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.
We performed a series of site-directed mutagenesis experiments of catalytic antibody, 6D9, which hydrolyzes a prodrug of chloramphenicol, based on our previous directed evolution study [Takahashi et al. (2001) Nat. Biotechnol. 19, 563-567]. Since we previously found that the variants with a mutation of Ser(L27e)Tyr afforded a one order of magnitude increase in catalytic rate, we created a site-directed mutant containing this mutation. The resulting mutant, 6D9-Ser(L27e)Tyr, had 6.5-fold higher k(cat)/k(uncat) and 9.8-fold higher k(cat)/K(m) than wild-type 6D9. We also created 6D9-Thr(L27a)Pro, since this mutation occurred frequently in the previous directed evolution, and it had 2.1-fold higher k(cat)/k(uncat) and k(cat)/K(m) than 6D9. Kinetic and computational analyses suggest that Tyr at L27e contributes to transition-state stabilization, while Pro at L27a does not interact with the transition-state structure directly, but obviously contributes to enhanced catalytic activity. Including double mutants that combined favourable substitutions, we created seven site-directed mutants. However, none of them had higher catalytic activities than some of highly improved variants obtained in the previous directed evolution. The present study gives direct evidence that not only a specific amino acid residue which obviously contributes to transition-state stabilization, but also a group of amino acid residues working in concert is important for efficient catalysis of a given transformation.  相似文献   

2.
Enzymes enhance chemical reaction rates by lowering the activation energy, the energy barrier of the reaction leading to products. This occurs because enzymes bind the high-energy intermediate of the reaction (the transition state) more strongly than the substrate. We studied details of this process by determining the substrate binding energy (DeltaG(s), calculated from K(m) values) and the activation energy (DeltaG(T), determined from k(cat)/K(m) values) for the trypsin-catalyzed hydrolysis of oligopeptides. Plots of DeltaG(T) versus DeltaG(s) for oligopeptides with 15 amino acid replacements at each of the positions P(1)', P(1), and P(2) were straight lines, as predicted by a derived equation that relates DeltaG(T) and DeltaG(s). The data led to the conclusion that the trypsin active site has subsites that bind moieties of substrate and of transition state in characteristic ratios, whichever substrate is used. This was unexpected and means that each subsite characteristically favors substrate binding or catalysis.  相似文献   

3.
Antibody AZ-28 was generated against the chairlike transition-state analogue (TSA) 1 and catalyzes the oxy-Cope rearrangement of substrate 2 to product 3. The germline precursor to AZ-28 catalyzes the reaction with a 35-fold higher rate (k(cat)/k(uncat) = 163 000), despite a 40-fold lower binding affinity for TSA.1 (K(D) = 670 nM). To determine the structural basis for the differences in the binding and catalytic properties of the germline and affinity-matured antibodies, the X-ray crystal structures of the unliganded and TSA.1 complex of antibody AZ-28 have been determined at 2.8 and 2.6 A resolution, respectively; the structures of the unliganded and TSA.1 complex of the germline precursor to AZ-28 were both determined at 2. 0 A resolution. In the affinity-matured antibody.hapten complex the TSA is fixed in a catalytically unfavorable conformation by a combination of van der Waals and hydrogen-bonding interactions. The 2- and 5-phenyl substituents of TSA.1 are almost perpendicular to the cyclohexyl ring, leading to decreased orbital overlap and decreased stabilization of the putative transition state. The active site of the germline antibody appears to have an increased degree of flexibility-CDRH3 moves 4.9 A outward from the active site upon binding of TSA.1. We suggest that this conformational flexibility in the germline antibody, which results in a lower binding affinity for TSA.1, allows dynamic changes in the dihedral angle of the 2-phenyl substituent along the reaction coordinate. These conformational changes in turn lead to enhanced orbital overlap and increased catalytic rate. These studies suggest that protein and substrate dynamics play a key role in this antibody-catalyzed reaction.  相似文献   

4.
Considering the dynamic nature of CYPs, methods that reveal information about substrate and enzyme dynamics are necessary to generate predictive models. To compare substrate dynamics in CYP2E1 and CYP2A6, intramolecular isotope effect experiments were conducted, using deuterium labeled substrates: o-xylene, m-xylene, p-xylene, 2,6-dimethylnaphthalene, and 4,4'-dimethylbiphenyl. Competitive intermolecular experiments were also conducted using d(0)- and d(6)-labeled p-xylene. Both CYP2E1 and CYP2A6 displayed full isotope effect expression for o-xylene oxidation and almost complete suppression for dimethylbiphenyl. Interestingly, (k(H)/k(D))(obs) for d(3)-p-xylene oxidation ((k(H)/k(D))(obs)=6.04 and (k(H)/k(D))(obs)=5.53 for CYP2E1 and CYP2A6, respectively) was only slightly higher than (k(H)/k(D))(obs) for d(3)-dimethylnaphthalene ((k(H)/k(D))(obs)=5.50 and (k(H)/k(D))(obs)=4.96, respectively). One explanation is that in some instances (k(H)/k(D))(obs) values are generated by the presence of two substrates-bound simultaneously to the CYP. Speculatively, if this explanation is valid, then intramolecular isotope effect experiments should be useful in the mechanistic investigation of P450 cooperativity.  相似文献   

5.
This report described that a hapten of racemic phosphonate 3 designed as the mimic of the transition state of hydrolysis of naproxen ethyl ester was successfully synthesized from easily available 2-acetyl-6-methoxy-naphthalene 5. Then BALB/C mice were immunized and one of the monoclonal catalytic antibodies, N116-27, which enantioselectively accelerated the hydrolysis of the R-(-)-naproxen ethyl ester was given. The Michaelis-Menton parameter for the catalyzed reaction was K(M)=6.67 mM and k(cat)/k(uncat)=5.8 x 10(4). This enantioselective result was explained by the fact that the R-isomer of rac-hapten was more immunogenic than the S-isomer.  相似文献   

6.
Histamine dehydrogenase (NSHADH) can be isolated from cultures of Nocardioides simplex grown with histamine as the sole nitrogen source. A previous report suggested that NSHADH might contain the quinone cofactor tryptophan tryptophyl quinone (TTQ). Here, the hdh gene encoding NSHADH is cloned from the genomic DNA of N. simplex, and the isolated enzyme is subjected to a full spectroscopic characterization. Protein sequence alignment shows NSHADH to be related to trimethylamine dehydrogenase (TMADH: EC 1.5.99.7), where the latter contains a bacterial ferredoxin-type [4Fe-4S] cluster and 6-S-cysteinyl FMN cofactor. NSHADH has no sequence similarity to any TTQ containing amine dehydrogenases. NSHADH contains 3.6+/-0.3 mol Fe and 3.7+/-0.2 mol acid labile S per subunit. A comparison of the UV/vis spectra of NSHADH and TMADH shows significant similarity. The EPR spectrum of histamine reduced NSHADH also supports the presence of the flavin and [4Fe-4S] cofactors. Importantly, we show that NSHADH has a narrow substrate specificity, oxidizing only histamine (K(m)=31+/-11 microM, k(cat)/K(m)=2.1 (+/-0.4)x10(5)M(-1)s(-1)), agmatine (K(m)=37+/-6 microM, k(cat)/K(m)=6.0 (+/-0.6)x10(4)M(-1)s(-1)), and putrescine (K(m)=1280+/-240 microM, k(cat)/K(m)=1500+/-200 M(-1)s(-1)). A kinetic characterization of the oxidative deamination of histamine by NSHADH is presented that includes the pH dependence of k(cat)/K(m) (histamine) and the measurement of a substrate deuterium isotope effect, (D)(k(cat)/K(m) (histamine))=7.0+/-1.8 at pH 8.5. k(cat) is also pH dependent and has a reduced substrate deuterium isotope of (D)(k(cat))=1.3+/-0.2.  相似文献   

7.
The catalytic antibody 6D9, which was raised against a transition-state analogue (TSA), catalyzes the hydrolysis of a non-bioactive chloramphenicol monoester to generate chloramphenicol. It has been shown that 6D9 utilizes the binding affinity in the catalysis; the differential affinity of the TSA relative to the substrate is equal to the rate enhancement. To reveal the recognition mechanism of 6D9 for the TSA and the substrate, we performed NMR analysis of the Fv fragment of 6D9 (6D9-Fv), together with site-directed mutagenesis and stopped-flow kinetic analyses. Among six 6D9-Fv mutants, Y58(H)A and W100i(H)A displayed significant reductions in their affinities to the TSA, while their substrate-binding affinities were identical with that of the wild-type 6D9-Fv. The stopped-flow kinetic studies revealed that the TSA binding to 6D9-Fv occurred by an induced-fit mechanism. In contrast, no induced-fit type of TSA-binding mechanism was observed for Y58(H)A and W100i(H)A. From NMR experiments, we identified the residues with chemical shifts that were perturbed by the ligand-binding. The residues affected by the TSA binding were located on the TSA-binding site determined by the X-ray study, and on the regions far from the binding site. On the other hand, the residues affected by the substrate binding were localized on the TSA-binding site. As for W100i(H)A, no residue other than those in the binding site was affected by the ligand binding. On the basis of these results and the crystal structure, we concluded that the TSA binding induced a conformational change involving the formation of aromatic-aromatic interactions and a hydrogen bond. These interactions can account for the differential affinity for the TSA relative to the substrate. W100i(H) probably plays an important role in inducing the conformational changes. The present NMR studies have enabled us to visualize the concept of transition-state stabilization in enzymatic catalysis, in which the transition-state contacts are better than those of the substrate.  相似文献   

8.
The solvent kinetic isotope effects (SKIE) on the yeast alpha-glucosidase-catalyzed hydrolysis of p-nitrophenyl and methyl-d-glucopyranoside were measured at 25 degrees C. With p-nitrophenyl-D-glucopyranoside (pNPG), the dependence of k(cat)/K(m) on pH (pD) revealed an unusually large (for glycohydrolases) solvent isotope effect on the pL-independent second-order rate constant, (DOD)(k(cat)/K(m)), of 1.9 (+/-0.3). The two pK(a)s characterizing the pH profile were increased in D(2)O. The shift in pK(a2) of 0.6 units is typical of acids of comparable acidity (pK(a)=6.5), but the increase in pK(a1) (=5.7) of 0.1 unit in going from H(2)O to D(2)O is unusually small. The initial velocities show substrate inhibition (K(is)/K(m) approximately 200) with a small solvent isotope effect on the inhibition constant [(DOD)K(is)=1.1 (+/-0.2)]. The solvent equilibrium isotope effects on the K(is) for the competitive inhibitors D-glucose and alpha-methyl D-glucoside are somewhat higher [(DOD)K(i)=1.5 (+/-0.1)]. Methyl glucoside is much less reactive than pNPG, with k(cat) 230 times lower and k(cat)/K(m) 5 x 10(4) times lower. The solvent isotope effect on k(cat) for this substrate [=1.11 (+/-0. 02)] is lower than that for pNPG [=1.67 (+/-0.07)], consistent with more extensive proton transfer in the transition state for the deglucosylation step than for the glucosylation step.  相似文献   

9.
To elucidate the interaction between substrate inhibition and substrate transglycosylation of retaining glycoside hydrolases (GHs), a steady-state kinetic study was performed for the GH family 3 glucan (1-->3)-beta-glucosidase from the white-rot fungus Phanerochaete chrysosporium, using laminarioligosaccharides as substrates. When laminaribiose was incubated with the enzyme, a transglycosylation product was detected by thin-layer chromatography. The product was purified by size-exclusion chromatography, and was identified as a 6-O-glucosyl-laminaribiose (beta-D-Glcp-(1-->6)-beta-D-Glcp-(1-->3)-D-Glc) by 1H NMR spectroscopy and electrospray ionization mass spectrometry analysis. In steady-state kinetic studies, an apparent decrease of laminaribiose hydrolysis was observed at high concentrations of the substrate, and the plots of glucose production versus substrate concentration were thus fitted to a modified Michaelis-Menten equation including hydrolytic and transglycosylation parameters (K(m), K(m2), k(cat), k(cat2)). The rate of 6-O-glucosyl-laminaribiose production estimated by high-performance anion-exchange chromatography coincided with the theoretical rate calculated using these parameters, clearly indicating that substrate inhibition of this enzyme is fully explained by substrate transglycosylation. Moreover, when K(m), k(cat), and affinity for glucosyl-enzyme intermediates (K(m2)) were estimated for laminarioligosaccharides (DP=3-5), the K(m) value of laminaribiose was approximately 5-9 times higher than those of the other oligosaccharides (DP=3-5), whereas the K(m2) values were independent of the DP of the substrates. The kinetics of transglycosylation by the enzyme could be well interpreted in terms of the subsite affinities estimated from the hydrolytic parameters (K(m) and k(cat)), and a possible mechanism of transglycosylation is proposed.  相似文献   

10.
Case A  Stein RL 《Biochemistry》2006,45(7):2443-2452
Ubiquitin C-terminal hydrolases (UCHs) cleave Ub-X bonds (Ub is ubiquitin and X an alcohol, an amine, or a protein) through a thioester intermediate that is produced by nucleophilic attack of the Cys residue of a Cys-SH/His-Im catalytic diad. We are studying the mechanism of UCH-L1, a UCH that is implicated in Parkinson's disease, and now wish to report our initial findings. (i) Pre-steady-state kinetic studies for UCH-L1-catalyzed hydrolysis of Ub-AMC (AMC, 7-amido-4-methylcoumarin) indicate that k(cat) is rate-limited by acyl-enzyme formation. Thus, K(m) = K(s), the dissociation constant for the Michaelis complex, and k(cat) = k(2), the rate constant for acyl-enzyme formation. (ii) For K(assoc) (=K(s)(-)(1)), DeltaC(p) = -0.8 kcal mol(-)(1) deg(-)(1) and is consistent with coupling between substrate association and a conformational change of the enzyme. For k(2), DeltaS(++) = 0 and suggests that in the E-S, substrate and active site residues are precisely aligned for reaction. (iii) Solvent isotope effects are (D)K(assoc) = 0.5 and (D)k(2) = 0.9, suggesting that the substrate binds to a form of free enzyme in which the active site Cys exists as the thiol. In the resultant Michaelis complex, the diad has tautomerized to ion pair Cys-S(-)/His-ImH(+). Subsequent attack of thiolate produces the acyl-enzyme species. In contrast, isotope effects for association of UCH-L1 with transition-state analogue ubiquitin aldehyde suggest that an alternative mechanistic pathway can sometimes be available to UCH-L1 involving general base-catalyzed attack of Cys-SH by His-Im.  相似文献   

11.
We investigate the effects of detergent on the kinetics and oligomeric state of allene oxide synthase (AOS) from Arabidopsis thaliana (CYP74A1). We show that detergent-free CYP74A1 is monomeric and highly water soluble with dual specificity, but has relatively low activity. Detergent micelles promote a 48-fold increase in k(cat)/K(m) (to 5.9 x 10(7)M(-1)s(-1)) with concomitant changes in the spin state equilibrium of the haem-iron due to the binding of a single detergent micelle to the protein monomer, which is atypical of P450 enzymes. This mechanism is shown to be an important determinant of the substrate specificity of CYP74A1. CYP74A1 may be suited for structural resolution of the first plant cytochrome P450 and its 9-AOS activity and behaviour in vitro has implications for its role in planta.  相似文献   

12.
Sierks MR  Svensson B 《Biochemistry》2000,39(29):8585-8592
Molecular recognition using a series of deoxygenated maltose analogues was used to determine the substrate transition-state binding energy profiles of 10 single-residue mutants at the active site of glucoamylase from Aspergillus niger. The individual contribution of each substrate hydroxyl group to transition-state stabilization with the wild type and each mutant GA was determined from the relation Delta(DeltaG()) = -RT ln[(k(cat)/K(M))(x)/(k(cat)/K(M))(y)], where x represents either a mutant enzyme or substrate analogue and y the wild-type enzyme or parent substrate. The resulting binding energy profiles indicate that disrupting an active site hydrogen bond between enzyme and substrate, as identified in crystal structures, not only sharply reduces or eliminates the energy contributed from that particular hydrogen bond but also perturbs binding contributions from other substrate hydroxyl groups. Replacing the active site acidic groups, Asp55, Glu180, or Asp309, with the corresponding amides, and the neutral Trp178 with the basic Arg, all substantially reduced the binding energy contribution of the 4'- and 6'-OH groups of maltose at subsite -1, even though both Glu180 and Asp309 are localized at subsite 1. In contrast, the substitution, Asp176 --> Asn, located near subsites -1 and 1, did not substantially perturb any of the individual hydroxyl group binding energies. Similarly, the substitutions Tyr116 --> Ala, Ser119 --> Tyr, or Trp120 --> Phe also did not substantially alter the energy profiles even though Trp120 has a critical role in directing conformational changes necessary for activity. Since the mutations at Trp120 and Asp176 reduced k(cat) values by 50- and 12-fold, respectively, a large effect on k(cat) is not necessarily accompanied by changes in hydroxyl group binding energy contributions. Two substitutions, Asn182 --> Ala and Tyr306 --> Phe, had significant though small effects on interactions with 3- and 4'-OH, respectively. Binding interactions between the enzyme and the glucosyl group in subsite -1, particularly with the 4'- and 6'-OH groups, play an important role in substrate binding, while subsite 1 interactions may play a more important role in product release.  相似文献   

13.
To obtain a clearer understanding of the forces involved in transition state stabilization by Escherichia coli cytidine deaminase, we investigated the thermodynamic changes that accompany substrate binding in the ground state and transition state for substrate hydrolysis. Viscosity studies indicate that the action of cytidine deaminase is not diffusion-limited. Thus, K(m) appears to be a true dissociation constant, and k(cat) describes the chemical reaction of the ES complex, not product release. Enzyme-substrate association is accompanied by a loss of entropy and a somewhat greater release of enthalpy. As the ES complex proceeds to the transition state (ES), there is little further change in entropy, but heat is taken up that almost matches the heat that was released with ES formation. As a result, k(cat)/K(m) (describing the overall conversion of the free substrate to ES is almost invariant with changing temperature. The free energy barrier for the enzyme-catalyzed reaction (k(cat)/K(m)) is much lower than that for the spontaneous reaction (k(non)) (DeltaDeltaG = -21.8 kcal/mol at 25 degrees C). This difference, which also describes the virtual binding affinity of the enzyme for the activated substrate in the transition state (S), is almost entirely enthalpic in origin (DeltaDeltaH = -20.2 kcal/mol), compatible with the formation of hydrogen bonds that stabilize the ES complex. Thus, the transition state affinity of cytidine deaminase increases rapidly with decreasing temperature. When a hydrogen bond between Glu-91 and the 3'-hydroxyl moiety of cytidine is disrupted by truncation of either group, k(cat)/K(m) and transition state affinity are each reduced by a factor of 10(4). This effect of mutation is entirely enthalpic in origin (DeltaDeltaH approximately 7.9 kcal/mol), somewhat offset by a favorable change in the entropy of transition state binding. This increase in entropy is attributed to a loss of constraints on the relative motions of the activated substrate within the ES complex. In an Appendix, some objections to the conventional scheme for transition state binding are discussed.  相似文献   

14.
Protein engineers use a variety of mutagenic strategies to adapt enzymes to novel substrates. Directed evolution techniques (random mutagenesis and high-throughput screening) offer a systematic approach to the management of protein complexity. This sub-discipline was galvanized by the invention of DNA shuffling, a procedure that randomly recombines point mutations in vitro. In one influential study, Escherichia coli beta-galactosidase (BGAL) variants with enhanced beta-fucosidase activity (tenfold increase in k(cat)/K(M) in reactions with the novel para-nitrophenyl-beta-d-fucopyranoside substrate; 39-fold decrease in reactivity with the "native"para-nitrophenyl-beta-d-galactopyranoside substrate) were evolved in seven rounds of DNA shuffling and screening. Here, we show that a single round of site-saturation mutagenesis and screening enabled the identification of beta-fucosidases that are significantly more active (180-fold increase in k(cat)/K(M) in reactions with the novel substrate) and specific (700,000-fold inversion of specificity) than the best variants in the previous study. Site-saturation mutagenesis thus proved faster, less resource-intensive and more effective than DNA shuffling for this particular evolutionary pathway.  相似文献   

15.
Thermal denaturation curves of ribonuclease-A were measured by monitoring changes in the far-UV circular dichroism (CD) spectra in the presence of different concentrations of six sugars (glucose, fructose, galactose, sucrose, raffinose and stachyose) and mixture of monosaccharide constituents of each oligosaccharide at various pH values in the range of 6.0-2.0. These measurements gave values of T(m) (midpoint of denaturation), DeltaH(m) (enthalpy change at T(m)), DeltaC(p) (constant-pressure heat capacity change) under a given solvent condition. Using these values of DeltaH(m), T(m) and DeltaC(p) in appropriate thermodynamic relations, thermodynamic parameters at 25 degrees C, namely, DeltaG(D)(o) (Gibbs energy change), DeltaH(D)(o) (enthalpy change), and DeltaS(D)(o) (entropy change) were determined at a given pH and concentration of each sugar (including its mixture of monosaccharide constituents). Our main conclusions are: (i) each sugar stabilizes the protein in terms of T(m) and DeltaG(D)(o), and this stabilization is under enthalpic control, (ii) the protein stabilization by the oligosaccharide is significantly less than that by the equimolar concentration of the constituent monosaccharides, and (iii) the stabilization by monosaccharides in a mixture is fully additive. Furthermore, measurements of the far- and near-UV CD spectra suggested that secondary and tertiary structures of protein in their native and denatured states are not perturbed on the addition of sugars.  相似文献   

16.
NAD(+)-dependent 15-hydroxyprostaglandin dehydrogenase (15-PGDH), a member of the short-chain dehydrogenase/reductase (SDR) family, catalyzes the first step in the catabolic pathways of prostaglandins and lipoxins, and is believed to be the key enzyme responsible for the biological inactivation of these biologically potent eicosanoids. The enzyme utilizes NAD(+) specifically as a coenzyme. Potential amino acid residues involved in binding NAD(+) and facilitating enzyme catalysis have been partially identified. In this report, we propose that three more residues in 15-PGDH, Ile-17, Asn-91, and Val-186, are also involved in the interaction with NAD(+). Site-directed mutagenesis was used to examine their roles in binding NAD(+). Several mutants (I17A, I17V, I17L, I17E, I17K, N91A, N91D, N91K, V186A, V186I, V186D, and V186K) were prepared, expressed as glutathione S-transferase (GST) fusion enzymes in Escherichia coli, and purified by GSH-agarose affinity chromatography. Mutants I17E, I17K, N91L, N91K, and V186D were found to be inactive. Mutants N91A, N91D, V186A, and V186K exhibited comparable activities to the wild type enzyme. However, mutants I17A, I17V, I17L, and V186I had higher activity than the wild type. Especially, the activities of I17L and V186I were increased nearly 4- and 5-fold, respectively. The k(cat)/K(m) ratios of all active mutants for PGE(2) were similar to that of the wild type enzyme. However, the k(cat)/K(m) ratios of mutants I17A and N91A for NAD(+) were decreased 5- and 10-fold, respectively, whereas the k(cat)/K(m) ratios of mutants I17V, N91D, V186I, and V186K for NAD(+) were comparable to that of the wild type enzyme. The k(cat)/K(m) ratios of mutants I17L and V186A for NAD(+) were increased over nearly 2-fold. These results suggest that Ile-17, Asn-91, and Val-186 are involved in the interaction with NAD(+) and contribute to the full catalytic activity of 15-PGDH.  相似文献   

17.
Clouet A  Darbre T  Reymond JL 《Biopolymers》2006,84(1):114-123
A 65,536-member combinatorial library of peptide dendrimers was prepared by split-and-mix synthesis and screened on solid support for esterolytic activity in aqueous buffer using 8-butyryloxypyrene-1,3,6-trisulfonate (2) as a fluorogenic substrate. Active sequences were identified by analysis of fluorescent beads. The corresponding dendrimers were resynthesized by solid-phase synthesis, cleaved from the resin, and purified by preparative reverse-phase HPLC. The dendrimers showed the expected catalytic activity in aqueous buffer. Catalysis was studied against a pannel of fluorogenic 8-acyloxypyrene-1,3,6-trisulfonate substrates. The catalytic peptide dendrimers display enzyme-like kinetics in aqueous buffer with substrate binding in the range K(M) approximately 0.1 mM, catalytic rate constants k(cat) approximately 0.1 min(-1), and specific rate accelerations over background up to k(cat)/k(uncat) = 10,000.  相似文献   

18.
Hydrostatic pressure causes biphasic effects on the oxidation of alcohols by yeast alcohol dehydrogenase as expressed on the kinetic parameter V/K which measures substrate capture. Moderate pressure increases capture by activating hydride transfer, whose transition-state must therefore have a smaller volume than the free alcohol plus the capturing form of enzyme, with DeltaV(double dagger)=-30 mL mol(-1) for isopropanol. A comparison of these effects with those on the oxidation of deutero-isopropanol generates a monophasic decrease in the intrinsic isotope effect; therefore, the volume of activation for the transition-state of deuteride transfer must be even more negative, by 7.6 mL mol(-1). The pressure data extrapolate and factor the kinetic isotope effect into a semi-classical reactant-state component, with a null value of k(H)/k(D)=1, and a transition-state component of Q(H)/Q(D)=4, suggestive of hydrogen tunneling. Pressures above 1.5 kbar decrease capture by favoring a minor conformation of enzyme which binds nicotinamide adenine dinucleotide (NAD(+)) less tightly. This inactive conformation has a smaller volume than active E-NAD(+), with a difference of 74 mL mol(-1) and an equilibrium constant of 93 between them, at one atmosphere of pressure. These results are virtually identical to those obtained with benzyl alcohol and give credence to this method of analysis. Moreover, qualitatively similar results with greater pressure sensitivity but less precision are obtained using ethanol as a substrate, only with pressure driving the value of the isotope effect to a value less than (D)k=1.03 directly, without extrapolation. The ethanol data verify the most surprising finding of these studies, namely that the entire kinetic isotope effect arises from a transition-state phenomenon.  相似文献   

19.
MDL 72527 was considered a selective inhibitor of FAD-dependent polyamine oxidases. In the present communication, we demonstrate that MDL 72527 inactivates bovine serum amine oxidase, a copper-containing, TPQ-enzyme, time-dependently at 25 degrees C. In striking contrast, the enzyme remained active after incubation with excessive MDL 72527 at 37 degrees C, even after 70 h of incubation. Inactivation of BSAO with MDL 72527 at 25 degrees C did not involve the cofactor, as was shown by spectroscopy and by reaction with phenylhydrazine. Docking of MDL 72527 is difficult, owing to its size and two lipophilic moieties, and it has been shown that minor changes in reaction rate of substrates cause major changes in K(m) and k(cat)/K(m). We hypothesise that subtle conformational changes between 25 and 37 degrees C impair MDL 72527 from productive binding and prevent the nucleophilic group from reacting with the double bond system.  相似文献   

20.
Bovine pancreatic ribonuclease A (RNase A) catalyzes the cleavage of the P-O(5') bond in RNA. Although this enzyme has been the object of much landmark work in bioorganic chemistry, the nature of its rate-limiting transition state and its catalytic rate enhancement had been unknown. Here, the value of k(cat)/K(m) for the cleavage of UpA by wild-type RNase A was found to be inversely related to the concentration of added glycerol. In contrast, the values of k(cat)/K(m) for the cleavage of UpA by a sluggish mutant of RNase A and the cleavage of the poor substrate UpOC(6)H(4)-p-NO(2) by wild-type RNase A were found to be independent of glycerol concentration. Yet, UpA cleavage by the wild-type and mutant enzymes was found to have the same dependence on sucrose concentration, indicating that catalysis of UpA cleavage by RNase A is limited by desolvation. The rate of UpA cleavage by RNase A is maximal at pH 6.0, where k(cat) = 1.4 × 10(3) s(-1) and k(cat)/K(m) = 2.3 × 10(6) M(-1)s(-1) at 25°C. At pH 6.0 and 25°C, the uncatalyzed rate of [5,6-(3)H]Up[3,5,8-(3)H]A cleavage was found to be k(uncat) = 5 × 10(-9) s(-1) (t(1/2) = 4 years). Thus, RNase A enhances the rate of UpA cleavage by 3 × 10(11)-fold by binding to the transition state for P-O(5') bond cleavage with a dissociation constant of <2 × 10(-15) M.  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