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.
Association and dissociation rate constants obtained by stopped-flow spectroscopy have permitted definition of a kinetic scheme for recombinant human dihydrofolate reductase that correctly predicts full time course kinetics of the enzymatic reaction over a wide range of substrate and product concentrations. The scheme is complex compared with that for the bacterial enzyme and involves branched pathways. It successfully accounts for observed rapid hysteresis preceding steady state and for the nonhyperbolic dependence of steady-state rate on substrate and product concentrations. The major branch point in the catalytic cycle occurs at E.NADP.H4folate because either NADP or H4folate can dissociate from the ternary product complex (koff = 84 s-1 and 46 s-1, respectively). The rate of conversion of enzyme-bound substrates to products is very fast (k = 1360 s-1) and nearly unidirectional (Kequ = 37) so that other steps limit the catalytic rate. At saturating substrate concentrations these steps include release of NADP and H4folate from E.NADP.H4folate and release of products from the two abortive complexes E.NADPH.H4folate (koff = 225 s-1) and E.NADP.H4folate (koff = 4.6 s-1). Since NADP dissociates slowly from E.NADP.H2folate nearly 90% of the enzyme accumulates as this complex at steady state. Nonetheless, the catalytic rate is maintained at 12 s-1 by rapid flux of a small portion of the enzyme through an alternate branch. At physiological concentrations of substrates and products the steady-state rate is limited primarily by the rate of H2folate binding to E.NADPH so that the enzyme is extremely efficient.  相似文献   

2.
We have found previously (Fahien, L.A., Kmiotek, E.H., MacDonald, M. J., Fibich, B., and Mandic, M. (1988) J. Biol. Chem. 263, 10687-10697) that glutamate-malate oxidation can be enhanced by cooperative binding of mitochondrial aspartate aminotransferase and malate dehydrogenase to the alpha-ketoglutarate dehydrogenase complex. The present results demonstrate that glutamate dehydrogenase, which forms binary complexes with these enzymes, adds to this ternary complex and thereby increases binding of the other enzymes. Kinetic evidence for direct transfer of alpha-ketoglutarate and NADH, within these complexes, has been obtained by measuring steady-state rates of E2 when most of the substrate or coenzyme is bound to the aminotransferase or glutamate dehydrogenase (E1). Rates significantly greater than those which can be accounted for by the concentration of free ligand, calculated from the measured values of the E1-ligand dissociation constants, require that the E1-ligand complex serve as a substrate for E2 (Srivastava, D. K., and Bernhard, S. A. (1986) Curr. Tops. Cell Regul. 28, 1-68). By this criterion, NADH is transferred directly from glutamate dehydrogenase to malate dehydrogenase and alpha-ketoglutarate is channeled from the aminotransferase to both glutamate dehydrogenase and the alpha-ketoglutarate dehydrogenase complex. Similar evidence indicates that GTP bound to an allosteric site on glutamate dehydrogenase functions as a substrate for succinic thiokinase. The potential physiological advantages to channeling of activators and inhibitors as well as substrates within multienzyme complexes organized around the alpha-ketoglutarate dehydrogenase complex are discussed.  相似文献   

3.
The pathway for the transfer of NADH from one dehydrogenase (E1) to another dehydrogenase (E2) has been investigated by studying the E2-catalyzed reduction of S2 by NADH. The experimental conditions are that the concentration of E1 exceeds that of NADH, which in turn is very much greater than E2; hence, the concentration of free (aqueous) NADH is exceedingly low. The rate of reduction of S2 will hence be very slow if unliganded aqueous NADH is required for the E2-catalyzed reaction. Our results with eight dehydrogenases are entirely consistent with the direct transfer of NADH between E1 and E2 whenever the two enzymes transfer hydrogen via opposite faces (A and B) of the nicotinamide ring. Whenever the two enzymes are both A or both B, NADH transfer occurs only via the aqueous solvent. Some mechanistic inferences and their possible physiological significance are discussed.  相似文献   

4.
A kinetic scheme is presented for Escherichia coli dihydrofolate reductase that predicts steady-state kinetic parameters and full time course kinetics under a variety of substrate concentrations and pHs. This scheme was derived from measuring association and dissociation rate constants and pre-steady-state transients by using stopped-flow fluorescence and absorbance spectroscopy. The binding kinetics suggest that during steady-state turnover product dissociation follows a specific, preferred pathway in which tetrahydrofolate (H4F) dissociation occurs after NADPH replaces NADP+ in the ternary complex. This step, H4F dissociation from the E X NADPH X H4F ternary complex, is proposed to be the rate-limiting step for steady-state turnover at low pH because koff = VM. The rate constant for hydride transfer from NADPH to dihydrofolate (H2F), measured by pre-steady-state transients, has a deuterium isotope effect of 3 and is rapid, khyd = 950 s-1, essentially irreversible, Keq = 1700, and pH dependent, pKa = 6.5, reflecting ionization of a single group in the active site. This scheme accounts for the apparent pKa = 8.4 observed in the steady state as due to a change in the rate-determining step from product release at low pH to hydride transfer above pH 8.4. This kinetic scheme is a necessary background to analyze the effects of single amino acid substitutions on individual rate constants.  相似文献   

5.
1. The steady-state parameters kcat and Km and the rate constants of hydride transfer for the substrates isopropanol/acetone; (S)-2-butanol, (R)-2-butanol/2-butanone; (S)-2-pentanol, (R)-2-pentanol/2-pentanone; 3-pentanol/3-pentanone; (S)-2-octanol and (R)-2-octanol have been determined for the native Zn(II)-containing horse-liver alcohol dehydrogenase (LADH) and the specific active-site-substituted Co(II)LADH. 2. A combined evaluation of steady-state kinetic data and rate constants obtained from stopped-flow measurements, allowed the determination of all rate constants of the following ordered bi-bi mechanism: E in equilibrium E.NAD in equilibrium E.NAD.R1R2 CHOH in equilibrium E.NADH.R1R2CO in equilibrium E.NADH in equilibrium E. 3. On the basis of the different substrate specificities of LADH and yeast alcohol dehydrogenase (YADH), a procedure has been developed to evaluate the enantiomeric product composition of ketone reductions. 2-Butanone and 2-pentanone reductions revealed (S)-2-butanol (86%) and (S)-2-pentanol (95%) as the major products. 4. The observed enantioselectivity implies the existence of two productive ternary complexes; E.NADH.(pro-S) 2-butanone and E.NADH.(pro-R) 2-butanone. All rate constants describing the kinetic pathways of the system (S)-2-butanol, (R)-2-butanol/2-butanone have been determined. These data have been used to estimate the expected enantiomer product composition of 2-butanone reductions using apparent kcat/Km values for the two different ternary-complex configurations of 2-butanone. Additionally, these data have been used for computer simulations of the corresponding reaction cycles. Calculated, simulated and experimental data were found to be in good agreement. Thus, the system (S)-2-butanol, (R)-2-butanol/2-butanone is the first example of a LADH-catalyzed reaction for which the stereochemical course could be described in terms of rate constants of the underlying mechanism. 5. The effects of Co(II) substitution on the different steps of the kinetic pathway have been investigated. The free energy of activation is higher for alcohol oxidation and lower for ketone reduction when catalyzed by Co(II)LADH in comparison to Zn(II)LADH. However, the free energies of binding are affected by metal substitution in such a way that the enantioselectivity of ketone reduction is not significantly changed by the substitution of Co(II) for Zn(II). 6. Evaluation of the data shows that substrate specificity and stereoselectivity result from combination of the free energies of binding and activation, with differences in binding energies as the dominating factors. In this regard, the interactions of substrate molecules with the protein moiety are dominant over the interactions with the catalytic metal ion.  相似文献   

6.
C J Kay  L P Solomonson  M J Barber 《Biochemistry》1991,30(48):11445-11450
Assimilatory nitrate reductase (NR) from Chlorella is homotetrameric, each subunit containing FAD, heme, and Mo-pterin in a 1:1:1 stoichiometry. Measurements of NR activity and steady-state reduction of the heme component under conditions of NADH limitation or competitive inhibition by nitrite suggested intramolecular electron transfer between heme and Mo-pterin was a rate-limiting step and provided evidence that heme is an obligate intermediate in the transfer of electrons between FAD and Mo-pterin. In addition to the physiological substrates NADH and nitrate, various redox mediators undergo reactions with one or more of the prosthetic groups. These reactions are coupled by NR to NADH oxidation or nitrate reduction. To test whether intramolecular redox reactions of NR were rate-determining, rate constants for redox reactions between NR and several chemically diverse mediators were measured by cyclic voltammetry in the presence of NADH or nitrate. Reduction of ferrocenecarboxylic acid, dichlorophenolindophenol, and cytochrome c by NADH-reduced NR was coupled to reoxidation at a glassy carbon electrode (ferrocene and dichlorophenolindophenol) or at a bis(4-pyridyl) disulfide modified gold electrode (cytochrome c), yielding rate constants of 10.5 x 10(6), 1.7 x 10(6), and 2.7 x 10(6) M-1 s-1, respectively, at pH 7. Kinetics were consistent with a second-order reaction, implying that intramolecular heme reduction by NADH and endogenous FAD was not limiting. In contrast, reduction of methyl viologen and diquat at a glassy carbon electrode, coupled to oxidation by NR and nitrate, yielded similar kinetics for the two dyes. In both cases, second-order kinetics were not obeyed, and reoxidation of dye-reduced Mo-pterin of NR by nitrate became limiting at low scan rates.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

7.
The direct transfer of NADH between A-B pairs of dehydrogenases and also the dissociation of NADH from individual E-NADH complexes have been investigated by transient stopped-flow kinetic techniques. Such A-B transfers of NADH occur without the intermediate dissociation of coenzyme into the aqueous solvent environment [Srivastava, D.K., & Bernhard, S.A. (1985) Biochemistry 24, 623-628]. The equilibrium distributions of limiting NADH among aqueous solvent and A and B dehydrogenase sites have also been determined. At sufficiently high but realizable concentrations of dehydrogenases, both the transfer rate and the equilibrium distribution of bound NADH are virtually independent of the excessive enzyme concentrations; at excessive E2 concentration, substantial NADH is bound to the E1 site. These results further substantiate earlier kinetic arguments for the preferential formation of an EA-NADH-EB complex, within which coenzyme is directly transferred between sites. The unimolecular specific rates of coenzyme transfer from site to site are nearly invariant among different A-B dehydrogenase pairs. The equilibrium constants for the distribution of coenzyme within the EA X EB complexes are near unity. At high [E2] and for [E2] greater than [E1] greater than [NADH], E1-NADH X E2 and E1 X NADH-E2 are virtually the only coenzyme-contained species. In contrast to the nearly invariant unimolecular NADH transfer rates within EA X EB complexes, unimolecular specific rates of dissociation of NADH from E-NADH into aqueous solution are highly variable.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

8.
Rubach JK  Ramaswamy S  Plapp BV 《Biochemistry》2001,40(42):12686-12694
The participation of Val-292 in catalysis by alcohol dehydrogenase and the involvement of dynamics were investigated. Val-292 interacts with the nicotinamide ring of the bound coenzyme and may facilitate hydride transfer. The substitution of Val-292 with Ser (V292S) increases the dissociation constants for the coenzymes (NAD(+) by 50-fold, NADH by 75-fold) and the turnover numbers by 3-7-fold. The V292S enzyme crystallized in the presence of NAD(+) and 2,3,4,5,6-pentafluorobenzyl alcohol has an open conformation similar to the structure of the wild-type apo-enzyme, rather than the closed conformation observed for ternary complexes with wild-type enzyme. The V292S substitution perturbs the conformational equilibrium of the enzyme and decreases the kinetic complexity, which permits study of the hydride transfer step with steady-state kinetics. Eyring plots show that the DeltaH for the oxidation (V(1)) of the protio and deuterio benzyl alcohols is 13 kcal/mol and that the kinetic isotope effect of 4.1 is essentially temperature-independent. Eyring plots for the catalytic efficiency for reduction of benzaldehyde (V(2)/K(p)) with NADH or NADD are distinctly convex, being temperature-dependent from 5 to 25 degrees C and temperature-independent from 25 to 50 degrees C; the kinetic isotope effect of 3.2 for V(2)/K(p) is essentially independent of the temperature. The temperature dependencies and isotope effects for V(1) and V(2)/K(p) are not adequately explained by semiclassical transition state theory and are better explained by hydride transfer occurring through vibrationally assisted tunneling.  相似文献   

9.
The binding of NADH and NAD+ to the human liver cytoplasmic, E1, and mitochondrial, E2, isozymes at pH 7.0 and 25 degrees C was studied by the NADH fluorescence enhancement technique, the sedimentation technique, and steady-state kinetics. The binding of radiolabeled [14C]NADH and [14C]NAD+ to the E1 isozyme when measured by the sedimentation technique yielded linear Scatchard plots with a dissociation constant of 17.6 microM for NADH and 21.4 microM for NAD+ and a stoichiometry of ca. two coenzyme molecules bound per enzyme tetramer. The dissociation constant, 19.2 microM, for NADH as competitive inhibitor was found from steady-state kinetics. With the mitochondrial E2 isozyme, the NADH fluorescence enhancement technique showed only one, high-affinity binding site (KD = 0.5 microM). When the sedimentation technique and radiolabeled coenzymes were used, the binding studies showed nonlinear Scatchard plots. A minimum of two binding sites with lower affinity was indicated for NADH (KD = 3-6 microM and KD = 25-30 microM) and also for NAD+ (KD = 5-7 microM and KD = 15-30 microM). A fourth binding site with the lowest affinity (KD = 184 microM for NADH and KD = 102 microM for NAD+) was observed from the steady-state kinetics. The dissociation constant for NAD+, determined by the competition with NADH via fluorescence titration, was found to be 116 microM. The number of binding sites found by the fluorescence titration (n = 1 for NADH) differs from that found by the sedimentation technique (n = 1.8-2.2 for NADH and n = 1.2-1.6 for NAD+).(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

10.
In this work, we have postulated a comprehensive and unified chemical mechanism of action for yeast alcohol dehydrogenase (EC 1.1.1.1, constitutive, cytoplasmic), isolated from Saccharomyces cerevisiae. The chemical mechanism of yeast enzyme is based on the integrity of the proton relay system: His-51....NAD+....Thr-48....R.CH2OH(H2>O)....Zn++, stretching from His-51 on the surface of enzyme to the active site zinc atom in the substrate-binding site of enzyme. Further, it is based on extensive studies of steady-state kinetic properties of enzyme which were published recently. In this study, we have reported the pH-dependence of dissociation constants for several competitive dead-end inhibitors of yeast enzyme from their binary complexes with enzyme, or their ternary complexes with enzyme and NAD+ or NADH; inhibitors include: pyrazole, acetamide, sodium azide, 2-fluoroethanol, and 2,2,2-trifluorethanol. The unified mechanism describes the structures of four dissociation forms of apoenzyme, two forms of the binary complex E.NAD+, three forms of the ternary complex E.NAD+.alcohol, two forms of the ternary complex E.NADH.aldehyde and three binary complexes E.NADH. Appropriate pKa values have been ascribed to protonation forms of most of the above mentioned complexes of yeast enzyme with coenzymes and substrates.  相似文献   

11.
The use of the steady-state treatment in the study of rapid kinetics was illustrated with experiments on horse liver alcohol dehydrogenase using a stopped-flow spectrophoto-fluorimeter. The amplitude of the “burst” formation of NADH fluorescence observed in the transient reaction of horse liver alcohol dehydrogenase, NAD+, and ethanol corresponded mainly to the steady-state concentration of the binary complex, horse liver alcohol dehydrogenase-NADH. The results on the forward and reverse reactions are shown to be consistent with a Theorell-Chance mechanism. The formation of the ternary complexes appeared to decrease the “burst” formation of the binary complex in the benzylalcoholbenzaldehyde system. There was no evidence for the participation of nonequivalent states of the two active sites in the enzyme molecule. It is shown that the equilibrium constants and rate constants involving the mechanisms of LADH reactions can be evaluated using the data of the amplitude of the “burst” reaction in similar manner to that of usual steady-state kinetics.  相似文献   

12.
Enterobacter cloacae NAD(P)H:nitroreductase (NR; EC 1.6.99.7) catalyzes two-electron reduction of a series of quinoidal compounds according to a "ping-pong" scheme, with marked substrate inhibition by quinones. The steady-state catalytic constants (k(cat)) range from 0.1 to 1600s(-1), and bimolecular rate constants (k(cat)/K(m)) range from 10(3) to 10(8)M(-1)s(-1). Quinones, nitroaromatic compounds and competitive to NADH inhibitor dicumarol, quench the flavin mononucleotide (FMN) fluorescence of nitroreductase. The reactivity of NR with single-electron acceptors is consistent with an "outer-sphere" electron transfer model, taking into account high potential of FMN semiquinone/FMNH(-) couple and good solvent accessibility of FMN. However, the single-electron acceptor 1,1(')-dibenzyl-4,4(')-bipyridinium was far less reactive than quinones possessing similar single-electron reduction potentials (E(1)(7)). For all quinoidal compounds except 2-hydroxy-1,4-naphthoquinones, there existed parabolic correlations between the log of rate constants of quinone reduction and their E(1)(7) or hydride-transfer potential (E(7)(Q/QH(-))). Based on pH dependence of rate constants, a single-step hydride transfer seems to be a more feasible quinone reduction mechanism. The reactivities of 2-hydroxy-1,4-naphthoquinones were much higher than expected from their reduction potential. Most probably, their enhanced reactivity was determined by their binding at or close to the binding site of NADH and dicumarol, whereas other quinones used the alternative, currently unidentified binding site.  相似文献   

13.
We report the synthesis of fluorescently labeled ubiquitin (Ub) and its use for following ubiquitin transfer to various proteins. Using Oregon green (Og) succinimidyl ester, we prepared a population of Ub mainly labeled by a single Og molecule; greater than 95% of the Og label is associated with Lys 6 of Ub. We demonstrate that Og-Ub is efficiently accepted by Ub-utilizing enzymes, such as the human ubiquitin-activating enzyme (E1). We used this fluorescent substrate to follow the steady-state kinetics of human E1-catalyzed Ub-transfer to the ubiquitin-carrier enzyme Ubc4. In this reaction, E1 uses three substrates: ATP, Ubc4, and Ub. The steady-state kinetics of Og-Ub utilization by E1 is presented. We have also used analytical ultracentrifugation methods to establish that E1 is monomeric under our assay condition (low salt) as well as under physiological condition (150 mM NaCl).  相似文献   

14.
1. The kinetics of 1,3-bisphosphoglycerate binding to glyceraldehyde-3-phosphate dehydrogenase have been examined by stopped-flow techniques in the absence and presence of phosphoglycerate kinase, using enzyme concentrations in the range 0.5-40 microM. Rate and equilibrium constant estimates for the interaction of the ligand with the two enzymes are reported. 2. The kinetics of ligand transfer from the binary complex of bisphosphoglycerate and phosphoglycerate kinase to the binary complex of NAD+ and glyceraldehyde-3-phosphate dehydrogenase conform excellently to the predictions of a standard free-diffusion mechanism and exhibit no detectable contributions from a mechanism of direct (channelized) transfer of bisphosphoglycerate between the two enzymes. 3. Previously reported evidence that the binary complex of bisphosphoglycerate and phosphoglycerate kinase may act (in the presence of NADH) as a substrate for glyceraldehyde-3-phosphate dehydrogenase according to Michaelis-Menten kinetics is based on a misinterpretation of the experimental observations that can be attributed to neglect of the autocatalytic effect of NAD+ produced during the reaction. Experiments performed under conditions where the autocatalytic effect of NAD+ is eliminated provide clear evidence that the kinetics of utilization of the kinase-bisphosphoglycerate complex for enzymic NADH reduction are consistent with prior dissociation of the complex according to a free-diffusion mechanism of metabolite transfer and incompatible with a mechanism of direct metabolite transfer. 4. A kinetic argument is presented which renders implausible the very idea that direct metabolite transfer between 'soluble' consecutive enzymes in metabolic pathways may offer any catalytic advantages in comparison to metabolite transfer by free diffusion. A mechanism of direct metabolite transfer seems intuitively attractive only because one tends to disregard the diffusional processes required to bring the consecutive enzymes together and to separate them when the transfer has been completed. Direct metabolite transfer would be expected to be catalytically advantageous only in tightly bound multienzyme complexes showing no kinetically significant tendency to dissociate. 5. It is concluded that mechanisms of direct metabolite transfer have not been convincingly demonstrated to apply, nor are they likely to apply, between 'soluble' consecutive enzymes in metabolic pathways, at least not in the glycolytic sequence of reactions.  相似文献   

15.
The steady-state kinetics of oxidation of the mitochondrial NADH: ubiquinone oxidoreductase (complex I, EC 1.6.99.3) by artificial electron acceptors--p-quinones and inorganic complexes has been investigated. A limiting stage in the NADH: ferricyanide reductase reaction is a reductive half-reaction. Ferricyanide interacts with negative-charged protein groups taking part in the NADH binding. The rate constants of the quinone reduction by complex I vary from 1.10(6) to 4.10(3) M-1s-1. The NADH, NAD+ and ADP-ribose inhibition data indicate that oxidizers in the rotenono-insensitive reaction interact with the redox centre near the NAD+/NADH binding site, most probably with FMN.  相似文献   

16.
Trimmer EE  Ballou DP  Matthews RG 《Biochemistry》2001,40(21):6205-6215
The flavoprotein methylenetetrahydrofolate reductase (MTHFR) from Escherichia coli catalyzes the reduction of 5,10-methylenetetrahydrofolate (CH(2)-H(4)folate) to 5-methyltetrahydrofolate (CH(3)-H(4)folate) using NADH as the source of reducing equivalents. The enzyme also catalyzes the transfer of reducing equivalents from NADH or CH(3)-H(4)folate to menadione, an artificial electron acceptor. Here, we have determined the midpoint potential of the enzyme-bound flavin to be -237 mV. We have examined the individual reductive and oxidative half-reactions constituting the enzyme's activities. In an anaerobic stopped-flow spectrophotometer, we have measured the rate constants of flavin reduction and oxidation occurring in each half-reaction and have compared these with the observed catalytic turnover numbers measured under steady-state conditions. We have shown that, in all cases, the half-reactions proceed at rates sufficiently fast to account for overall turnover, establishing that the enzyme is kinetically competent to catalyze these oxidoreductions by a ping-pong Bi-Bi mechanism. Reoxidation of the reduced flavin by CH(2)-H(4)folate is substantially rate limiting in the physiological NADH-CH(2)-H(4)folate oxidoreductase reaction. In the NADH-menadione oxidoreductase reaction, the reduction of the flavin by NADH is rate limiting as is the reduction of flavin by CH(3)-H(4)folate in the CH(3)-H(4)folate-menadione oxidoreductase reaction. We conclude that studies of individual half-reactions catalyzed by E. coli MTHFR may be used to probe mechanistic questions relevant to the overall oxidoreductase reactions.  相似文献   

17.
Pig heart lactate dehydrogenase was studied in the direction of pyruvate and NADH formation by recording rapid changes in extinction, proton concentration, nucleotide fluorescence and protein fluorescence. Experiments measuring extinction changes show that there is a very rapid formation of NADH within the first millisecond and that the amplitude of this phase (phase 1) increases threefold over the pH range 6-8. A second transient rate (phase 2) can also be distinguished (whose rate is pH-dependent), followed by a steady-state rate (phase 3) of NADH production. The sum of the amplitudes of the first two phases corresponds to 1mol of NADH produced/mol of active sites of lactate dehydrogenase. Experiments that measured the liberation of protons by using Phenol Red as an indicator show that no proton release occurs during the initial very rapid formation of NADH (phase 1), but protons are released during subsequent phases of NADH production. Fluorescence experiments help to characterize these phases, and show that the very rapid phase 1 corresponds to the establishment of an equilibrium between E(NAD) (Lactate) right harpoon over left harpoon H(+)E(NADH) (Pyruvate). This equilibrium can be altered by changing lactate concentration or pH, and the H(+)E(NADH) (Pyruvate) species formed has very low nucleotide fluorescence and quenched protein fluorescence. Phase 2 corresponds to the dissociation of pyruvate and a proton from the complex with a rate constant of 1150s(-1). The observed rate constant is slower than this and is proportional to the position of the preceding equilibrium. The E(NADH) formed has high nucleotide fluorescence and quenched protein fluorescence. The reaction, which is rate-limiting during steady-state turnover, must then follow this step and be involved with dissociation of NADH from the enzyme or some conformational change immediately preceding dissociation. Several inhibitory complexes have also been studied including E(NAD+) (Oxamate) and E(NADH) (Oxamate') and the abortive ternary complex E(NADH) (Lactate). The rate of NADH dissociation from the enzyme was measured and found to be the same whether measured by ligand displacement or by relaxation experiments. These results are discussed in relation to the overall mechanism of lactate dehydrogenase turnover and the independence of the four binding sites in the active tetramer.  相似文献   

18.
Nidetzky B  Klimacek M  Mayr P 《Biochemistry》2001,40(34):10371-10381
Microbial xylose reductase, a representative aldo-keto reductase of primary sugar metabolism, catalyzes the NAD(P)H-dependent reduction of D-xylose with a turnover number approximately 100 times that of human aldose reductase for the same reaction. To determine the mechanistic basis for that physiologically relevant difference and pinpoint features that are unique to the microbial enzyme among other aldo/keto reductases, we carried out stopped-flow studies with wild-type xylose reductase from the yeast Candida tenuis. Analysis of transient kinetic data for binding of NAD(+) and NADH, and reduction of D-xylose and oxidation of xylitol at pH 7.0 and 25 degrees C provided estimates of rate constants for the following mechanism: E + NADH right arrow over left arrow E.NADH right arrow over left arrow E.NADH + D-xylose right arrow over left arrow E.NADH.D-xylose right arrow over left arrow E.NAD(+).xylitol right arrow over left arrow E.NAD(+) right arrow over left arrow E.NAD(+) right arrow over left arrow E + NAD(+). The net rate constant of dissociation of NAD(+) is approximately 90% rate limiting for k(cat) of D-xylose reduction. It is controlled by the conformational change which precedes nucleotide release and whose rate constant of 40 s(-)(1) is 200 times that of completely rate-limiting E.NADP(+) --> E.NADP(+) step in aldehyde reduction catalyzed by human aldose reductase [Grimshaw, C. E., et al. (1995) Biochemistry 34, 14356-14365]. Hydride transfer from NADH occurs with a rate constant of approximately 170 s(-1). In reverse reaction, the E.NADH --> E.NADH step takes place with a rate constant of 15 s(-1), and the rate constant of ternary-complex interconversion (3.8 s(-1)) largely determines xylitol turnover (0.9 s(-1)). The bound-state equilibrium constant for C. tenuis xylose reductase is estimated to be approximately 45 (=170/3.8), thus greatly favoring aldehyde reduction. Formation of productive complexes, E.NAD(+) and E.NADH, leads to a 7- and 9-fold decrease of dissociation constants of initial binary complexes, respectively, demonstrating that 12-fold differential binding of NADH (K(i) = 16 microM) vs NAD(+) (K(i) = 195 microM) chiefly reflects difference in stabilities of E.NADH and E.NAD(+). Primary deuterium isotope effects on k(cat) and k(cat)/K(xylose) were, respectively, 1.55 +/- 0.09 and 2.09 +/- 0.31 in H(2)O, and 1.26 +/- 0.06 and 1.58 +/- 0.17 in D(2)O. No deuterium solvent isotope effect on k(cat)/K(xylose) was observed. When deuteration of coenzyme selectively slowed the hydride transfer step, (D)()2(O)(k(cat)/K(xylose)) was inverse (0.89 +/- 0.14). The isotope effect data suggest a chemical mechanism of carbonyl reduction by xylose reductase in which transfer of hydride ion is a partially rate-limiting step and precedes the proton-transfer step.  相似文献   

19.
The turnover numbers and other kinetic constants for human alcohol dehydrogenase (ADH) 4 ("stomach" isoenzyme) are substantially larger (10-100-fold) than those for human class I and horse liver alcohol dehydrogenases. Comparison of the primary amino acid sequences (69% identity) and tertiary structures of these enzymes led to the suggestion that residue 317, which makes a hydrogen bond with the nicotinamide amide nitrogen of the coenzyme, may account for these differences. Ala-317 in the class I enzymes is substituted with Cys in human ADH4, and locally different conformations of the peptide backbones could affect coenzyme binding. This hypothesis was tested by making the A317C substitution in horse liver ADH1E and comparisons to the wild-type ADH1E. The steady-state kinetic constants for the oxidation of benzyl alcohol and the reduction of benzaldehyde catalyzed by the A317C enzyme were very similar (up to about 2-fold differences) to those for the wild-type enzyme. Transient kinetics showed that the rate constants for binding of NAD(+) and NADH were also similar. Transient reaction data were fitted to the full Ordered Bi Bi mechanism and showed that the rate constants for hydride transfer decreased by about 2.8-fold with the A317C substitution. The structure of A317C ADH1E complexed with NAD(+) and 2,3,4,5,6-pentafluorobenzyl alcohol at 1.2 ? resolution is essentially identical to the structure of the wild-type enzyme, except near residue 317 where the additional sulfhydryl group displaces a water molecule that is present in the wild-type enzyme. ADH is adaptable and can tolerate internal substitutions, but the protein dynamics apparently are affected, as reflected in rates of hydride transfer. The A317C substitution is not solely responsible for the larger kinetic constants in human ADH4; thus, the differences in catalytic activity must arise from one or more of the other hundred substitutions in the enzyme.  相似文献   

20.
NADH oxidation by quinone electron acceptors   总被引:1,自引:0,他引:1  
The rate constants of NADH oxidation by quinones are increased with the oxidation potential increase: log kox (M-1 X s-1) = -0.25 + 12.2 E0(7) (V) for o-quinones and log kox (M-1 X s-1) = -3.06 + 13.5 E0(7) (V) for p-quinones (pH 7.0, 25 degrees C). It is assumed that the oxidation proceeds via the hydride-ion transfer. The rate constants of NADH oxidation by single-electron quinone acceptors are also increased with the oxidizer potential increase; log kox (M-1 X s-1) = -0.64 + 9.34 E0(7) (V) and correlate with the constants of NADH oxidation by quinone radicals obtained earlier (Grodkowski, J., Neta, P., Carlson, B.W. and Miller, L. (1983) J. Phys. Chem. 87, 3135-3138). Single-electron transfer is the limiting stage of the process.  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