/mL and 3.9 mM, respectively. A synergistic effect of the liposomal encapsulation and the presence of NAD+ was examined on the thermal stability of YADH at 45 and 50 degrees C. The enzyme stability of the YADH-NADL was compared to the stabilities of the liposomal YADH (YADHL) containing 3.3 mg/mL YADH without NAD+ as well as the free YADH with and without NAD+. Free YADH was increasingly deactivated during its incubation at 45 degrees C for 2 h with decrease of the enzyme concentration from 3.3 to 0.01 mg/mL because of the dissociation of tetrameric YADH into its subunits. At that temperature, the coexistence of free NAD+ at 3.9 mM improved the stability of free YADH at 2.3 mg/mL through forming their thermostable complex, although the stabilization effect of NAD+ was lowered at 50 degrees C. The turbidity measurements for the above free YADH solution with and without NAD+ revealed that the change in the enzyme tertiary structure was much more pronounced at 50 degrees C than at 45 degrees C even in the presence of NAD+. This suggests that YADH was readily deactivated in free solution due to a decrease in the inherent affinity of YADH with NAD+. On the other hand, both liposomal enzyme systems, YADH-NADL and YADHL, showed stabilities at both 45 and 50 degrees C much higher than those of the above free enzyme systems, YADH/NAD+ and YADH. These results imply that the liposome membranes stabilized the enzyme tertiary and thus quaternary structures. Furthermore, the enzyme activity of the YADH-NADL showed a stability higher than that of the YADHL with a more remarkable effect of NAD+ at 50 degrees C than at 45 degrees C. This was considered to be because even at 50 degrees C the stabilization effect of lipid membranes on the tertiary and quaternary structures of the liposomal YADH allowed the enzyme to form its thermostable complex with NAD+ in liposomes.  相似文献   

7.
Initial-rate studies were made of the oxidation of L-glutamate by NAD+ and NADP+ catalysed by highly purified preparations of dogfish liver glutamate dehydrogenase. With NAD+ as coenzyme the kinetics show the same features of coenzyme activation as seen with the bovine liver enzyme [Engel & Dalziel (1969) Biochem. J. 115, 621--631]. With NADP+ as coenzyme, initial rates are much slower than with NAD+, and Lineweaver--Burk plots are linear over extended ranges of substrate and coenzyme concentration. Stopped-flow studies with NADP+ as coenzyme give no evidence for the accumulation of significant concentrations of NADPH-containing complexes with the enzyme in the steady state. Protection studies against inactivation by pyridoxal 5'-phosphate indicate that NAD+ and NADP+ give the same degree of protection in the presence of sodium glutarate. The results are used to deduce information about the mechanism of glutamate oxidation by the enzyme. Initial-rate studies of the reductive amination of 2-oxoglutarate by NADH and NADPH catalysed by dogfish liver glutamate dehydrogenase showed that the kinetic features of the reaction are very similar with both coenzymes, but reactions with NADH are much faster. The data show that a number of possible mechanisms for the reaction may be discarded, including the compulsory mechanism (previously proposed for the enzyme) in which the sequence of binding is NAD(P)H, NH4+ and 2-oxoglutarate. The kinetic data suggest either a rapid-equilibrium random mechanism or the compulsory mechanism with the binding sequence NH4+, NAD(P)H, 2-oxoglutarate. However, binding studies and protection studies indicate that coenzyme and 2-oxoglutarate do bind to the free enzyme.  相似文献   

8.
Ligand binding and stabilization of malate- and lactate dehydrogenase   总被引:1,自引:0,他引:1  
Binding of coenzymes, coenzyme fragments and phenolate ligands to malate- and lactate dehydrogenase was studied. From linear competition in titration experiments, the coenzyme binding site was concluded to bind all the ligands employed. The analogy between the phenolate ligands and tetraiodofluorescein which is known to bind at the adenosine binding site suggests binding of phenolates at this site. Coenzymes and coenzyme fragments retard the irreversible thermal inactivation of the enzymes. The retardation effect decreases in the order NADH greater than NAD greater than ADPR greater than or equal to AMP for both enzymes. Phenolate ligands binding to the adenosine pocket do not stabilize the enzymes. The stabilization is concluded to originate from the interaction of coenzyme phosphate and nicotinamide with the enzymes. The interactions with the adenosine moiety and with the second ribose seem to be ineffective in retardation of thermal denaturation.  相似文献   

9.
D Chen  K T Yue  C Martin  K W Rhee  D Sloan  R Callender 《Biochemistry》1987,26(15):4776-4784
We report the Raman spectra of reduced and oxidized nicotinamide adenine dinucleotide (NADH and NAD+, respectively) and adenosine 5'-diphosphate ribose (ADPR) when bound to the coenzyme site of liver alcohol dehydrogenase (LADH). The bound NADH spectrum is calculated by taking the classical Raman difference spectrum of the binary complex, LADH/NADH, with that of LADH. We have investigated how the bound NADH spectrum is affected when the ternary complexes with inhibitors are formed with dimethyl sulfoxide (Me2SO) or isobutyramide (IBA), i.e., LADH/NADH/Me2SO or LADH/NADH/IBA. Similarly, the difference spectra of LADH/NAD+/pyrazole or LADH/ADPR with LADH are calculated. The magnitude of these difference spectra is on the order of a few percent of the protein Raman spectrum. We report and discuss the experimental configuration and control procedures we use in reliably calculating such small difference signals. These sensitive difference techniques could be applied to a large number of problems where the classical Raman spectrum of a "small" molecule, like adenine, bound to the active site of a protein is of interest. The spectrum of bound ADPR allows an assignment of the bands of the bound NADH and NAD+ spectra to normal coordinates located primarily on either the nicotinamide or the adenine moiety. By comparing the spectra of the bound coenzymes with model compound data and through the use of deuterated compounds, we confirm and characterize how the adenine moiety is involved in coenzyme binding and discuss the validity of the suggestion that the adenine ring is protonated upon binding. The nicotinamide moiety of NADH shows significant molecular changes upon binding.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

10.
The formation of binary complexes between sturgeon apoglyceralddhyde-3-phosphate dehydrogenase, coenzymes (NAD+ and NADH) and substrates (phosphate, glyceraldehyde 3-phosphate and 1,3-bisphosphoglycerate) has been studied spectrophotometrically and spectrofluorometrica-ly. Coenzyme binding to the apoenzyme can be characterized by several distinct spectroscopic properties: (a) the low intensity absorption band centered at 360 nm which is specific of NAD+ binding (Racker band); (b) the quenching of the enzyme fluorescence upon coenzyme binding; (c) the quenching of the fluorescence of the dihydronicotinamide moiety of the reduced coenzyme (NADH); (D) the hypochromicity and the red shift of the absorption band of NADH centered at 338 nm; (e) the coenzyme-induced difference spectra in the enzyme absorbance region. The analysis of these spectroscopic properties shows that up to four molecules of coenzyme are bound per molecule of enzyme tetramer. In every case, each successively bound coenzyme molecule contributes identically to the total observed change. Two classes of binding sites are apparent at lower temperatures for NAD+ Binding. Similarly, the binding of NADH seems to involve two distinct classes of binding sites. The excitation fluorescence spectra of NADH in the binary complex shows a component centered at 260 nm as in aqueous solution. This is consistent with a "folded" conformation of the reduced coenzyme in the binary complex, contradictory to crystallographic results. Possible reasons for this discrepancy are discussed. Binding of phosphorylated substrates and orthophosphate induce similar difference spectra in the enzyme absorbance region. No anticooperativity is detectable in the binding of glyceraldehyde 3-phosphate. These results are discussed in light of recent crystallographic studies on glyceraldehyde-3-phosphate dehydrogenases.  相似文献   

11.
The binding of NADH and NAD+ by cytoplasmic aldehyde dehydrogenase was studied by various direct and indirect methods. At pH 7.0 at 25 degrees C there appears to be approx. 1 binding site for both nucleotides per 200 000 daltons of protein, although the NAD+-binding results are rather uncertain. Estimates of the dissociation constants of the E . NADH and E . NAD+ complexes under the stated conditions are also presented. Preparations of enzyme are sometimes found to contain significant amounts of very tightly bound NAD+ and NADH. The implications of these findings are discussed.  相似文献   

12.
1. The binding parameters for NADH and NAD+ to rabbit-muscle glyceraldehyde-phosphate dehydrogenase (D-glyceraldehyde-3-phosphate:NAD+ oxidoreductase (phosphorylating), EC 1.2.1.12) have been measured by quenching of the flourescence of the protein and the NADH. 2. The fact that the degree of protein fluorescence quenching by bound NAD+ or NADH, excited at 285 nm and measured at 340 nm ('blue' tryptophans), is not linearly related to the saturation functions of these nucleotides, leads to a slight overestimation of the interaction energy and an underestimation of the concentration of sites, if linearity is assumed. 3. This is also the case for NADH, but not for NAD+, when the protein fluorescence is excited at 305 nm and measured at 390 nm ('red' tryptophans). 4. The binding of NAD+ can be described by a model in which the binding of NAD+, via negative interactions within the dimer, induces weaker binding sites, with the result that the microscopic dissociation constant is 0.08 microM at low saturation and 0.18 microM for the holoenzyme. 5. The binding of NADH can be described on the basis of the same model, the dissociation constant at low saturation being 0.5 microM and of the holoenzyme 1.0 microM. 6. The fluorescence of bound NADH is not sensitive to the conformational changes that cause the decrease in affinity of bound NAD+ or NADH. 7. The binding of NAD+ to the 3-phosphoglyceroyl enzyme can be described by a dissociation constant that is at least two orders of magnitude greater than the dissociation constants of the unacylated enzyme. The affinity of NAD+ to this form of the enzyme is in agreement with the Ki calculated from product inhibition by NAD+ of the reductive dephosphorylation of 1,3-diphosphoglycerate.  相似文献   

13.
As shown by X-ray crystallography, horse liver alcohol dehydrogenase undergoes a global conformational change upon binding of NAD+ or NADH, involving a rotation of the catalytic domain relative to the coenzyme binding domain and the closing up of the active site to produce a catalytically efficient enzyme. The conformational change requires a complete coenzyme and is affected by various chemical or mutational substitutions that can increase the catalytic turnover by altering the kinetics of the isomerization and rate of dissociation of coenzymes. The binding of NAD+ is kinetically limited by a unimolecular isomerization (corresponding to the conformational change) that is controlled by deprotonation of the catalytic zinc-water to produce a negatively-charged zinc-hydroxide, which can attract the positively-charged nicotinamide ring. The deprotonation is facilitated by His-51 acting through a hydrogen-bonded network to relay the proton to solvent. Binding of NADH also involves a conformational change, but the rate is very fast. After the enzyme binds NAD+ and closes up, the substrate displaces the hydroxide bound to the catalytic zinc; this exchange may involve a double displacement reaction where the carboxylate group of a glutamate residue first displaces the hydroxide (inverting the tetrahedral coordination of the zinc), and then the exogenous ligand displaces the glutamate. The resulting enzyme-NAD+-alcoholate complex is poised for hydrogen transfer, and small conformational fluctuations may bring the reactants together so that the hydride ion is transferred by quantum mechanical tunneling. In the process, the nicotinamide ring may become puckered, as seen in structures of complexes of the enzyme with NADH. The conformational changes of alcohol dehydrogenase demonstrate the importance of protein dynamics in catalysis.  相似文献   

14.
The influence of aspartate-168 on the proton-donating and -accepting properties of histidine-195 (the active site acid/base catalyst in lactate dehydrogenase) was evaluated by use of site-directed mutagenesis to change the residue to asparagine and to alanine. Despite the fact that asparagine could form a hydrogen bond to histidine while alanine could not, the two mutant enzymes have closely similar catalytic and ligand-binding properties. Both bind pyruvate and its analogue (oxamate) 200 times more weakly than the wild-type enzyme but show little disruption in their binding of lactate and its unreactive analogue, trifluorolactate. Neither mutation alters the binding of coenzymes (NADH and NAD+) or the pK of the histidine-195 residue in the enzyme-coenzyme complex. We conclude that a strong histidine-aspartate interaction is only formed when both coenzyme and substrate are bound. Deletion of the negative charge of aspartate shifts the equilibrium between enzyme-NADH-pyruvate (protonated histidine) and enzyme-NAD+-lactate (unprotonated histidine) toward the latter. In contrast to the wild-type enzyme, the rate of catalysis in both directions in the mutants is limited by a slow hydride ion transfer step.  相似文献   

15.
The substrate binding site of pig mitochondrial malate dehydrogenase was characterized using complexes of the enzyme with the substrate analogue 8-hydroxypyrene-1,3,6-trisulfonate with and without the addition of coenzymes. The rotational mobility of the fluorescent dye within the binding site was examined with the aid of a multi-frequency phase-fluorimeter. Together with absorption, circular dichroism and fluorescence spectroscopy, conformational changes of the substrate binding site could be defined. The dye was generally found to be immobilized in the binding site. Addition of NADH to the binary complex caused strengthening of a hydrogen bond and further loss of mobility, whereas NAD enlarged the space available for motion of the dye with concomitant loss of the hydrogen bridge.  相似文献   

16.
Equilibrium, thermochemical, and time-resolved fluorescence measurements have been carried out in order to compare pig heart lactate dehydrogenase (LDH) and cytoplasmic malate dehydrogenase (MDH). The differences in the thermodynamic parameters for binding of NADH and NAD+ show the same pattern for both enzymes. The stronger binding of NADH is entropy-based, which can be understood as reflecting electrostatic interactions. The tryptophan fluorescence of MDH and LDH differ for the free enzymes and in quenching by NADH. The differences can be accounted for in terms of a single long-lived tryptophan residue present in LDH and not in MDH.  相似文献   

17.
Amino acid racemases inherently catalyze the exchange of alpha-hydrogen of amino acids with deuterium during racemization in 2H2O. When the reactions catalyzed by alanine racemase (EC 5.1.1.1) and L-alanine dehydrogenase (EC 1.4.1.1), which is pro-R specific for the C-4 hydrogen transfer of NADH, are coupled in 2H2O, [4R-2H]NADH is exclusively produced. Similarly, [4S-2H]NADH is made in 2H2O with amino-acid racemase with low substrate specificity (EC 5.1.1.10) and L-leucine dehydrogenase (EC 1.4.1.9), which is pro-S specific. We have established a simple procedure for the in situ analysis of stereospecificity of C-4 hydrogen transfer of NADH by an NAD-dependent dehydrogenase by combination with either of the above two couples of enzymes in the same reaction mixture. When the C-4 hydrogen of NAD+ is fully retained after sufficient incubation, the stereospecificity of hydrogen transfer by a dehydrogenase is the same as that of alanine dehydrogenase or leucine dehydrogenase. However, when the C-4 hydrogen of NAD+ is exchanged with deuterium, the enzyme to be examined shows the different stereospecificity from alanine dehydrogenase or leucine dehydrogenase. Thus, we can readily determine the stereospecificity by 1H NMR measurement without isolation of the coenzymes and products.  相似文献   

18.
P A Tipton  J Peisach 《Biochemistry》1991,30(3):739-744
Mn2+.tartrate dehydrogenase.substrate complexes have been examined by electron spin echo envelope modulation spectroscopy. The occurrence of dipolar interactions between Mn2+ and 2H on [2H]pyruvate and [4-2H]NAD(H) confirms that Mn2+ binds at the enzyme active site. The 2H signal arising from labeled pyruvate was lost if the sample was incubated at room temperature, indicating that the enzyme catalyzes exchange between the pyruvate methyl protons and solvent protons. Mn-133Cs dipolar coupling was also observed, which suggests that the monovalent cation cofactor also binds in the active site. The tartrate analogue oxalate was observed to have a significant effect on the binding of NAD(H). Oxalate appears to constrain the binding of NAD(H) so that the nicotinamide portion of the cofactor is held in close proximity to Mn2+. Spectra of enzyme complexes prepared with (R)-[4-2H]NADH showed a more intense 2H signal than analogous complexes prepared with (S)-[4-2H]NADH, demonstrating that the pro-R position of NADH is closer to Mn2+ than the pro-S position and suggesting that tartrate dehydrogenase is an A-side-specific dehydrogenase. Oxalate also affected Cs+ binding; the intensity of the 133Cs signal increased in the presence of oxalate, which suggest that oxalate facilitates binding of Cs+ to the active site or that Cs+ binds closer to Mn2+ when oxalate is present. In addition to signals from substrates, electron spin echo envelope modulation spectra revealed 14N signals that arose from coordination to Mn2+ by nitrogen-containing ligands from the protein; however, the identity of this ligand or ligands remains obscure.  相似文献   

19.
Starting from 6-chloropurine riboside and NAD+, different reactive analogues of NAD+ have been obtained by introducing diazoniumaryl or aromatic imidoester groups via flexible spacers into the nonfunctional adenine moiety of the coenzyme. The analogues react with different amino-acid residues of dehydrogenases and form stable amidine or azobridges, respectively. After the formation of a ternary complex by the coenzyme, the enzyme and a pseudosubstrate, the reactive spacer is anchored in the vicinity of the active site. Thus, the coenzyme remains covalently attached to the protein even after decomposition of the complex. On addition of substrates the covalently bound coenzyme is converted to the dihydro-form. In enzymatic tests the modified dehydrogenases show 80-90% of the specific activity of the native enzymes, but they need remarkably higher concentrations of free NAD+ to achieve these values. The dihydro-coenzymes can be reoxidized by oxidizing agents like phenazine methosulfate or by a second enzyme system. Various systems for coenzyme regeneration were investigated; the modified enzymes were lactate dehydrogenase from pig heart and alcohol dehydrogenase from horse liver; the auxiliary enzymes were alcohol dehydrogenase from yeast and liver, lactate dehydrogenase from pig heart, glutamate dehydrogenase and alanine dehydrogenase. Lactate dehydrogenase from heart muscle is inhibited by pyruvate. With alanine dehydrogenase as the auxiliary enzyme, the coenzyme is regenerated and the reaction product, pyruvate, is removed. This system succeeds to convert lactate quantitatively to L-alanine. The thermostability of the binary enzyme systems indicates an interaction of covalently bound coenzymes with both dehydrogenases; both binding sites seem to compete for the coenzyme. The comparison of dehydrogenases with different degrees of modifications shows that product formation mainly depends on the amount of incorporated coenzyme.  相似文献   

20.
R S Ehrlich  R F Colman 《Biochemistry》1992,31(49):12524-12531
The coenzyme selectivity of pig heart NAD-dependent and NADP-dependent isocitrate dehydrogenase has been investigated by nuclear magnetic resonance through the use of coenzyme analogues. For both isocitrate dehydrogenases, more than 10-fold lower maximal activity is observed with thionicotinamide adenine dinucleotide [sNAD(P)+] than with NAD(P)+ or acetylpyridine adenine dinucleotide [acNAD-(P)+] as coenzyme. Nuclear Overhauser effect measurements failed to reveal any differences in the adenine-ribose conformations among the enzyme-bound analogues. The 2'-phosphate resonance of the enzyme-bound NADP+ analogues showed the same change in chemical shift observed for the natural coenzyme and revealed the same lack of pH dependence in the range from pH 5.4 to 8.2. NADP-dependent isocitrate dehydrogenase exhibits only small differences in Michaelis constants for the coenzymes with various nicotinamide substituents, reflecting a predominant role for the adenosine moiety in binding. The conformation of the bound nicotinamide-ribose of the natural coenzymes was appreciably different from that of the coenzyme, sNAD(P)+, which shows low catalytic activity. For both isocitrate dehydrogenases, sNAD(P)+ bound to the enzymes exhibits a mixture of syn and anti conformations while only the anti conformation can be detected for NAD(P)+. Chemical shifts of NAD(P)+ enriched with 13C in the carboxamide indicate that interaction of this group with the enzymes may play a role in positioning the nicotinamide ring to participate in catalysis. Our results suggest that, although interaction of the nicotinamide moiety with the enzymes contributes relatively little to the energy of interaction in the binary complex, the enzymes must correctly position this group for the catalytic event.(ABSTRACT TRUNCATED AT 250 WORDS)  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