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 15 毫秒
1.
Saccharopine dehydrogenase [N6-(glutaryl-2)-L-lysine:NAD oxidoreductase (L-lysine forming)] catalyzes the final step in the alpha-aminoadipate pathway for lysine biosynthesis. It catalyzes the reversible pyridine nucleotide-dependent oxidative deamination of saccharopine to generate alpha-Kg and lysine using NAD+ as an oxidizing agent. The proton shuttle chemical mechanism is proposed on the basis of the pH dependence of kinetic parameters, dissociation constants for competitive inhibitors, and isotope effects. In the direction of lysine formation, once NAD+ and saccharopine bind, a group with a pKa of 6.2 accepts a proton from the secondary amine of saccharopine as it is oxidized. This protonated general base then does not participate in the reaction again until lysine is formed at the completion of the reaction. A general base with a pKa of 7.2 accepts a proton from H2O as it attacks the Schiff base carbon of saccharopine to form the carbinolamine intermediate. The same residue then serves as a general acid and donates a proton to the carbinolamine nitrogen to give the protonated carbinolamine. Collapse of the carbinolamine is then facilitated by the same group accepting a proton from the carbinolamine hydroxyl to generate alpha-Kg and lysine. The amine nitrogen is then protonated by the group that originally accepted a proton from the secondary amine of saccharopine, and products are released. In the reverse reaction direction, finite primary deuterium kinetic isotope effects were observed for all parameters with the exception of V2/K(NADH), consistent with a steady-state random mechanism and indicative of a contribution from hydride transfer to rate limitation. The pH dependence, as determined from the primary isotope effect on DV2 and D(V2/K(Lys)), suggests that a step other than hydride transfer becomes rate-limiting as the pH is increased. This step is likely protonation/deprotonation of the carbinolamine nitrogen formed as an intermediate in imine hydrolysis. The observed solvent isotope effect indicates that proton transfer also contributes to rate limitation. A concerted proton and hydride transfer is suggested by multiple substrate/solvent isotope effects, as well as a proton transfer in another step, likely hydrolysis of the carbinolamine. In agreement, dome-shaped proton inventories are observed for V2 and V2/K(Lys), suggesting that proton transfer exists in at least two sequential transition states.  相似文献   

2.
Saccharopine dehydrogenase (SDH) catalyzes the final reaction in the α-aminoadipate pathway, the conversion of l-saccharopine to l-lysine (Lys) and α-ketoglutarate (α-kg) using NAD? as an oxidant. The enzyme utilizes a general acid-base mechanism to conduct its reaction with a base proposed to accept a proton from the secondary amine of saccharopine in the oxidation step and a group proposed to activate water to hydrolyze the resulting imine. Crystal structures of an open apo form and a closed form of the enzyme with saccharopine and NADH bound have been determined at 2.0 and 2.2 ? resolution, respectively. In the ternary complex, a significant movement of domain I relative to domain II that closes the active site cleft between the two domains and brings H96 and K77 into the proximity of the substrate binding site is observed. The hydride transfer distance is 3.6 ?, and the side chains of H96 and K77 are properly positioned to act as acid-base catalysts. Preparation of the K77M and H96Q single-mutant and K77M/H96Q double-mutant enzymes provides data consistent with their role as the general acid-base catalysts in the SDH reaction. The side chain of K77 initially accepts a proton from the ε-amine of the substrate Lys and eventually donates it to the imino nitrogen as it is reduced to a secondary amine in the hydride transfer step, and H96 protonates the carbonyl oxygen as the carbinolamine is formed. The K77M, H976Q, and K77M/H96Q mutant enzymes give 145-, 28-, and 700-fold decreases in V/E(t) and >103-fold increases in V?/K(Lys)E(t) and V?/K(α-kg)E(t) (the double mutation gives >10?-fold decreases in the second-order rate constants). In addition, the K77M mutant enzyme exhibits a primary deuterium kinetic isotope effect of 2.0 and an inverse solvent deuterium isotope effect of 0.77 on V?/K(Lys). A value of 2.0 was also observed for (D)(V?/K(Lys))(D?O) when the primary deuterium kinetic isotope effect was repeated in D?O, consistent with a rate-limiting hydride transfer step. A viscosity effect of 0.8 was observed on V?/K(Lys), indicating the solvent deuterium isotope effect resulted from stabilization of an enzyme form prior to hydride transfer. A small normal solvent isotope effect is observed on V, which decreases slightly when repeated with NADD, consistent with a contribution from product release to rate limitation. In addition, V?/K(Lys)E(t) is pH-independent, which is consistent with the loss of an acid-base catalyst and perturbation of the pK(a) of the second catalytic group to a higher pH, likely a result of a change in the overall charge of the active site. The primary deuterium kinetic isotope effect for H96Q, measured in H?O or D?O, is within error equal to 1. A solvent deuterium isotope effect of 2.4 is observed with NADH or NADD as the dinucleotide substrate. Data suggest rate-limiting imine formation, consistent with the proposed role of H96 in protonating the leaving hydroxyl as the imine is formed. The pH-rate profile for V?/K(Lys)E(t) exhibits the pK(a) for K77, perturbed to a value of ~9, which must be unprotonated to accept a proton from the ε-amine of the substrate Lys so that it can act as a nucleophile. Overall, data are consistent with a role for K77 acting as the base that accepts a proton from the ε-amine of the substrate lysine prior to nucleophilic attack on the α-oxo group of α-ketoglutarate, and finally donating a proton to the imine nitrogen as it is reduced to give saccharopine. In addition, data indicate a role for H96 acting as a general acid-base catalyst in the formation of the imine between the ε-amine of lysine and the α-oxo group of α-ketoglutarate.  相似文献   

3.
Dihydrofolate reductase from Mycobacterium tuberculosis (MtDHFR) catalyzes the NAD(P)-dependent reduction of dihydrofolate, yielding NAD(P)(+) and tetrahydrofolate, the primary one-carbon unit carrier in biology. Tetrahydrofolate needs to be recycled so that reactions involved in dTMP synthesis and purine metabolism are maintained. In this work, we report the kinetic characterization of the MtDHFR. This enzyme has a sequential steady-state random kinetic mechanism, probably with a preferred pathway with NADPH binding first. A pK(a) value for an enzymic acid of approximately 7.0 was identified from the pH dependence of V, and the analysis of the primary kinetic isotope effects revealed that the hydride transfer step is at least partly rate-limiting throughout the pH range analyzed. Additionally, solvent and multiple kinetic isotope effects were determined and analyzed, and equilibrium isotope effects were measured on the equilibrium constant. (D(2)O)V and (D(2)O)V/K([4R-4-(2)H]NADH) were slightly inverse at pH 6.0, and inverse values for (D(2)O)V([4R-4-(2)H]NADH) and (D(2)O)V/K([4R-4-(2)H]NADH) suggested that a pre-equilibrium protonation is occurring before the hydride transfer step, indicating a stepwise mechanism for proton and hydride transfer. The same value was obtained for (D)k(H) at pH 5.5 and 7.5, reaffirming the rate-limiting nature of the hydride transfer step. A chemical mechanism is proposed on the basis of the results obtained here.  相似文献   

4.
Silva RG  Schramm VL 《Biochemistry》2011,50(42):9158-9166
The reversible phosphorolysis of uridine to generate uracil and ribose 1-phosphate is catalyzed by uridine phosphorylase and is involved in the pyrimidine salvage pathway. We define the reaction mechanism of uridine phosphorylase from Trypanosoma cruzi by steady-state and pre-steady-state kinetics, pH-rate profiles, kinetic isotope effects from uridine, and solvent deuterium isotope effects. Initial rate and product inhibition patterns suggest a steady-state random kinetic mechanism. Pre-steady-state kinetics indicated no rate-limiting step after formation of the enzyme-products ternary complex, as no burst in product formation is observed. The limiting single-turnover rate constant equals the steady-state turnover number; thus, chemistry is partially or fully rate limiting. Kinetic isotope effects with [1'-(3)H]-, [1'-(14)C]-, and [5'-(14)C,1,3-(15)N(2)]uridine gave experimental values of (α-T)(V/K)(uridine) = 1.063, (14)(V/K)(uridine) = 1.069, and (15,β-15)(V/K)(uridine) = 1.018, in agreement with an A(N)D(N) (S(N)2) mechanism where chemistry contributes significantly to the overall rate-limiting step of the reaction. Density functional theory modeling of the reaction in gas phase supports an A(N)D(N) mechanism. Solvent deuterium kinetic isotope effects were unity, indicating that no kinetically significant proton transfer step is involved at the transition state. In this N-ribosyl transferase, proton transfer to neutralize the leaving group is not part of transition state formation, consistent with an enzyme-stabilized anionic uracil as the leaving group. Kinetic analysis as a function of pH indicates one protonated group essential for catalysis and for substrate binding.  相似文献   

5.
Beta-ketoacyl-acyl carrier protein (ACP) reductase from Mycobacterium tuberculosis (MabA) is responsible for the second step of the type-II fatty acid elongation system of bacteria, plants, and apicomplexan organisms, catalyzing the NADPH-dependent reduction of beta-ketoacyl-ACP to generate beta-hydroxyacyl-ACP and NADP(+). In the present work, the mabA-encoded MabA has been cloned, expressed, and purified to homogeneity. Initial velocity studies, product inhibition, and primary deuterium kinetic isotope effects suggested a steady-state random bi-bi kinetic mechanism for the MabA-catalyzed reaction. The magnitudes of the primary deuterium kinetic isotope effect indicated that the C(4)-proS hydrogen is transferred from the pyridine nucleotide and that this transfer contributes modestly to the rate-limiting step of the reaction. The pH-rate profiles demonstrated groups with pK values of 6.9 and 8.0, important for binding of NADPH, and with pK values of 8.8 and 9.6, important for binding of AcAcCoA and for catalysis, respectively. Temperature studies were employed to determine the activation energy of the reaction. Solvent kinetic isotope effects and proton inventory analysis established that a single proton is transferred in a partially rate-limiting step and that the mechanism of carbonyl reduction is probably concerted. The observation of an inverse (D)2(O)V/K and an increase in (D)2(O)V when [4S-(2)H]NADPH was the varied substrate obscured the distinction between stepwise and concerted mechanisms; however, the latter was further supported by the pH dependence of the primary deuterium kinetic isotope effect. Kinetic and chemical mechanisms for the MabA-catalyzed reaction are proposed on the basis of the experimental data.  相似文献   

6.
Homoisocitrate dehydrogenase (HIcDH, 3-carboxy-2-hydroxyadipate dehydrogenase) catalyzes the fourth reaction of the alpha-aminoadipate pathway for lysine biosynthesis, the conversion of homoisocitrate to alpha-ketoadipate using NAD as an oxidizing agent. A chemical mechanism for HIcDH is proposed on the basis of the pH dependence of kinetic parameters, dissociation constants for competitive inhibitors, and isotope effects. According to the pH-rate profiles, two enzyme groups act as acid-base catalysts in the reaction. A group with a p K a of approximately 6.5-7 acts as a general base accepting a proton as the beta-hydroxy acid is oxidized to the beta-keto acid, and this residue participates in all three of the chemical steps, acting to shuttle a proton between the C2 hydroxyl and itself. The second group acts as a general acid with a p K a of 9.5 and likely catalyzes the tautomerization step by donating a proton to the enol to give the final product. The general acid is observed in only the V pH-rate profile with homoisocitrate as a substrate, but not with isocitrate as a substrate, because the oxidative decarboxylation portion of the isocitrate reaction is limiting overall. With isocitrate as the substrate, the observed primary deuterium and (13)C isotope effects indicate that hydride transfer and decarboxylation steps contribute to rate limitation, and that the decarboxylation step is the more rate-limiting of the two. The multiple-substrate deuterium/ (13)C isotope effects suggest a stepwise mechanism with hydride transfer preceding decarboxylation. With homoisocitrate as the substrate, no primary deuterium isotope effect was observed, and a small (13)C kinetic isotope effect (1.0057) indicates that the decarboxylation step contributes only slightly to rate limitation. Thus, the chemical steps do not contribute significantly to rate limitation with the native substrate. On the basis of data from solvent deuterium kinetic isotope effects, viscosity effects, and multiple-solvent deuterium/ (13)C kinetic isotope effects, the proton transfer step(s) is slow and likely reflects a conformational change prior to catalysis.  相似文献   

7.
B N Leichus  J S Blanchard 《Biochemistry》1992,31(12):3065-3072
Lipoamide dehydrogenase is a flavoprotein which catalyzes the reversible oxidation of dihydrolipoamide, Lip(SH)2, by NAD+. The ping-pong kinetic mechanism involves stable oxidized and two-electron-reduced forms. We have investigated the rate-limiting nature of proton transfer steps in both the forward and reverse reactions catalyzed by the pig heart enzyme by using a combination of alternate substrates and solvent kinetic isotope effect studies. With NAD+ as the variable substrate, and at a fixed, saturating concentration of either Lip(SH)2 or DTT, inverse solvent kinetic isotope effects of 0.68 +/- 0.05 and 0.71 +/- 0.05, respectively, were observed on V/K. Solvent kinetic isotope effects on V of 0.91 +/- 0.07 and 0.69 +/- 0.02 were determined when Lip(SH)2 or DTT, respectively, was used as reductant. When Lip(SH)2 or DTT was used as the variable substrate, at a fixed concentration of NAD+, solvent kinetic isotope effects of 0.74 +/- 0.06 and 0.51 +/- 0.04, respectively, were observed on V/K for these substrates. Plots of the kinetic parameters versus mole fraction D2O (proton inventories) were linear in all cases. Solvent kinetic isotope effect measurements performed in the reverse direction using NADH as the variable substrate showed equivalent, normal solvent kinetic isotope effects on V/KNADH when oxidized lipoamide, lipoic acid, or DTT were present at fixed, saturating concentrations. Solvent kinetic isotope effects on V were equal to 1.5-2.1. When solvent kinetic isotope effect measurements were performed using the disulfide substrates lipoamide, lipoic acid, or DTT as the variable substrates, normal kinetic isotope effects on V/K of 1.3-1.7 were observed.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

8.
The pH dependence of the kinetic parameters V, V/KNADH, and V/KH2O2 has been determined for the flavoenzyme NADH peroxidase. Both V/KNADH and V/KH2O2 decrease as groups exhibiting pK's of 9.2 and 9.9, respectively, are deprotonated. The V profile decreases by a factor of 5 as a group exhibiting a pK of 7.2 is deprotonated. Primary deuterium kinetic isotope effects on NADH oxidation are observed on V only, and the magnitude of DV is independent of H2O2 concentration at pH 7.5. DV/KNADH is pH independent and equal to 1.0 between pH 6 and pH 9.5, but DV is pH dependent, decreasing from a value of 7.2 at pH 5.5 to 1.9 at pH 9.5. The shape of the DV versus pH profile parallels that observed in the V profile and yields a similar pK of 6.6 for the group whose deprotonation decreases DV. Solvent kinetic isotope effects obtained with NADH or reduced nicotinamide hypoxanthine dinucleotide as the variable substrate are observed on V only, while equivalent solvent kinetic isotope effects on V and V/K are observed when H2O2 is used as the variable substrate. In all cases linear proton inventories are observed. Primary deuterium kinetic isotope effects on V for NADH oxidation decrease as the solvent isotopic composition is changed from H2O to D2O. These data are consistent with a change in the rate-limiting step from a step in the reductive half-reaction at low pH to a step in the oxidative half-reaction at high pH. Analysis of the multiple kinetic isotope effect data suggests that at high D2O concentrations the rate of a single proton transfer step in the oxidative half-reaction is slowed. These data are used to propose a chemical mechanism involving the pH-dependent protonation of a flavin hydroxide anion, following flavin peroxide bond cleavage.  相似文献   

9.
Fillgrove KL  Anderson VE 《Biochemistry》2001,40(41):12412-12421
The chemical mechanism of the 2,4-dienoyl-CoA reductase (EC 1.3.1.34) from rat liver mitochondria has been investigated. This enzyme catalyzes the NADPH-dependent reduction of 2,4-dienoyl-coenzyme A (CoA) thiolesters to the resulting trans-3-enoyl-CoA. Steady-state kinetic parameters for trans-2,trans-4-hexadienoyl-CoA and 5-phenyl-trans-2,trans-4-pentadienoyl-CoA were determined and demonstrated that the dienoyl-CoA and NADPH bind to the 2,4-dienoyl-CoA reductase via a sequential kinetic mechanism. Kinetic isotope effect studies and the transient kinetics of substrate binding support a random order of nucleotide and dienoyl-CoA addition. The large normal solvent isotope effects on V/K ((D)(2)(O)V/K) and V ((D)(2)(O)V) for trans-2,trans-4-hexadienoyl-CoA reduction indicate that a proton transfer step is rate limiting for this substrate. The stability gained by conjugating the phenyl ring to the diene in PPD-CoA results in the reversal of the rate-determining step, as evidenced by the normal isotope effects on V/K(CoA) ((D)V/K(CoA)) and V/K(NADPH) ((D)V/K(NADPH)). The reversal of the rate-determining step was supported by transient kinetics where a burst was observed for the reduction of trans-2,trans-4-hexadienoyl-CoA but not for 5-phenyl-trans-2,trans-4-pentadienoyl-CoA reduction. The chemical mechanism is stepwise where hydride transfer from NADPH occurs followed by protonation of the observable dienolate intermediate, which has an absorbance maximum at 286 nm. The exchange of the C alpha protons of trans-3-decenoyl-CoA, catalyzed by the 2,4-dienoyl-CoA reductase, in the presence of NADP(+) suggests that formation of the dienolate is catalyzed by the enzyme active site.  相似文献   

10.
Glutathione reductase catalyzes the NADPH-dependent reduction of oxidized glutathione (GSSG). The kinetic mechanism is ping-pong, and we have investigated the rate-limiting nature of proton-transfer steps in the reactions catalyzed by the spinach, yeast, and human erythrocyte glutathione reductases using a combination of alternate substrate and solvent kinetic isotope effects. With NADPH or GSSG as the variable substrate, at a fixed, saturating concentration of the other substrate, solvent kinetic isotope effects were observed on V but not V/K. Plots of Vm vs mole fraction of D2O (proton inventories) were linear in both cases for the yeast, spinach, and human erythrocyte enzymes. When solvent kinetic isotope effect studies were performed with DTNB instead of GSSG as an alternate substrate, a solvent kinetic isotope effect of 1.0 was observed. Solvent kinetic isotope effect measurements were also performed on the asymmetric disulfides GSSNB and GSSNP by using human erythrocyte glutathione reductase. The Km values for GSSNB and GSSNP were 70 microM and 13 microM, respectively, and V values were 62 and 57% of the one calculated for GSSG, respectively. Both of these substrates yield solvent kinetic isotope effects greater than 1.0 on both V and V/K and linear proton inventories, indicating that a single proton-transfer step is still rate limiting. These data are discussed in relationship to the chemical mechanism of GSSG reduction and the identity of the proton-transfer step whose rate is sensitive to solvent isotopic composition. Finally, the solvent equilibrium isotope effect measured with yeast glutathione reductase is 4.98, which allows us to calculate a fractionation factor for the thiol moiety of GSH of 0.456.  相似文献   

11.
Site-directed mutagenesis was used to change Lys 240 of yeast pyruvate kinase (Lys 269 in muscle PK) to Met. K240M has an absolute requirement for FBP for catalysis. K240M is 100- and 1000-fold less active than wild-type YPK in the presence of Mn(2+) and Mg(2+), respectively. Steady-state fluorescence titration data suggest that the substrate PEP binds to K240M with the same affinity as it does to wild-type YPK. The rate of phosphoryl transfer in K240M has been decreased >1000-fold compared to wild-type YPK. The detritiation of 3-[(3)H]pyruvate catalyzed by YPK occurs at a rate significantly greater than the spontaneous rate. Detritiation of pyruvate by wild-type YPK occurs as a divalent metal- and FBP-dependent process requiring ATP. There is no detectable detritiation of pyruvate catalyzed by K240M. The solvent deuterium isotope effect on k(cat) is 2.7 +/- 0.2 and 1.6 +/- 0.1 for the wild type and for K240M YPK, respectively. This suggests that the isotope sensitive step in the PK reaction does not involve Lys 240 and that the enolpyruvate intermediate is still protonated by K240M. Isotope trapping was used to characterize enolpyruvate protonation by K240M. While there was enrichment of the methyl protons of pyruvate from labeled solvent formed by catalysis with muscle PK and wild-type YPK, only background levels of tritium were trapped with K240M. In K240M, the proton donor exchanges protons with the solvent at a higher rate relative to turnover than does the proton donor in wild-type YPK. The pH-rate profile of K240M exhibits the loss of a pK(a) value of 8. 8 observed with wild-type YPK. The above data and recent crystal structure data suggest that Lys 240 interacts with the phosphoryl group of phosphoenolpyruvate and helps to stabilize the pentavalent phosphate transition state during phosphoryl transfer. Phosphoryl transfer is highly coupled to proton transfer, or Lys 240 also affects enolate protonation.  相似文献   

12.
Zheng R  Blanchard JS 《Biochemistry》2003,42(38):11289-11296
Ketopantoate reductase (EC 1.1.1.169), an enzyme in the pantothenate biosynthetic pathway, catalyzes the NADPH-dependent reduction of alpha-ketopantoate to form D-(-)-pantoate. The enzyme exhibits high specificity for ketopantoate, with V and V/K for ketopantoate being 5- and 365-fold higher than those values for alpha-ketoisovalerate and 20- and 648-fold higher than those values for alpha-keto-beta-methyl-n-valerate, respectively. For pyridine nucleotides, V/K for beta-NADPH is 3-500-fold higher than that for other nucleotide substrates. The magnitude of the primary deuterium kinetic isotope effects on V and V/K varied substantially when different ketoacid and pyridine nucleotide substrates were used. The small primary deuterium kinetic isotope effects observed using NADPH and NHDPH suggest that the chemical step is not rate-limiting, while larger primary deuterium isotope effects were observed for poor ketoacid and pyridine nucleotide substrates, indicating that the chemical reaction has become partially or completely rate-limiting. The pH dependence of (D)V using ketopantoate was observed to vary from a value of 1.1 at low pH to a value of 2.5 at high pH, while the magnitude of (D)V/K(NADPH) and (D)V/K(KP) were pH-independent. The value of (D)V is large and pH-independent when alpha-keto-beta-methyl-n-valerate was used as the ketoacid substrate. Solvent kinetic isotope effects of 2.2 and 1.2 on V and V/K, respectively, were observed with alpha-keto-beta-methyl-n-valerate. Rapid reaction analysis of NADPH oxidation using ketopantoate showed no "burst" phase, suggesting that product-release steps are not rate-limiting and the cause of the small observed kinetic isotope effects with this substrate pair. Large primary deuterium isotope effects on V and V/K using 3-APADPH in steady-state experiments, equivalent to the isotope effect observed in single turnover studies, suggests that chemistry is rate-limiting for this poorer reductant. These results are discussed in terms of a kinetic and chemical mechanism for the enzyme.  相似文献   

13.
Argyrou A  Washabaugh MW  Pickart CM 《Biochemistry》2000,39(34):10373-10384
Dihydroorotate dehydrogenase from Clostridium oroticum was purified to apparent homogeneity and found to be a heterotetramer consisting of two alpha (32 kDa) and two beta (28 kDa) polypeptides. This subunit composition, coupled with known cofactor requirements and the ability to transfer electrons from L-dihydroorotate to NAD(+), defines the C. oroticum enzyme as a family 1B dihydroorotate dehydrogenase. The results of steady-state kinetic analyses and isotope exchange studies suggest that this enzyme utilizes a ping-pong steady-state kinetic mechanism. The pH-k(cat) profile is bell-shaped with a pK(a) of 6.4 +/- 0.1 for the ascending limb and 8. 9 +/- 0.1 for the descending limb; the pH-k(cat)/K(m) profile is similar but somewhat more complex. The pK(a) values of 6.4 and 8.9 are likely to represent the ionizations of cysteine and lysine residues in the active site which act as a general base and an electrostatic catalyst, respectively. At saturating levels of NAD(+), the isotope effects on (D)V and (D)(V/K(DHO)), obtained upon deuteration at both the C(5)-proR and C(5)-proS positions of L-dihydroorotate, increase from a value of unity at pH >9.0 to sizable values at low pH due to a high commitment to catalysis at high pH. At pH = 6.5, the magnitude of the double isotope effects (D)V and (D)(V/K(DHO)), obtained upon additional deuteration at C(6), is consistent with a mechanism in which C(5)-proS proton transfer and C(6)-hydride transfer occur in a single, partially rate-limiting step.  相似文献   

14.
L J Hyland  T A Tomaszek  T D Meek 《Biochemistry》1991,30(34):8454-8463
The pH dependence of the peptidolytic reaction of recombinant human immunodeficiency virus type 1 protease has been examined over a pH range of 3-7 for four oligopeptide substrates and two competitive inhibitors. The pK values obtained from the pKis vs pH profiles for the unprotonated and protonated active-site aspartyl groups, Asp-25 and Asp-25', in the monoprotonated enzyme form were 3.1 and 5.2, respectively. Profiles of log V/K vs pH for all four substrates were "bell-shaped" in which the pK values for the unprotonated and protonated aspartyl residues were 3.4-3.7 and 5.5-6.5, respectively. Profiles of log V vs pH for these substrates were "wave-shaped" in which V was shifted to a constant lower value upon protonation of a residue of pK = 4.2-5.2. These results indicate that substrates bind only to a form of HIV-1 protease in which one of the two catalytic aspartyl residues is protonated. Solvent kinetic isotope effects were measured over a pH (D) range of 3-7 for two oligopeptide substrates, Ac-Arg-Ala-Ser-Gln-Asn-Tyr-Pro-Val-Val-NH2 and Ac-Ser-Gln-Asn-Tyr-Pro-Val-Val-NH2. The pH-independent value for DV/K was 1.0 for both substrates, and DV = 1.5-1.7 and 2.2-3.2 at low and high pH (D), respectively. The attentuation of both V and DV at low pH (D) is consistent with a change in rate-limiting step from a chemical one at high pH (D) to one in which a product release step or an enzyme isomerization step becomes partly rate-limiting at low pH (D). Proton inventory data is in accord with the concerted transfer of two protons in the transition state of a rate-limiting chemical step in which the enzyme-bound amide hydrate adduct collapses to form the carboxylic acid and amine products.  相似文献   

15.
Zheng R  Blanchard JS 《Biochemistry》2000,39(51):16244-16251
Ketopantoate reductase (EC 1.1.1.169) catalyzes the NADPH-dependent reduction of alpha-ketopantoate to D-(-)-pantoate in the biosynthesis of pantothenate. The pH dependence of V and V/K for the E. coli enzyme suggests the involvement of a general acid/base in the catalytic mechanism. To identify residues involved in catalysis and substrate binding, we mutated the following six strictly conserved residues to Ala: Lys72, Lys176, Glu210, Glu240, Asp248, and Glu256. Of these, the K176A and E256A mutant enzymes showed 233- and 42-fold decreases in V(max), and 336- and 63-fold increases in the K(m) value of ketopantoate, respectively, while the other mutants exhibited WT kinetic properties. The V(max) for the K176A and E256A mutant enzymes was markedly increased, up to 25% and 75% of the wild-type level, by exogenously added primary amines and formate, respectively. The rescue efficiencies for the K176A and E256A mutant enzymes were dependent on the molecular volume of rescue agents, as anticipated for a finite active site volume. The protonated form of the amine is responsible for recovery of activity, suggesting that Lys176 functions as a general acid in catalysis of ketopantoate reduction. The rescue efficiencies for the K176A mutant by primary amines were independent of the pK(a) value of the rescue agents (Bronsted coefficient, alpha = -0.004 +/-0.008). Insensitivity to acid strength suggests that the chemical reaction is not rate-limiting, consistent with (a) the catalytic efficiency of the wild-type enzyme (k(cat)/K(m) = 2x10(6) M(-1) s(-1) and (b) the small primary deuterium kinetic isotope effects, (D)V = 1.3 and (D)V/K = 1.5, observed for the wild-type enzyme. Larger primary deuterium isotope effects on V and V/K were observed for the K176A mutant ((D)V = 3.0, (D)V/K = 3.7) but decreased nearly to WT values as the concentration of ethylamine was increased. The nearly WT activity of the E256A mutant in the presence of formate argues for an important role for this residue in substrate binding. The double mutant (K176A/E256A) has no detectable ketopantoate reductase activity. These results indicate that Lys176 and Glu256 of the E. coli ketopantoate reductase are active site residues, and we propose specific roles for each in binding ketopantoate and catalysis.  相似文献   

16.
The neuronal nitric oxide synthase (nNOS) basal and calmodulin- (CaM-) stimulated reduction of 2,6-dichloroindophenol (DCIP) and cytochrome c(3+) follow ping-pong mechanisms [Wolthers and Schimerlik (2001) Biochemistry 40, 4722-4737]. Primary deuterium [NADPH(D)] and solvent deuterium isotope effects on the kinetic parameters were studied to determine rate-limiting step(s) in the kinetic mechanisms for the two substrates. nNOS was found to abstract the pro-R (A-side) hydrogen from NADPH. Values for (D)V and (D)(V/K)(NADPH) were similar for the basal (1.3-1.7) and CaM-stimulated (1.5-2.1) reduction of DCIP, while (D)V (2.1-2.8) was higher than (D)(V/K)(NADPH) (1.1-1.5) for cytochrome c(3+) reduction with and without CaM. This suggests that the rate of the reductive half-reaction (NADPH oxidation) rather than that of the oxidative half-reaction (reduction of DCIP or cytochrome c(3+)) limits the overall reaction rate. A value for (D)(V/K)(NADPH) close to 1 indicates the intrinsic isotope effect on hydride transfer is suppressed by a slower step in the reductive half-reaction. The oxidative half-reaction is insensitive to NADPD isotope effects as both (D)(V/K)(DCIP) and (D)(V/K)(cytc) equal 1 within experimental error. Large solvent kinetic isotope effects (SKIE) observed for (V/K)(cytc) for basal (approximately 8) and CaM-stimulated (approximately 31) reduction of cytochrome c(3+) suggest that proton uptake from the solvent limits the rate of the oxidative half-reaction. This step does not severely limit the overall reaction rate as (D2O)V equaled 2 and (D2O)(V/K)(NADPH) was between 0.9 and 1.3 for basal and CaM-stimulated cytochrome c(3+) reduction.  相似文献   

17.
Heredia VV  Penning TM 《Biochemistry》2004,43(38):12028-12037
3Alpha-hydroxysteroid dehydrogenases (3alpha-HSDs) catalyze the interconversion between 5alpha-dihydrotestosterone (5alpha-DHT), the most potent androgen, and 3alpha-androstanediol (3alpha-diol), a weak androgen metabolite. To identify the rate-determining step in this physiologically important reaction, rat liver 3alpha-HSD (AKR1C9) was used as the protein model for the human homologues in fluorescence stopped-flow transient kinetic and kinetic isotope effect studies. Using single and multiple turnover experiments to monitor the NADPH-dependent reduction of 5alpha-DHT, it was found that k(lim) and k(max) values were identical to k(cat), indicating that chemistry is rate-limiting overall. Kinetic isotope effect measurements, which gave (D)k(cat) = 2.4 and (D)2(O)k(cat) = 3.0 at pL 6.0, suggest that the slow chemical transformation is significantly rate-limiting. When the NADP(+)-dependent oxidation of 3alpha-diol was monitored, single and multiple turnover experiments showed a k(lim) and burst kinetics consistent with product release as being rate-limiting overall. When NAD(+) was substituted for NADP(+), burst phase kinetics was eliminated, and k(max) was identical to k(cat). Thus with the physiologically relevant substrates 5alpha-DHT plus NADPH and 3alpha-diol plus NAD(+), the slowest event is chemistry. R276 forms a salt-linkage with the phosphate of 2'-AMP, and when it is mutated, tight binding of NAD(P)H is no longer observed [Ratnam, K., et al. (1999) Biochemistry 38, 7856-7864]. The R276M mutant also eliminated the burst phase kinetics observed for the NADP(+)-dependent oxidation of 3alpha-diol. The data with the R276M mutant confirms that the release of the NADPH product is the slow event; and in its absence, chemistry becomes rate-limiting. W227 is a critical hydrophobic residue at the steroid binding site, and when it is mutated to alanine, k(cat)/K(m) for oxidation is significantly depressed. Burst phase kinetics for the NADP(+)-dependent turnover of 3alpha-diol by W227A was also abolished. In the W227A mutant, the slow release of NADPH is no longer observed since the chemical transformation is now even slower. Thus, residues in the cofactor and steroid-binding site can alter the rate-determining step in the NADP(+)-dependent oxidation of 3alpha-diol to make chemistry rate-limiting overall.  相似文献   

18.
We have investigated the dynamics of proton equilibration within the proton-transfer pathway of cytochrome c oxidase from bovine heart that is used for the transfer of both substrate and pumped protons during reaction of the reduced enzyme with oxygen (D-pathway). The kinetics of the slowest phase in the oxidation of the enzyme (the oxo-ferryl --> oxidized transition, F --> O), which is associated with proton uptake, were studied by monitoring absorbance changes at 445 nm. The rate constant of this transition, which is 800 s(-)(1) in H(2)O (at pH approximately 7.5), displayed a kinetic deuterium isotope effect of approximately 4 (i.e., the rate was approximately 200 s(-)(1) in 100% D(2)O). To investigate the kinetics of the onset of the deuterium isotope effect, fully reduced, solubilized CO-bound cytochrome c oxidase in H(2)O was mixed rapidly at a ratio of 1:5 with a D(2)O buffer saturated with oxygen. After a well-defined time period, CO was flashed off using a short laser flash. The time between mixing and flashing off CO was varied within the range 0. 04-10 s. The results show that for the bovine enzyme, the onset of the deuterium isotope effect takes place within two time windows of O transition is internal proton transfer from a site, proposed to be Glu (I-286) (R. sphaeroides amino acid residue numbering), to the binuclear center. The spontaneous equilibration of protons/deuterons with this site in the interior of the protein is slow (approximately 1 s).  相似文献   

19.
The steady-state kinetic parameters for epimerization of UDP-galactose by UDP-galactose 4-epimerase from Escherichia coli (GalE), Y149F-GalE, and S124A-GalE have been measured as a function of pH. The deuterium kinetic isotope effects for epimerization of UDP-galactose-C-d(7) by these enzymes have also been measured. The results show that the activity of wild-type GalE is pH-independent in the pH range of 5.5-9.3, and there is no significant deuterium kinetic isotope effect in the reaction of UDP-galactose-C-d(7). It is concluded that the rate-limiting step for epimerization by wild-type GalE is not hydride transfer and must be either a diffusional process or a conformational change. Epimerization of UDP-galactose-C-d(7) by Y149F-GalE proceeds with a pH-dependent deuterium kinetic isotope effect on k(cat) of 2.2 +/- 0.4 at pH 6.2 and 1.1 +/- 0.5 at pH 8.3. Moreover, the plot of log k(cat)/K(m) breaks downward on the acid side with a fitted value of 7.1 for the pK(a). It is concluded that the break in the pH-rate profile arises from a change in the rate-limiting step from hydride transfer at low pH to a conformational change at high pH. Epimerization of UDP-galactose-C-d(7) by S124A-GalE proceeds with a pH-independent deuterium kinetic isotope effect on k(cat) of 2.0 +/- 0.2 between pH 6 and 9. Both plots of log k(cat) and log k(cat)/K(m) display pH dependence. The plot of log k(cat) versus pH breaks downward with a pK(a) of 6.35 +/- 0.10. The plot of log k(cat)/K(m) versus pH is bell-shaped, with fitted pK(a) values of 6.76 +/- 0.09 and 9.32 +/- 0.21. It is concluded that hydride transfer is rate-limiting, and the pK(a) of 6.7 for free S124A-GalE is assigned to Tyr 149, which displays the same value of pK(a) when measured spectrophotometrically in this variant. Acid-base catalysis by Y149F-GalE is attributed to Ser 124, which is postulated to rescue catalysis of proton transfer in the absence of Tyr 149. The kinetic pK(a) of 7.1 for free Y149F-GalE is lower than that expected for Ser 124, as proven by the pH-dependent kinetic isotope effect. Epimerization by the doubly mutated Y149F/S124A-GalE proceeds at a k(cat) that is lower by a factor of 10(7) than that of wild-type GalE. This low rate is attributed to the synergistic actions of Tyr 149 and Ser 124 in wild-type GalE and to the absence of any internal catalysis of hydride transfer in the doubly mutated enzyme.  相似文献   

20.
Electron nuclear double resonance (ENDOR) of protons at Type 2 and Type 1 cupric active sites correlates with the enzymatic pH dependence, the mutation of nearby conserved, nonligating residues, and electron transfer in heterologously expressed Rhodobacter sphaeroides nitrite reductase. Wild-type enzyme showed a pH 6 activity maximum but no kinetic deuterium isotope effect, suggesting protons are not transferred in the rate-limiting step of nitrite reduction. However, protonatable Asp129 and His287, both located near the Type 2 center, modulated enzyme activity. ENDOR of the wild-type Type 2 center at pH 6.0 revealed an exchangeable proton with large hyperfine coupling. Dipolar distance estimates indicated that this proton was 2.50-2.75 or 2.25-2.45 A from Type 2 copper in the presence or absence of nitrite, respectively. This proton may provide a properly oriented hydrogen bond to enhance water formation upon nitrite reduction. This proton was eliminated at pH 5.0 and showed a diminished coupling at pH 7.5. Mutations of Asp129 and His287 reduced enzyme activity and altered the exchangeable proton hyperfine spectra. Mutation of Asp129 prevented a pH-dependent change at the Type 1 Cys167 ligand as observed by Cys C(beta) proton ENDOR, implying there is a Type 2 and pH-dependent alteration of the Type 1 center. Mutation of the Type 1 center ligand Met182 to Thr and mutation of Asp129 increased the activation energy for nitrite reduction. Involvement of both the Type 1 center and Asp129 in modulating activation energy shows that electron transfer from the Type 1 center to a nitrite-ligated Type 2 center is rate-limiting for nitrite reduction. Mutation of Ile289 to Ala and Val caused minor perturbation to enzyme activity, but as detected by ENDOR, allowed formate binding. Thus, bulky Ile289 may exclude non-nitrite ligands from the Type 2 active site.  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