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.
The binding of NAD+, NADH and adenosine diphosphoribose (Ado-PP-Rib) to a stable, highly active and nucleotide-free preparation of rabbit muscle glyceraldehyde-3-phosphate dehydrogenase (D-glyceraldehyde-3-phosphate: NAD+ oxidoreductase (phosphorylating), EC 1.2.1.12) has been studied. All three nucleotides quench the protein fluorescence to the same extent when they bind to the enzyme, and this property has been used to measure the dissociation constants for the two high-affinity binding sites for the nucleotides. The results indicate negative interactions between, or non-identify of, these two binding sites, to which NAD+ and NADH bind with similar affinity. The binding of NAD+ to the enzyme has been studied by spectrophotometric titrations at 360 nm. It appears that the binding of NAD+ to each of the four subunits of the enzyme contributes equally to the intensity of this 'Racker' band. The dissociation constants associated with the binding of the third and fourth molecules of NAD+ estimated from such titrations confirm some previous estimates. The binding of NADH to the enzyme causes a decrease of intensity of the absorbance of the coenzyme at 340 nm, and the dissociation constants for binding of the third and fourth molecules of NADH have been estimated from spectrophotometric titrations. They are the same as those for NAD+. Judging by the apparent dissociation constants, negative interactions on binding the third molecule of NAD+ or NADH are more marked than those associated with the binding of the second and fourth molecules, suggesting that a major conformational change occurs at half-saturation of the tetramer with coenzyme.  相似文献   

2.
The interaction of 3-aminopyridine-adenine dinucleotide, an NAD + 2 analogue which is fluorescent at the pyridine end of the molecule, with rabbit muscle glyceraldehyde-3-phosphate dehydrogenase was investigated. The fluorescence properties of the AAD+ molecule were used to monitor the nicotinamide subsites ou the GPDHase tetramer, the fluorescent aminopyridine moiety of the molecule serving as an intrinsic probe. Although the binding of AAD+ wag found to be negatively co-operative, no conformational changes induced at the nicotinamide subsite upon coenzyme binding were found to be transmitted to neighboring subunits. These findings, in conjunction with our earlier findings and with the observation that different NAD+ analogues which differ in the chemistry of the pyridine moiety bind with different extents of co-operativity, enable us to offer specific roles for the nicotinamide and the adenine subsites in generating the negative co-operativity.It is suggested that the structure of the pyridine moiety of the coenzyme controls the mode of binding of the pyridine moiety to the nicotinamide subsite. This, in turn, controls the orientation of the adenine moiety with respect to its subsite, thereby determining the mode of the interactions between the adenine and its binding domain. As the propagation of conformational changes caused by these interactions to neighboring subunits is believed to be the cause of the negative co-operativity exhibited by this enzyme towards coenzyme binding, the structure of the pyridine moiety controls this phenomenon.  相似文献   

3.
Enoyl acyl carrier protein reductase (ENR) is involved in fatty acid biosynthesis. In Escherichia coli this enzyme is the target for the experimental family of antibacterial agents, the diazaborines, and for triclosan, a broad spectrum antimicrobial agent. Biochemical studies have suggested that the mechanism of diazaborine inhibition is dependent on NAD(+) and not NADH, and resistance of Brassica napus ENR to diazaborines is thought to be due to the replacement of a glycine in the active site of the E. coli enzyme by an alanine at position 138 in the plant homologue. We present here an x-ray analysis of crystals of B. napus ENR A138G grown in the presence of either NAD(+) or NADH and the structures of the corresponding ternary complexes with thienodiazaborine obtained either by soaking the drug into the crystals or by co-crystallization of the mutant with NAD(+) and diazaborine. Analysis of the ENR A138G complex with diazaborine and NAD(+) shows that the site of diazaborine binding is remarkably close to that reported for E. coli ENR. However, the structure of the ternary ENR A138G-NAD(+)-diazaborine complex obtained using co-crystallization reveals a previously unobserved conformational change affecting 11 residues that flank the active site and move closer to the nicotinamide moiety making extensive van der Waals contacts with diazaborine. Considerations of the mode of substrate binding suggest that this conformational change may reflect a structure of ENR that is important in catalysis.  相似文献   

4.
Corrected fluorescence properties of yeast alcohol dehydrogenase and its coenzyme complexes have been investigated as a function of temperature. Dissociation constants have been obtained for binary and ternary complexes of NAD and NADH by following the enhancement of NADH fluorescence or the quenching of the protein fluorescence. It is found that the presence of pyrazole increases the affinity of NAD to the enzyme approximately 100-fold. The formation of the ternary enzyme - NAD - pyrazole complex is accompanied by a large change in the ultraviolet absorption properties, with a new band in the 290-nm region. Significant optical changes also accompany the formation of the ternary enzyme-NADH-acetamide complex. The possible origin for the quenching of the protein fluorescence upon coenzyme binding is discussed, and it is suggested that a coenzyme-induced conformational change can cause it. Thermodynamic parameters associated with NAD and NADH binding have been evaluated on the basis of the change of the dissociation constants with temperature. Optical and thermodynamic properties of binary and ternary complexes of yeast alcohol dehydrogenase are compared with the analogous properties of horse liver alcohol dehydrogenase.  相似文献   

5.
Transhydrogenase (TH) is a dimeric integral membrane enzyme in mitochondria and prokaryotes that couples proton translocation across a membrane with hydride transfer between NAD(H) and NADP(H) in soluble domains. Crystal structures of the NAD(H) binding alpha1 subunit (domain I) of Rhodospirillum rubrum TH have been determined at 1.8 A resolution in the absence of dinucleotide and at 1.9 A resolution with NADH bound. Each structure contains two domain I dimers in the asymmetric unit (AB and CD); the dimers are intimately associated and related by noncrystallographic 2-fold axes. NADH binds to subunits A and D, consistent with the half-of-the-sites reactivity of the enzyme. The conformation of NADH in subunits A and D is very similar; the nicotinamide is in the anti conformation, the A-face is exposed to solvent, and both N7 and O7 participate in hydrogen bonds. Comparison of subunits A and D to six independent copies of the subunit without bound NADH reveals multiple conformations for residues and loops surrounding the NADH site, indicating flexibility for binding and release of the substrate (product). The NADH-bound structure is also compared to the structures of R. rubrum domain I with NAD bound (PDB code 1F8G) and with NAD bound in complex with domain III of TH (PDB code 1HZZ). The NADH- vs NAD-bound domain I structures reveal conformational differences in conserved residues in the NAD(H) binding site and in dinucleotide conformation that are correlated with the net charge, i.e., oxidation state, of the nicotinamides. The comparisons illustrate how nicotinamide oxidation state can affect the domain I conformation, which is relevant to the hydride transfer step of the overall reaction.  相似文献   

6.
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   

7.
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   

8.
The binding of NAD+, NADH, and ADP-ribose to horse liver alcohol dehydrogenase has been studied calorimetrically as a function of pH at 25 degrees C. The enthalpy of NADH binding is 0 +/- 0.5 kcal mol-1 in the pH range 6 to 8.6. The enthalpy of NAD+ binding, however, varies with pH in a sigmoidal fashion and is -4.0 kcal mol(NAD)-1 at pH 6.0 and +4.5 kcal mol(NAD)-1 at pH 8.6 with an apparent pKa of 7.6 +/- 0.2. The enthalpy of proton ionization of the group on the enzyme is calculated to be in the range 8.8 to 9.8 kcal mol(H+)-1. In conjunction with the available thermodynamic data on the ionization of zinc-bound water in model compounds, it is concluded that the group with a pKa of 9.8 in the free enzyme and 7.6 in the enzyme . NAD+ binary complex is, most likely, the zinc-bound water molecule. Our studies with zinc-free enzyme provide further evidence for this conclusion. Therefore, the processes involving a conformational change of the enzyme upon NAD+ binding and the suggested mechanism of subsequent quenching of the fluorescence of Trp-314 implicating the participation of an ionized tyrosine group must be re-evaluated in the light of this thermodynamic study.  相似文献   

9.
Comparison of AMP and NADH binding to glycogen phosphorylase b   总被引:3,自引:0,他引:3  
The binding sites for the allosteric activator, AMP, to glycogen phosphorylase b are described in detail utilizing the more precise knowledge of the native structure obtained from crystallographic restrained least-squares refinement than has hitherto been available. Localized conformational changes are seen at the allosteric effector site that include shifts of between 1 and 2 A for residues Tyr75 and Arg309 and very small shifts for the region of residues 42 to 44 from the symmetry-related subunit. Kinetic studies demonstrate that NADH inhibits the AMP activation of glycogen phosphorylase b. Crystallographic binding studies at 3.5 A resolution show that NADH binds to the same sites on the enzyme as AMP, i.e. the allosteric effector site N, which is close to the subunit-subunit interface, and the nucleoside inhibitor site I, which is some 12 A from the catalytic site. The conformations of NADH at the two sites are different but both conformations are "folded" so that the nicotinamide ring is close (approx. 6 A) to the adenine ring. These conformations are compared with those suggested from solution studies and with the extended conformations observed in the single crystal structure of NAD+ and for NAD bound to dehydrogenases. Possible mechanisms for NADH inhibition of phosphorylase activation are discussed.  相似文献   

10.
J Ovádi  I R Osman  J Batke 《Biochemistry》1982,21(25):6375-6382
Covalent binding of FITC up to 2 mol/mol of tetrameric enzyme does not affect the enzymatic activity and dissociation properties of pig muscle D-glyceraldehyde-3-phosphate dehydrogenase (GAPD). The binding of NAD to dehydrogenase-FITC complex partially reverts the quenching caused by the binding of dye to apo-GAPD. This phenomenon, as well as the formation of a characteristic absorption difference spectrum caused by the binding of NAD, makes it possible to follow the NAD-induced local conformational changes near the dye-binding region. The time course of NAD-induced spectral changes shows biphasic kinetics: a burst and a slow phase. The amplitude of burst phase as a function of NAD equivalents has sigmoidal shape due to the cooperative interaction between subunits. The same conclusion could be drawn from fluorescence anisotropy measurements. In the presence of excess NAD a slow conformational change can be detected, the amplitude of which is a function of NAD concentration. This phenomenon can be attributed to the binding of further NAD molecules to the holoenzyme. The slow phase follows first-order kinetics, and the rate constant depends on enzyme concentration. The specific fluorescence intensity and the fluorescence anisotropy of fluorescent dye labeled apo-GAPD and GAPD saturated with NAD are also dependent on enzyme concentration. We suggest that NAD binding induces major changes in the steric structure of tetrameric enzyme without influencing remarkably the interacting forces between the contact surfaces of subunits. Data are quantitatively interpreted in terms of a two-step dissociation model.  相似文献   

11.
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   

12.
Pre-steady state, stopped flow analysis of Escherichia coli D-3-phosphoglycerate dehydrogenase was performed by following the fluorescence of protein tryptophan and the fluorescence resonance energy transfer from protein tryptophan to bound NADH. The results indicate that binding of substrates is ordered, with coenzyme, NADH, binding first. Furthermore, the analysis indicated that there are two sets of sites on the tetrameric enzyme that can be differentiated by their kinetic behavior. NADH binding was consistent with an initial binding event followed by a slow conformational change for each site. The slow conformational change is responsible for the apparent tight binding of NADH to the apoenzyme but is too slow to participate in the catalytic cycle when the enzyme is rapidly turning over. Subsequent binding of the substrate, alpha-ketoglutarate, was characterized by a rapid equilibrium binding event followed by a conformational change for each site. Catalysis in the direction of NAD(+) reduction showed a distinct burst of activity followed by a slow rate of turnover, indicating that the rate-limiting step is after hydride transfer. Catalysis in the direction of NADH oxidation did not display burst kinetics, indicating that the rate-limiting step is at or before the hydride transfer step. The burst data indicated that the rate of NAD(+) reduction (3.8 s(-1)) is similar to the k(cat) of the enzyme (2-3 s(-1)) in that direction. However, analysis of the reaction with deuterated NADH failed to show an effect on the velocity of the reaction with a V(H)/V(D)=1.07+/-0.06. None of the other rates determined by stopped flow analysis could account for the k(cat) of the enzyme in either direction (forward k(cat)=0.01 s(-1), reverse k(cat)=2-3 s(-1)), suggesting that the rate-limiting step in both directions is a conformational change in the enzyme that is not detected optically.  相似文献   

13.
The crystal structure of the mitochondrial NAD-malic enzyme from Ascaris suum, in a quaternary complex with NADH, tartronate, and magnesium has been determined to 2.0-A resolution. The structure closely resembles the previously determined structure of the same enzyme in binary complex with NAD. However, a significant difference is observed within the coenzyme-binding pocket of the active site with the nicotinamide ring of NADH molecule rotating by 198 degrees over the C-1-N-1 bond into the active site without causing significant movement of the other catalytic residues. The implications of this conformational change in the nicotinamide ring to the catalytic mechanism are discussed. The structure also reveals a binding pocket for the divalent metal ion in the active site and a binding site for tartronate located in a highly positively charged environment within the subunit interface that is distinct from the active site. The tartronate binding site, presumably an allosteric site for the activator fumarate, shows striking similarities and differences with the activator site of the human NAD-malic enzyme that has been reported recently. Thus, the structure provides additional insights into the catalytic as well as the allosteric mechanisms of the enzyme.  相似文献   

14.
The tetrameric glyceraldehyde-3-phosphate dehydrogenase from rabbit muscle binds NAD+ and some of its analogues in a negatively cooperative manner, whereas other NAD+ analogues bind non-cooperatively to this enzyme. Subsequent to alkylation of a fraction of the active sites of the enzyme with the fluorescent SH reagent N-iodoacetyl-N'-(5-sulfo-1-naphthyl)-ethylenediamine, it was found that the alkylated sites bind NAD+ and NAD+ analogues with a markedly reduced affinity as compared with non-alkylated sites. It was therefore feasible to measure the fluorescence and the circular polarization of the luminescence of the enzyme-bound alkyl groups as a function of binding of NAD+ and of NAD+ analogues to the non-alkylated sites. The changes observed indicate that ligand binding to the non-alkylated sites induces changes in the fluorescence properties of the alkyl groups bound to neighbouring subunits, most likely through the protein moiety. The nature of these changes appears to depend on the structure of the coenzyme analogue. The binding of the non-cooperative binders acetyl-pyridine--adenine dinucleotide, ATP and ADP-ribose induce different conformational changes in the neighbouring vacant subunit, as monitored by the spectroscopic properties of the bound alkyl group. These results in conjunction with other data support the view that the negative cooperativity in NAD+ binding to glyceraldehyde-3-phosphate dehydrogenase results from ligand-induced conformational changes. Furthermore, these results further support the view that subtle structural changes in the coenzyme molecule determine the nature of the conformational changes induced within the enzyme tetramer.  相似文献   

15.
The luminescence quenching and conformational behavior of alcohol dehydrogenase from horse liver upon substrate binding has been studied. It was shown that the binding of NADH and NAD+ to the enzyme resulted in the quenching of Trp-314 luminescence, whereas the luminescence of Trp-15 was not quenched. In this case non-radiating energy transfer from Trp-314 to NADH was observed. An essential energy transfer from Trp-15 to NADH and between the two Trp-314 of both subunits of the enzyme was not revealed. The quenching of the enzyme luminescence upon NAD+ binding was, mainly, caused by NAD+ reduction up to NADH. It was assumed, that the release of the proton upon NAD+ binding occurred due to the reduction. Binding of ethanol, ADP or adenosine did not result in essential conformational changes of the enzyme.  相似文献   

16.
The effect of NAD on the binding of 1-anilino-8-naphthalene sulfonate (ANS) to yeast glyceraldehyde-3-phosphate dehydrogenase has been studied using difference spectrophotometric and fluorescence techniques. Coenzyme addition causes the displacement of ANS from its complex with the dehydrogenase, as suggested by the effect of NAD on the fluorescence of the enzyme--ANS complex, as well as on the magnitude of the difference spectrum of the complex. Adenine containing NAD fragments, adenosine, 5'-AMP, and ADP were shown to compete with ANS for the common site on the enzyme using fluorimetric technique; in the case of adenosine and 5'-AMP a direct method of analytical ultracentrifugation was also employed. The results obtained by both methods suggest the dye binding at the adenine subsite of the dehydrogenase. The interaction with ANS causes no detectable conformational changes of the protein. The fluorescence of the dye-enzyme complex increases and the emission maximum shifts to shorter wavelengths on addition of nicotinamide mononucleotide. This suggest some conformational changes to occur in the microenvironment of the bound dye in response to the interaction with the ligand in the nicotinamide subsite. The participation of the nicotinamide subsite of the active center in determining the character of conformational transitions associated with coenzyme binding to glyceraldehyde-3-phosphate dehydrogenase is discussed.  相似文献   

17.
The mechanism coupling electron transfer and proton pumping in respiratory complex I (NADH-ubiquinone oxidoreductase) has not been established, but it has been suggested that it involves conformational changes. Here, the influence of substrates on the conformation of purified complex I from Escherichia coli was studied by cross-linking and electron microscopy. When a zero-length cross-linking reagent was used, the presence of NAD(P)H, in contrast to that of NAD+, prevented the formation of cross-links between the hydrophilic subunits of the complex, including NuoB, NuoI, and NuoCD. Comparisons using different cross-linkers suggested that NuoB, which is likely to coordinate the key iron-sulfur cluster N2, is the most mobile subunit. The presence of NAD(P)H led also to enhanced proteolysis of subunit NuoG. These data indicate that upon NAD(P)H binding, the peripheral arm of the complex adopts a more open conformation, with increased distances between subunits. Single particle analysis showed the nature of this conformational change. The enzyme retains its L-shape in the presence of NADH, but exhibits a significantly more open or expanded structure both in the peripheral arm and, unexpectedly, in the membrane domain also.  相似文献   

18.
Supernatant malate dehydrogenase from pig heart, a dimeric protein containing two very similar or identical subunits, shows negatively cooperative (anticooperative) interactions between NADH binding sites in the presence, but not in the absence, of 0.1 M L-malate. This behavior is observed consitently whether the technique used employs protein fluorescence quenching, NADH fluorescence enhancement, or ultrafiltration dialysis. Fluorescence titration shows that L-malate is also anticooperatively bound in the presence of saturating concentrations of NADH. The data are consistent with an "induced asymmetry" model in which conformational change accompanies the formation of the ternary complex. Two of the three chromatographically resolvable forms of the enzyme have been tested and found to have anticooperative behavior.  相似文献   

19.
Karsten WE  Cook PF 《Biochemistry》2007,46(50):14578-14588
The NAD-malic enzyme catalyzes the oxidative decarboxylation of l-malate. Structures of the enzyme indicate that arginine 181 (R181) is within hydrogen bonding distance of the 1-carboxylate of malate in the active site of the enzyme and interacts with the carboxamide side chain of the nicotinamide ring of NADH, but not with NAD+. Data suggested R181 might play a central role in binding and catalysis in malic enzyme, and it was thus changed to lysine and glutamine to probe its potential function. A nearly 100-fold increase in the Km for malate and a 30-fold increase in the Ki for oxalate, an analogue of the enolpyruvate intermediate, in the R181Q and R181K mutants are consistent with a role for R181 in binding substrates. The mutant enzymes also exhibit a >10-fold increase in KiNADH, but only a slight or no change in KNAD, consistent with rotation of the nicotinamide ring into the malate binding site upon reduction of NAD+ to NADH. The activity of the R181Q mutant can be rescued by ammonium ion likely by binding in the pocket vacated by the guanidinium group of R181. Results suggest 2 mol of ammonia bind per mole of active sites with a high-affinity KNH4 of 0.7 +/- 0.1 mM and a low-affinity KNH4 of approximately 420 mM. Occupancy of the high-affinity site, likely by NH4+, results in an increase in the affinity of malate, oxalate, and NADH (with no change in NAD affinity), consistent with the above-proposed roles for R181. The second molecule to bind is likely neutral NH3, and its binding increases V/Et approximately 20-fold. Primary deuterium and 13C isotope effects measured in the absence and presence of ammonium ion suggest R181Q predominantly affects the rate of the reaction by changing the rate of the precatalytic conformational change. The isotope effects do not change upon binding the second mole of ammonia in spite of the 20-fold increase in V/Et. Thus, the R181Q mutant enzyme exists as an equilibrium mixture between active and less active forms, and NH3 stabilizes the more active conformation of the enzyme.  相似文献   

20.
Steady-state kinetic studies including initial velocity for mannitol oxidation and fructose reduction and product inhibition for mannitol oxidation using fructose and reduced nicotinamide adenine dinucleotide (NADH) are in accord with a reaction mechanism best described as ordered Bi-Bi with NAD+ and NADH designated as the first substrate, last product, respectively at pH 8.8. All replots of slopes and intercepts from product inhibition studies were linear. Dead-end inhibition studies using mannitol 1-phosphate gave slope-parabolic, intercept-linear noncompetitive inhibition for both NAD+ and mannitol as substrates. The dead-end inhibitor is capable of binding multiply to the E, EA, and EQ forms of the enzyme to an extent that is controlled by the concentration of substrates. The EQ complex is inferred to undergo a conformational change, E'Q equilibrium EQ, since (V1/E1) greater than (KiqV2)/(KqE1), and no evidence for dead-end complex formation with NADH can be adduced. This is interpreted to mean that the release of fructose from the central complex is faster than the isomerization of the E-NADH complex. When mannitol is saturating, the noncompetitive inhibition against NAD+, as the variable substrate, becomes parabolic uncompetitive. A replot of the slopes of the parabola against mannitol 1-phosphate remains concave upward. This situation could arise if the conformational change we infer in the EQ complex opens up additional sites on the protein which can interact with the dead-end inhibitor.  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