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 46 毫秒
1.
Oxamate competes with pyruvate for the substrate binding site on the E(NADH) complex of pig skeletal muscle lactate dehydrogenase. When this enzyme was mixed with saturating concentrations of NAD(+) and lactate in a stopped-flow rapid-reaction spectrophotometer there was no transient accumulation of enzyme complexes with the reduced nucleotide. The steady-state rate of formation of free NADH was reached within the dead-time of the instrument (3ms). When oxamate was added to inhibit the steady state and to uncouple the equilibration: [Formula: see text] through the rapid formation of E(NADH) (Oxamate), the rate of formation of E(NADH) could be measured by observation of the first turnover. This pH-dependent transient is controlled by the rate of dissociation of pyruvate and the fraction of the enzyme in the form E(NADH) (Pyruvate).  相似文献   

2.
The tetrameric molecule of pig skeletal muscle lactate dehydrogenase binds a cationic fluorescent probe, auramine O, at four equal non-interacting sites with a dissociation constant of (1.25 +/- 0.2) X 10(-4) M. Fluorescence of the dye/enzyme mixture is strongly pH-dependent, with a maximum at pH 6.3-6.8. Auramine O-binding sites are located outside the active center of the enzyme. The microenvironment of the bound dye changes upon interaction of lactate dehydrogenase with NAD+, NADH, ADP and pyruvate. The binding of specific ligands induces an increase in fluorescence of auramine O-enzyme complex. This effect was used to determine the dissociation constants of the complexes of lactate dehydrogenase with specific ligands. Pyruvate was demonstrated to bind to the apoenzyme-auramine O complex with a dissociation constant of 5.2 X 10(-4) M. With the use of auramine O, it became possible to reveal subunit interactions within the tetrameric molecule of lactate dehydrogenase. They are manifested in the changes of the microenvironment of a dye-binding site located on one of the subunits induced by the binding of ligands in the active center of a neighboring subunit.  相似文献   

3.
1. The reaction of pig heart lactate dehydrogenase (EC 1.1.1.27) with NAD(+) and lactate to form pyruvate and NADH was followed by rapid spectrophotometric methods. The distinct spectrum of enzyme-bound NADH permits the measurement of the rate of dissociation of this compound. 2. The reduction of the first mole equivalent of NAD(+) per mole of enzyme sites can also be observed, and is much more rapid than the steady-state rate of NADH production. 3. At pH8 the dissociation of the enzyme-NADH complex is rate-determining for the steady-state oxidation of lactate. At lower pH some other step after the interconversion of the ternary complex and before the dissociation of NADH is rate-determining. Other evidence for a compulsory-order mechanism is provided.  相似文献   

4.
The interactions of the essential divalent cation, Zn2+, with the binary complex formed between glycerol dehydrogenase (glycerol:NAD+ 2-oxidoreductase, EC 1.1.1.6) and its coenzyme NADH have been examined by fluorescence spectroscopy. Both the metallo and non-metallo form of the enzyme bind the coenzyme NADH. The addition of Zn2+ ions to a solution of the binary complex formed between metal-depleted enzyme and NADH results in a rapid increase in fluorescence emission at 430 nm. This has been used to determine the on rate for Zn2+ to the enzyme/binary complex. A dissociation constant of 3.02 +/- 0.25.10(-9) M for the equilibrium between Zn2+ ions and the enzyme has been determined.  相似文献   

5.
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   

6.
本文比较了大然乳酸脱氢酶和硫酸铵稳定的乳酸脱氢酶在盐酸胍性过程式中失活与内源荧光的变化速度.酶失活表现为三相反应,即极快相,其速度常数用停流装置也无法测定;快相和慢相,1M胍变性时,此二相的一级反应速度常数分别为2.7×10~(-3)秒~(-1)和4.17×10~(-4)秒~(-1).在2M硫酸铵存在条件下,用2M胍更性时,快相和慢相的一极反应速度常数分别为6.16×10~(-3)秒~(-1)和1.88×10~(-3)秒~(-1).内源荧光强度的变化表现为二相反应,即极快相,相当酶失活的极快相,但变化幅度远小于酶失活的变化幅度;快相,相当于酶失活的快相,其速度常数为失活速度常数的1/3倍.上述结果表明,类似肌酸激酶,乳酸脱氢酶的失活速度快于酶分子整体构象的变化,相对于整个酶分子来说,活性中心的构象变化对变性剂更加敏感.  相似文献   

7.
The preceding paper (Südi, 1974) reports partial success in describing the conversion of E(NADH) plus pyruvate into E(NAD+) plus lactate in terms of a simple Haldane-type scheme which involves two intermediates (E(NADH) (Pyr) and E(NAD+) (Lac)), where E represents lactate dehydrogenase. This information is completed here by reporting kinetic results obtained by carrying out the same reaction in the opposite direction. The combined results of these two papers confirm the findings of Holbrook & Gutfreund (1973) that the observed spectral changes do take place at the level of resolution of this simple two-intermediate scheme. The following numerical values for the rate (and equilibrium) constants involved in their formation and decomposition are reported: [Formula: see text] It is shown that although the precision of estimation of some of these numerical values is subject to some experimental uncertainty, their derivation from direct experimental observations only involves the principle of microscopic reversibility. This paper describes stopped-flow kinetic observations made with E(NAD+) and lactate as the two reactants. It is shown that fluorescence and u.v.-absorption measurements yield the same experimental rate constant for the last reaction step in which E(NADH) is generated. On the other hand, the generation of E(NADH) (Pyr) can only be indirectly observed, as a less than stoicheiometric ;burst', and by u.v.-absorption measurements only. It is shown that the stoicheiometry of this partial ;burst reaction', and a pre-equilibrium factor in the directly observed rate of E(NADH)-production, yield equivalent information about the reversible oxidation-reduction step. It is further shown that the pre-equilibrium factor that is involved in the generation of E(NADH) can be determined because k(+4)=222s(-1) is already known (Südi, 1974). Since the fluorescence measurements yield much more precise estimations, and their interpretation is considered by the author to be free of ambiguity, the presented quantitative analysis is based on the fluorescence observations.  相似文献   

8.
Digits JA  Hedstrom L 《Biochemistry》1999,38(8):2295-2306
IMP dehydrogenase (IMPDH) catalyzes the oxidation of IMP to XMP with conversion of NAD+ to NADH. This reaction is the rate-limiting step in de novo guanine nucleotide biosynthesis. IMPDH is a target for antitumor, antiviral, and immunosuppressive chemotherapy. We have determined the complete kinetic mechanism for IMPDH from Tritrichomonas foetus using ligand binding, isotope effect, pre-steady-state kinetic, and rapid quench kinetic experiments. Both substrates bind to the free enzyme, which suggests a random mechanism. IMP binds to the enzyme in two steps. Two steps are also involved when IMP binds to a mutant IMPDH in which the active site Cys is substituted with a Ser. This observation suggests that this second step may be a conformational change of the enzyme. No Vm isotope effect is observed when [2-2H]IMP is the substrate which indicates that hydride transfer is not rate-limiting. This result is confirmed by the observation of a pre-steady-state burst of NADH production when monitored by absorbance. However, when NADH production was monitored by fluorescence, the rate constant for the exponential phase is 5-10-fold lower than when measured by absorbance. This observation suggests that the fluorescence of enzyme-bound NADH is quenched and that this transient represents NADH release from the enzyme. The time-dependent formation and decay of [14C]E-XMP intermediates was monitored using rapid quench kinetics. These experiments indicate that both NADH release and E-XMP hydrolysis are rate-limiting and suggest that NADH release precedes hydrolysis of E-XMP.  相似文献   

9.
1. The mechanisms of the reduction of oxaloacetate and of 3-fluoro-oxaloacetate by NADH catalysed by cytoplasmic pig heart malate dehydrogenase (MDH) were investigated. 2. One mol of dimeric enzyme produces 1.7+/-0.4 mol of enzyme-bound NADH when mixed with saturating NAD+ and L-malate at a rate much higher than the subsequent turnover at pH 7.5. 3. Transient measurements of protein and nucleotide fluorescence show that the steady-state complex in the forward direction is MDH-NADH and in the reverse direction MDH-NADH-oxaloacetate. 4. The rate of dissociation of MDH-NADH was measured and is the same as Vmax. in the forward direction at pH 7.5. Both NADH-binding sites are kinetically equivalent. The rate of dissociation varies with pH, as does the equilibrium binding constant for NADH. 5. 3-Fluoro-oxaloacetate is composed of three forms (F1, F2 and S) of which F1 and F2 are immediately substrates for the enzyme. The third form, S, is not a substrate, but when the F forms are used up form S slowly and non-enzymically equilibrates to yield the active substrate forms. S is 2,2-dihydroxy-3-fluorosuccinate. 6. The steady-state compound during the reduction of form F1 is an enzyme form that does not contain NADH, probably MDH-NAD+-fluoromalate. The steady-state compound for form F2 is an enzyme form containing NADH, probably MDH-NADH-fluoro-oxaloacetate. 7. The rate-limiting reaction in the reduction of form F2 shows a deuterium isotope rate ratio of 4 when NADH is replaced by its deuterium analogue, and the rate-limiting reaction is concluded to be hydride transfer. 8. A novel titration was used to show that dimeric cytoplasmic malate dehydrogenase contains two sites that can rapidly reduce the F1 form of 3-fluoro-oxaloacetate. The enzyme shows 'all-of-the-sites' behaviour. 9. Partial mechanisms are proposed to explain the enzyme-catalysed transformations of the natural and the fluoro substrates. These mechanisms are similar to the mechanism of pig heart lactate dehydrogenase and this, and the structural results of others, can be explained if the two enzymes are a product of divergent evolution.  相似文献   

10.
1. Lactate oxidation catalysed by pig heart lactate dehydrogenase was studied in the presence of inhibitory concentrations of pyruvate. Experimental results show the presence of an intermediate which occurs immediately after the hydride transfer step, but before the dissociation of pyruvate and the H+ produced by the reaction. The rate constant for pyruvate dissociation and the dissociation constant for pyruvate from the ternary complex differ from those obtained in pyruvate reduction experiments. 2.In single-turnover pyruvate reduction by pig heart lactate dehydrogenase at pH8.0 pyruvate can bind to the enzyme before a H+ is taken up, and the subsequent uptake of a H+ is governed by a step that is also rate-limiting for single-turnover and steady-state NADH oxidation. 3. Observation of various intermediates in the single-turnover pyruvate reduction experiments has made it possible to determine separately the dissociation constant and Km value for pyruvate at pH8.0, and also the catalytic turnover rate and Km for pyruvate under first-order conditions at different pH values. 4. Further studies on single-turnover pyruvate reduction carried out in 2H2O, or in water at low temperature, show another step which, under these conditions, is slower than that controlling H+ uptake and rate-limiting for NADH oxidation. A scheme is presented which explains these results.  相似文献   

11.
Bao H  Kasten SA  Yan X  Hiromasa Y  Roche TE 《Biochemistry》2004,43(42):13442-13451
Pyruvate dehydrogenase kinase 2 (PDK2) activity is stimulated by NADH and NADH plus acetyl-CoA via the reduction and reductive acetylation of the lipoyl groups of the dihydrolipoyl acetyltransferase (E2) component. Elevated K(+) and Cl(-) were needed for significant stimulation. Stimulation substantially increased both k(cat) and the K(m) for ATP; the fractional stimulation increased with the level of ATP. With an E2 structure lacking the pyruvate dehydrogenase (E1) binding domain, stimulation of PDK2 was retained, the K(m) for E1 decreased, and the equilibrium dissociation constant for ATP increased but remained much lower than the K(m) for ATP. Stimulation of PDK2 activity greatly reduced the fraction of bound ADP. These results fit an ordered reaction mechanism with ATP binding before E1 and stimulation increasing the rate of dissociation of ADP. Conversion of all of the lipoyl groups in the E2 60mer to the oxidized form (E2(ox)) greatly reduced k(cat) and the K(m) of PDK2 for ATP. Retention over an extended period of time of a low portion of reduced lipoyl groups maintains E2 in a state that supported much higher PDK2 activity than short-term (5 min) reduction of a large portion of lipoyl groups of E2(ox), but reduction of E2(ox) produced a larger fold stimulation. Reduction and to a greater extent reductive acetylation increased PDK2 binding to E2; conversion to E2(ox) did not significantly hinder binding. We suggest that passing even limited reducing equivalents among lipoyl groups maintains E2 lipoyl domains in a conformation that aids kinase function.  相似文献   

12.
J W Burgner  W J Ray 《Biochemistry》1984,23(16):3620-3626
The binary complex of NAD and lactate dehydrogenase reacts reversibly with cyanide to produce a complex (E X NAD-CN) whose noncovalent interactions are similar to those in the E X NADH complex (where E is one-fourth of the tetrameric dehydrogenase). The reaction apparently is a simple bimolecular nucleophilic addition at the 4 position of the bound nicotinamide ring; viz., cyanide does not bind to the enzyme prior to reaction. The value of the dissociation constant for E X NAD-CN is about 1 X 10(-6) M and is independent of pH over the range of 6-8. The equilibrium constant for the reaction of cyanide with E X NAD is about 400-fold larger than that for the nonenzymic process after a statistical correction. This increment in Ke is accounted for by a 220-fold increase in the rate of the forward enzymic reaction (20 M-1 s-1) as compared with an approximately 2-fold decrease for the reverse process (9 X 10(-5) s-1). Thus, the increased value of the rate constant for bond formation in the enzymic reaction is attributed to an equilibrium binding effect that is translated almost entirely into a rate effect on that step (bond formation). Since the nonenzymic reaction is sensitive to solvent composition, this equilibrium binding effect likely is produced by environmental effects at the nicotinamide/dehydronicotinamide part of the coenzyme binding site on the enzyme.  相似文献   

13.
Transient-kinetic studies of pig muscle lactate dehydrogenase   总被引:9,自引:8,他引:1  
1. The very fast pre-steady-state formation of NADH catalysed by pig M(4) lactate dehydrogenase was equivalent to the enzyme-site concentration at pH values greater than 8.0 and to one-half the site concentration at pH6.8. 2. The rate of dissociation of NADH from the enzyme at pH8.0 (450s(-1)) in the absence of other substrates is faster than the steady-state oxidation of lactate (80s(-1)). The latter process is therefore controlled by a step before NADH dissociation but subsequent to the hydride transfer. 3. The oxidation of enzyme-NADH by excess of pyruvate was studied as a first-order process at pH9.0. There was no effect of NADD on this reaction and it was concluded that the ternary complex undergoes a rate-limiting change before the hydride-transfer step. 4. Some conclusions about the reactions catalysed by the M(4) isoenzyme were drawn from a comparison of these results with those obtained with the H(4) isoenzyme and liver alcohol dehydrogenase.  相似文献   

14.
15.
The use of organothallium compounds for protein/macromolecule modification and as probes for n.m.r. and fluorescence is introduced. Lactate dehydrogenase from a number of species was rapidly and specifically inhibited by o-carboxyphenylthallium(III) bistrifluoroacetate and p-methylphenylthallium(III) bistrifluoroacetate. Inhibition of rabbit muscle lactate dehydrogenase by o-carboxyphenylthallium(III) bistrifluoroacetate was time-dependent and not reversible by gel filtration. A small degree of re-activation was possible by incubation with dithiothreitol. The time course of the inactivation kinetics showed two phases, only the first, and faster, of which was efficiently prevented by the presence of cofactor, NADH. Inhibition rates depended on the structure of the thallium reagent, its concentration and the temperature. No significant inhibition was found by thallous acetate or thallic trifluoroacetate. Saturation kinetics were observed for the inhibition by o-carboxyphenylthallium(III) bistrifluoroacetate of the pig heart enzyme. The possibilities of various cross-linking activities of these reagents are addressed. Mechanisms of the inhibition are discussed.  相似文献   

16.
At the normal pH of the cytosol (7.0 to 7.1) and in the presence of physiological (1.0 mM) levels of free Mg2+, the Vmax of the NADPH oxidation is only slightly lower than the Vmax of NADH oxidation in the cytosolic glycerol-3-phosphate dehydrogenase (E.C. 1.1.1.8) reaction. Under these conditions physiological (30 microM) levels of cytosolic malate dehydrogenase (E.C. 1.1.1.37) inhibited oxidation of 20 microM NADH but had no effect on oxidation of 20 microM NADPH by glycerol-3-phosphate dehydrogenase. Consequently malate dehydrogenase increased the ratio of NADPH to NADH oxidation of glycerol-3-phosphate dehydrogenase. On the basis of the measured KD of complexes between malate dehydrogenase and these reduced pyridine nucleotides, and their Km in the glycerol-3-phosphate dehydrogenase reactions, it could be concluded that malate dehydrogenase would have markedly inhibited NADPH oxidation and inhibited NADH oxidation considerably more than observed if its only effect were to decrease the level of free NADH or NADPH. This indicates that due to the opposite chiral specificity of the two enzymes with respect to reduced pyridine nucleotides, complexes between malate dehydrogenase and NADH or NADPH can function as substrates for glycerol-3-phosphate dehydrogenase, but the complex with NADH is less active than free NADH, while the complex with NADPH is as active as free NADPH. Mg2+ enhanced the interactions between malate dehydrogenase and glycerol-3-phosphate dehydrogenase described above. Lactate dehydrogenase (E.C. 1.1.1.27) had effects similar to those of malate dehydrogenase only in the presence of Mg2+. In the absence of Mg2+, there was no evidence of interaction between lactate dehydrogenase and glycerol-3-phosphate dehydrogenase.  相似文献   

17.
The binding of pig skeletal muscle lactate dehydrogenase by F-actin has been studied using the sedimentation method in 10 mM Tris-acetate buffer, pH 6.0 at 20 degrees C. Adsorption capacity of F-actin is equal to (1 +/- 0.1) . 10(-5) moles of lactate dehydrogenase per 1 g of actin. NADH decreases the affinity of F-actin with respect to lactate dehydrogenase. The binding of lactate dehydrogenase by F-actin in diminishing the rate of enzymatic reduction of alpha-ketoglutarate. The microscopic dissociation constant for the complex of the enzyme with F-actin which is estimated from the dependence of the enzymatic reaction rate of F-actin concentration at saturating NADH concentrations is equal (3.0 +2- 0.5) . 10(-7) M. It has been shown that the bound enzyme is characterized by the greater value of Km and the lower value of Vmax in comparison to the free enzyme.  相似文献   

18.
Sheep liver cytoplasmic aldehyde dehydrogenase was purified to homogeneity to give a sample with a specific activity of 380 nmol NADH min(-1) mg(-1). An amino acid analysis of the enzyme gave results similar to those reported for aldehyde dehydrogenases from other sources. The isoelectric point was at pH 5.25 and the enzyme contained no significant amounts of metal ions. On the binding of NADH to the enzyme there is a shift in absorption maximum of NADH to 344 nm, and a 5.6-fold enhancement of nucleotide fluorescence. The protein fluorescence (lambdaexcit = 290 nm, lambdaemisson = 340 nm) is quenched on the binding of NAD+ and NADH. The enhancement of nucleotide fluorescence on the binding of NADH has been utilised to determine the dissociation constant for the enzyme . NADH complex (Kd = 1.2 +/- 0.2 muM). A Hill plot of the data gave a straight line with a slope of 1.0 +/- 0.3 indicating the absence of co-operative effects. Ellman's reagent reacted only slowly with the enzyme but in the presence of sodium dodecylsulphate complete reaction occurred within a few minutes to an extent corresponding to 36 thiol groups/enzyme. Molecular weights were determined for both cytoplasmic and mitochondrial aldehyde dehydrogenases and were 212 000 +/- 8 000 and 205 000 respectively. Each enzyme consisted of four subunits with molecular weight of 53 000 +/- 2 000. Properties of the cytoplasmic and mitochondrial aldehyde dehydrogenases from sheep liver were compared with other mammalian liver aldehyde dehydrogenases.  相似文献   

19.
The metabolism of [2-3H]lactate was studied in isolated hepatocytes from fed and starved rats metabolizing ethanol and lactate in the absence and presence of fructose. The yields of 3H in ethanol, water, glucose and glycerol were determined. The rate of ethanol oxidation (3 mumol/min per g wet wt.) was the same for fed and starved rats with and without fructose. From the detritiation of labelled lactate and the labelling pattern of ethanol and glucose, we calculated the rate of reoxidation of NADH catalysed by lactate dehydrogenase, alcohol dehydrogenase and triosephosphate dehydrogenase. The calculated flux of reducing equivalents from NADH to pyruvate was of the same order of magnitude as previously found with [3H]ethanol or [3H]xylitol as the labelled substrate [Vind & Grunnet (1982) Biochim. Biophys. Acta 720, 295-302]. The results suggest that the cytoplasm can be regarded as a single compartment with respect to NAD(H). The rate of reduction of acetaldehyde and pyruvate was correlated with the concentration of these metabolites and NADH, and was highest in fed rats and during fructose metabolism. The rate of reoxidation of NADH catalysed by lactate dehydrogenase was only a few per cent of the maximal activity of the enzymes, but the rate of reoxidation of NADH catalysed by alcohol dehydrogenase was equal to or higher than the maximal activity as measured in vitro, suggesting that the dissociation of enzyme-bound NAD+ as well as NADH may be rate-limiting steps in the alcohol dehydrogenase reaction.  相似文献   

20.
Stopped flow spectrophotometry was used to investigate the kinetics of the transition of the phosphoglycerate dehydrogenase (3-phosphoglycerate: NAD oxidoreductase, EC 1.1.1.95) reaction from the active to the inhibited rate upon the addition of the physiological inhibitor serine. The transition was characterized by a single first order rate constant (kobs,i) which was independent of enzyme concentration. At pH 8.5, kobs,i increased in a hyperbolic manner with serine concentration from 2 to 8 s-1. The increase in kobs,i occurred at serine concentrations where the steady state inhibition was virtually complete. These results indicate that serine inhibition is an allosteric process involving a conformational change in the enzyme. A model is presented in which serine at low concentrations binds exclusively to the inhibited state of the enzyme and shifts the equilibrium toward that state; at high serine concentrations, serine binds to the active state, facilitating its conversion to the inhibited state. An alternative model, which we favor, proposes two classes of inhibitor binding sites. The kinetics of the fluorescence quenching of enzyme-bound NADH by serine (Sugimoto, E., and Pizer, L.I. (1968) J. Biol. Chem. 243, 2090-2098), measured by stopped flow fluorimetry, was also characterized by a single first order rate constant (kobs,f.q.) which was independent of enzyme concentration. At pH 8.5, kobs,f.q. ranged from 0.4 s-1 at low serine concentrations to 1.1 s-1 at high serine concentrations. These results indicate that the fluorescence quenching induced by serine is a manifestation of a structural change in the enzyme. Enzyme and excess NADH were mixed with substrate and serine in the stopped flow instrument, and enzyme-bound NADH fluorescence was monitored by exciting through the protein at 285 nm. A rapid fluorescence quenching process, which occurred within the mixing time, was followed by a slower fluorescence enhancement process which terminated in a steady state level corresponding to the quenched fluorescence of the enzyme NADH serine complex. The rapid quenching was the result of substrate binding (Dubrow, R., and Pizer, L.I. (1977) J. Biol. Chem. 252, 1539-1551). The fluorescence enhancement was characterized by a single first order rate constant whose value for a given serine concentration corresponded with Kobs,j. This data shows that the quenched state of the enzyme-NADH-complex is the state which is directly responsible for the inhibition of enzyme activity. During catalysis the quenched state is achieved from a different initial conformation, and consequently at a different rate, than in the absence of substrate. kobs,j and kobs,f.q. were also measured using glycine, another inhibitor. The ultraviolet difference spectrum between enzyme and enzyme plus serine was determined and proposed to be the result of the same structural change which is responsible for the fluorescence quenching by serine.  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