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.
Catalytic reaction pathway for the mitogen-activated protein kinase ERK2   总被引:2,自引:0,他引:2  
Prowse CN  Hagopian JC  Cobb MH  Ahn NG  Lew J 《Biochemistry》2000,39(20):6258-6266
The structural, functional, and regulatory properties of the mitogen-activated protein kinases (MAP kinases) have long attracted considerable attention owing to the critical role that these enzymes play in signal transduction. While several MAP kinase X-ray crystal structures currently exist, there is by comparison little mechanistic information available to correlate the structural data with the known biochemical properties of these molecules. We have employed steady-state kinetic and solvent viscosometric techniques to characterize the catalytic reaction pathway of the MAP kinase ERK2 with respect to the phosphorylation of a protein substrate, myelin basic protein (MBP), and a synthetic peptide substrate, ERKtide. A minor viscosity effect on k(cat) with respect to the phosphorylation of MBP was observed (k(cat) = 10 +/- 2 s(-1), k(cat)(eta) = 0.18 +/- 0.05), indicating that substrate processing occurs via slow phosphoryl group transfer (12 +/- 4 s(-1)) followed by the faster release of products (56 +/- 4 s(-1)). At an MBP concentration extrapolated to infinity, no significant viscosity effect on k(cat)/K(m(ATP)) was observed (k(cat)/K(m(ATP)) = 0.2 +/- 0.1 microM(-1) s(-1), k(cat)/K(m(ATP))(eta) = -0.08 +/- 0.04), consistent with rapid-equilibrium binding of the nucleotide. In contrast, at saturating ATP, a full viscosity effect on k(cat)/K(m) for MBP was apparent (k(cat)/K(m(MBP)) = 2.4 +/- 1 microM(-1) s(-1), k(cat)/K(m(MBP))(eta) = 1.0 +/- 0.1), while no viscosity effect was observed on k(cat)/K(m) for the phosphorylation of ERKtide (k(cat)/K(m(ERKtide)) = (4 +/- 2) x 10(-3) microM(-1) s(-1), k(cat)/K(m(ERKtide))(eta) = -0.02 +/- 0.02). This is consistent with the diffusion-limited binding of MBP, in contrast to the rapid-equilibrium binding of ERKtide, to form the ternary Michaelis complex. Calculated values for binding constants show that the estimated value for K(d(MBP)) (/= 1.5 mM). The dramatically higher catalytic efficiency of MBP in comparison to that of ERKtide ( approximately 600-fold difference) is largely attributable to the slow dissociation rate of MBP (/=56 s(-1)), from the ERK2 active site.  相似文献   

2.
The control of the catalytic power and fidelity of DNA polymerases involves the complex combined effect of the protein residues, the Mg2+ ions, and the interaction between the DNA bases. In an attempt to advance the understanding of catalytic control, we analyze the effect of the protein residues, taking human DNA polymerase beta as a model system. Specifically, we examine the ability of different theoretical models to reproduce the effect of ionized residues on the transition state (TS) binding energy and the corresponding k(pol)/KD. We also explore the role of the Mg2+ ions in the binding and catalysis processes. The application of the microscopic linear response approximation (LRA) and the semimacroscopic PDLD/S-LRA methods to a benchmark of mutational studies produces a semiquantitative correlation and indicates that these methods can provide predictive power. However, pre-steady-state and steady-state kinetic studies currently available do not give a unique benchmark, owing principally to widely varying experimental conditions. We believe that a more uniform experimental benchmark is needed for further refinement of the theoretical models. The analysis of the correlation between the results obtained by a rigorous thermodynamic cycle and by simpler approximations indicates that the protein reorganization between the open, i.e., unbound, form and the closed form does not change the magnitude of the calculated mutational effects in a major way for the experimental data used in this study. The use of the PDLD/S-LRA group contributions allows us to construct energy-based correlation diagrams that can help toward understanding the coupling, i.e., transfer of information, between the base-binding and catalytic sites and to gain a deeper insight into the molecular basis of DNA replication fidelity. Our analysis suggests that the allosteric matrix obtained by subtracting the correlation matrix of the correct and incorrect base pairs should prove useful in exploring the information transfer occurring between the base-binding and catalytic sites. This type of treatment should be especially effective when coupled with structural studies of polymerase-DNA-base mispair ternary complexes and studies using polymerase double mutants. We discuss the potential of direct calculations of binding energy of the TS in a rational design of TS analogues and in drug design.  相似文献   

3.
Kumar K  Walz FG 《Biochemistry》2001,40(12):3748-3757
Combinatorial random mutageneses involving either Asn43 with Asn44 (set 1) or Glu46 with an adjacent insertion (set 2) were undertaken to explore the functional perfection of the guanine recognition loop of ribonuclease T(1) (RNase T(1)). Four hundred unique recombinants were screened in each set for their ability to enhance enzyme catalysis of RNA cleavage. After a thorough selection procedure, only six variants were found that were either as active or more active than wild type which included substitutions of Asn43 by Gly, His, Leu, or Thr, an unplanned Tyr45Ser substitution and Glu46Pro with an adjacent Glu47 insertion. Asn43His-RNase T(1) has the same loop sequence as that for RNases Pb(1) and Fl(2). None of the most active mutants were single substitutions at Asn44 or double substitutions at Asn43 and Asn44. A total of 13 variants were purified, and these were subjected to kinetic analysis using RNA, GpC, and ApC as substrates. Modestly enhanced activities with GpC and RNA involved both k(cat) and K(M) effects. Mutants having low activity with GpC had proportionately even lower relative activity with RNA. Asn43Gly-RNase T(1) and all five of the purified mutants in set 2 exhibited similar values of k(cat)/K(M) for ApC which were the highest observed and about 10-fold that for wild type. The specificity ratio [(k(cat)/K(M))(GpC)/(k(cat)/K(M))(ApC)] varied over 30 000-fold including a 10-fold increase [Asn43His variant; mainly due to a low (k(cat)/K(M))(ApC)] and a 3000-fold decrease (Glu46Ser/(insert)Gly47 variant; mainly due to a low (k(cat)/K(M))(GpC)) as compared with wild type. It is interesting that k(cat) (GpC) for the Tyr45Ser variant was almost 4-fold greater than for wild type and that Pro46/(insert)Glu47 RNase T(1) is 70-fold more active than the permuted variant (insert)Pro47-RNase T(1) which has a conserved Glu46. In any event, the observation that only 6 out of 800 variants surveyed had wild-type activity supports the view that functional perfection of the guanine recognition loop of RNase T(1) has been achieved.  相似文献   

4.
A new thermostable dipeptidase gene was cloned from the thermophile Brevibacillus borstelensis BCS-1 by genetic complementation of the D-Glu auxotroph Escherichia coli WM335 on a plate containing D-Ala-D-Glu. Nucleotide sequence analysis revealed that the gene included an open reading frame coding for a 307-amino-acid sequence with an M(r) of 35,000. The deduced amino acid sequence of the dipeptidase exhibited 52% similarity with the dipeptidase from Listeria monocytogenes. The enzyme was purified to homogeneity from recombinant E. coli WM335 harboring the dipeptidase gene from B. borstelensis BCS-1. Investigation of the enantioselectivity (E) to the P(1) and P(1)' site of Ala-Ala revealed that the ratio of the specificity constant (k(cat)/K(m)) for L-enantioselectivity to the P(1) site of Ala-Ala was 23.4 +/- 2.2 [E = (k(cat)/K(m))(L,D)/(k(cat)/K(m))(D,D)], while the D-enantioselectivity to the P(1)' site of Ala-Ala was 16.4 +/- 0.5 [E = (k(cat)/K(m))(L,D)/(k(cat)/K(m))(L,L)] at 55 degrees C. The enzyme was stable up to 55 degrees C, and the optimal pH and temperature were 8.5 and 65 degrees C, respectively. The enzyme was able to hydrolyze L-Asp-D-Ala, L-Asp-D-AlaOMe, Z-D-Ala-D-AlaOBzl, and Z-L-Asp-D-AlaOBzl, yet it could not hydrolyze D-Ala-L-Asp, D-Ala-L-Ala, D-AlaNH(2), and L-AlaNH(2.) The enzyme also exhibited beta-lactamase activity similar to that of a human renal dipeptidase. The dipeptidase successfully synthesized the precursor of the dipeptide sweetener Z-L-Asp-D-AlaOBzl.  相似文献   

5.
Wang GP  Hansen MR  Grubmeyer C 《Biochemistry》2012,51(22):4406-4415
Residue-to-alanine mutations and a two-amino acid deletion have been made in the highly conserved catalytic loop (residues 100-109) of Salmonella typhimurium OMP synthase (orotate phosphoribosyltransferase, EC 2.4.2.10). As described previously, the K103A mutant enzyme exhibited a 10(4)-fold decrease in k(cat)/K(M) for PRPP; the K100A enzyme suffered a 50-fold decrease. Alanine mutations at His105 and Glu107 produced 40- and 7-fold decreases in k(cat)/K(M), respectively, and E101A, D104A, and G106A were slightly faster than the wild-type (WT) in terms of k(cat), with minor effects on k(cat)/K(M). Equilibrium binding of OMP or PRPP in binary complexes was affected little by loop mutation, suggesting that the energetics of ground-state binding have little contribution from the catalytic loop, or that a favorable binding energy is offset by costs of loop reorganization. Pre-steady-state kinetics for mutants showed that K103A and E107A had lost the burst of product formation in each direction that indicated rapid on-enzyme chemistry for WT, but that the burst was retained by H105A. Δ102Δ106, a loop-shortened enzyme with Ala102 and Gly106 deleted, showed a 10(4)-fold reduction of k(cat) but almost unaltered K(D) values for all four substrate molecules. The 20% (i.e., 1.20) intrinsic [1'-(3)H]OMP kinetic isotope effect (KIE) for WT is masked because of high forward and reverse commitment factors. K103A failed to express intrinsic KIEs fully (1.095 ± 0.013). In contrast, H105A, which has a smaller catalytic lesion, gave a [1'-(3)H]OMP KIE of 1.21 ± 0.0005, and E107A (1.179 ± 0.0049) also gave high values. These results are interpreted in the context of the X-ray structure of the complete substrate complex for the enzyme [Grubmeyer, C., Hansen, M. R., Fedorov, A. A., and Almo, S. C. (2012) Biochemistry 51 (preceding paper in this issue, DOI 10.1021/bi300083p )]. The full expression of KIEs by H105A and E107A may result from a less secure closure of the catalytic loop. The lower level of expression of the KIE by K103A suggests that in these mutant proteins the major barrier to catalysis is successful closure of the catalytic loop, which when closed, produces rapid and reversible catalysis.  相似文献   

6.
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   

7.
The six mutations, referred to as the Hex mutations, that together have been shown to convert Escherichia coli aspartate aminotransferase (AATase) specificity to be substantially like that of E. coli tyrosine aminotransferase (TATase) are dissected into two groups, (T109S/N297S) and (V39L/K41Y/T47I/N69L). The letters on the left and right of the numbers designate AATase and TATase residues, respectively. The T109S/N297S pair has been investigated previously. The latter group, the "Grease" set, is now placed in the AATase framework, and the retroGrease set (L39V/Y41K/I47T/L69N) is substituted into TATase. The Grease mutations in the AATase framework were found primarily to lower K(M)s for both aromatic and dicarboxylic substrates. In contrast, retroGrease TATase exhibits lowered k(cat)s for both substrates. The six retroHex mutations, combining retroGrease and S109T/S297N, were found to invert the substrate specificity of TATase, creating an enzyme with a nearly ninefold preference (k(cat)/K(M)) for aspartate over phenylalanine. The retroHex mutations perturb the electrostatic environment of the pyridoxal phosphate cofactor, as evidenced by a spectrophotometric titration of the internal aldimine, which uniquely shows two pK(a)s, 6.1 and 9.1. RetroHex was also found to have impaired dimer stability, with a K(D) for dimer dissociation of 350 nM compared with the wild type K(D) of 4 nM. Context dependence and additivity analyses demonstrate the importance of interactions of the Grease residues with the surrounding protein framework in both the AATase and TATase contexts, and with residues 109 and 297 in particular. Context dependence and cooperativity are particularly evident in the effects of mutations on k(cat)/K(M)(Asp). Effects on k(cat)/K(M)(Phe) are more nearly additive and context independent.  相似文献   

8.
Enzymes capable of benzoate ester hydrolysis have several potential medical and industrial applications. A variant of human carbonic anhydrase II (HCAII) was constructed, by rational design, that is capable of hydrolysing para-nitrophenyl benzoate (pNPBenzo) with an efficiency comparable to some naturally occurring esterases. The design was based on a previously developed strategy [G. H?st, L.G. M?rtensson, B.H. Jonsson, Redesign of human carbonic anhydrase II for increased esterase activity and specificity towards esters with long acyl chains, Biochim. Biophys. Acta 1764 (2006) 1601-1606.], in which docking of a transition state analogue (TSA) to the active site of HCAII was used to predict mutations that would allow the reaction. A triple mutant, V121A/V143A/T200A, was thus constructed and shown to hydrolyze pNPBenzo with k(cat)/K(M)=625 (+/- 38) M(-1) s(-1). It is highly active with other ester substrates as well, and hydrolyzes para-nitrophenyl acetate with k(cat)/K(M)=101,700 (+/- 4800) M(-1) s(-1), which is the highest esterase efficiency so far for any CA variant. A parent mutant (V121A/V143A) has measurable K(M) values for para-nitrophenyl butyrate (pNPB) and valerate (pNPV), but for V121A/V143A/T200A no K(M) could be determined, showing that the additional T200A mutation has caused a decreased substrate binding. However, k(cat)/K(M) is higher with both substrates for the triple mutant, indicating that binding energy has been diverted from substrate binding to transition state stabilization.  相似文献   

9.
Protease-substrate interactions are governed by a variety of structural features. Although the substrate sequence specificities of numerous proteases have been established, "topological specificities," whereby proteases may be classified based on recognition of distinct three-dimensional structural motifs, have not. The aggrecanase members of the ADAMTS (a disintegrin and metalloproteinase with thrombospondin motifs) family cleave a variety of proteins but do not seem to possess distinct sequence specificities. In the present study, the topological substrate specificity of ADAMTS-4 (aggrecanase-1) was examined using triple-helical or single-stranded poly(Pro) II helical peptides. Substrate topology modulated the affinity and sequence specificity of ADAMTS-4 with K(m) values indicating a preference for triple-helical structure. In turn, non-catalytic ADAMTS-4 domains were critical for hydrolysis of triple-helical and poly(Pro) II helical substrates. Comparison of ADAMTS-4 with MMP-1 (collagenase 1), MMP-13 (collagenase 3), trypsin, and thermolysin using triple-helical peptide (THP) and single-stranded peptide (SSP) substrates demonstrated that all five proteases possessed efficient "triple-helical peptidase" activity and fell into one of two categories: (k(cat)/K(m))(SSP) > (k(cat)/K(m))(THP) (thermolysin, trypsin, and MMP-13) or (k(cat)/K(m))(THP) > or = (k(cat)/K(m))(SSP) and (K(m))(SSP) > (K(m))(THP) (MMP-1 and ADAMTS-4). Overall these results suggest that topological specificity may be a guiding principle for protease behavior and can be utilized to design specific substrates and inhibitors. The triple-helical and single-stranded poly(Pro) II helical peptides represent the first synthetic substrates successfully designed for aggrecanases.  相似文献   

10.
Cellobiose dehydrogenase is an extracellular flavocytochrome, which catalyzes the oxidation of cellobiose and other soluble oligosaccharides to their respective lactones, while reducing various one- and two-electron acceptors. Two residues at the active site of the flavin domain, His689 and Asn732, have been proposed to play critical roles in the oxidation of the substrate. To test these proposals, each residue was substituted with either a Gln, Asn, Glu, Asp, Val, Ala, and/or a His residue by site-directed mutagenesis, using a homologous expression system previously developed in our laboratory. This enabled an examination of the functional, stereochemical, and electrostatic constraints for binding and oxidation of the substrate. The steady-state kinetic parameters for the variant proteins were compared using cellobiose and its epimer, lactose, as the substrates. The H689 variants all exhibit >1000-fold lower k(cat) values, while the K(m) values for both substrates in these variants are similar to that of the wild-type enzyme. This supports the proposed role of this His residue as a general base in catalysis. The N732 variants exhibit a range of kinetic parameters: the k(cat) values for oxidation are 5-4000-fold lower than that for the wild-type enzyme, while the K(m) values vary between similar to and 60-fold higher than that for the wild-type. The difference in binding energy between cellobiose and lactose was calculated using the relationship delta(delta G) = -RT ln[(k(cat)/K(m))(lactose)/(k(cat)/K(m))(cellobiose)]. This calculation for the wild-type enzyme suggests that lactose binds considerably more weakly than cellobiose (7.2 kJ/mol difference), which corresponds to one extra (cumulative) hydrogen bond for cellobiose over lactose. Mutations at Asn732 result in a further weakening of lactose binding over cellobiose (2-4 kJ/mol difference). The results support a role for Asn732 in the binding of the substrate.  相似文献   

11.
Meyer-Almes FJ  Auer M 《Biochemistry》2000,39(43):13261-13268
A new mathematical formalism is deduced which allows for the calculation of the k(cat) over K(M) ratio based on measurements of the enzyme kinetics with substrate concentrations much lower than K(M). The equations are also applied on the action of an inhibitor on enzyme activity yielding the binding constant, K(i), of an inhibitor molecule. For practical evaluation of the new theoretical approach, the enzymatic reaction of CD45 phosphatase was used as a well-characterized model system with known inhibitors for testing the K(i) value determination scheme. The k(cat)/K(M) ratio was calulated to be 4.7 x 10(5) M(-)(1) s(-)(1), the K(i) of the inhibitor molecule PKF52-524 was estimated to be (1-2) x 10(-)(7) M and the association rate of the inhibitor PKF52-524 to CD45 phosphatase was estimated to be 59 M(-)(1) s(-)(1).  相似文献   

12.
Friedreich's ataxia (FRDA) is a progressive neurodegenerative disease associated with the loss of function of the protein frataxin (FXN) that results from low FXN levels due to a GAA triplet repeat expansion or, occasionally, from missense mutations in the FXN gene. Here biochemical and structural properties of FXN variants, including three FRDA missense mutations (N146K, Q148R, and R165C) and three related mutants (N146A, Q148G, and Q153A), were determined in an effort to understand the structural basis for the loss of function. In vitro assays revealed that although the three FRDA missense mutations exhibited similar losses of cysteine desulfurase and Fe-S cluster assembly activities, the causes for these activation defects were distinct. The R165C variant exhibited a k(cat)/K(M) higher than that of native FXN but weak binding to the NFS1, ISD11, and ISCU2 (SDU) complex, whereas the Q148R variant exhibited the lowest k(cat)/K(M) of the six tested FXN variants and only a modest binding deficiency. The order of the FXN binding affinities for the SDU Fe-S assembly complex was as follows: FXN > Q148R > N146A > Q148G > N146K > Q153A > R165C. Four different classes of FXN variants were identified on the basis of their biochemical properties. Together, these structure-function studies reveal determinants for the binding and allosteric activation of the Fe-S assembly complex and provide insight into how FRDA missense mutations are functionally compromised.  相似文献   

13.
Y L Jiang  J T Stivers 《Biochemistry》2001,40(25):7710-7719
The DNA repair enzyme uracil DNA glycosylase (UDG) is a powerful N-glycohydrolase that cleaves the glycosidic bond of deoxyuridine in DNA. We have investigated the role of substrate binding energy in catalysis by systematically dismantling the optimal substrate Ap(+1)UpA(-1)pA(-2) by replacing the nucleotides at the +1, -1, or -2 position with a tetrahydrofuran abasic site nucleotide (D), a 3-hydroxypropyl phosphodiester spacer (S), a phosphate monoester (p), or a hydroxyl group (h). Contrary to previous reports, the minimal substrate for UDG is 2'-deoxyuridine (hUh). UDG has a significant catalytic efficiency (CE) for hUh of 4 x 10(7) M(-1) [CE = (k(cat)/K(m))(1/k(non)), where k(non) is the rate of the spontaneous hydrolysis reaction of hUh at 25 degrees C]. Addition of +1 and -1 phosphate monoanions to form pUp increases k(cat)/K(m) by 45-fold compared to that of hUh. The k(cat)/K(m) for pUp, but not pU or Up, is found to decrease by 20-fold over the pH range of 6-9 with a pK(a) of 7.1, which is identical to the pK(a) values for deprotonation of the +1 and -1 phosphate groups determined by the pH dependence of the (31)P NMR chemical shifts. This pH dependence indicates that binding of the pUp tetraanion is disfavored, possibly due to unfavorable desolvation or electrostatic properties of the highly charged +1 and -1 phosphate groups. Addition of flexible hydroxypropyl groups to the +1 and -1 positions to make SpUpS increases k(cat)/K(m) by more than 10(5)-fold compared to that of hUh, which is a 20-fold greater effect than observed with rigid D substituents in these positions (i.e., DpUpD). The -2 phosphoester or nucleotide is found to increase the reactivity of trimer substrates with rigid furanose rings or nucleotides in the +1 and -1 positions by 1300-270000-fold (i.e., DpUpD --> DpUpDpA or ApUpA --> ApUpApA). In contrast, the -2 nucleotide provides only an 8-fold rate enhancement when appended to the substrate containing the more flexible +1 and -1 S substituents (SpUpS --> SpUpSpA). These context-dependent effects of a -2 nucleotide are interpreted in terms of a mechanism in which the binding energy of this "handle" is used drive the rigid +1 and -1 A or D substituents into their binding pockets, resulting in a net catalytic benefit of -4.3 to -7.5 kcal/mol. Taken together, these results systematically track how UDG uses distant site binding interactions to produce an overall four billion-fold increase in CE compared to that of the minimal substrate hUh.  相似文献   

14.
The role of the Lys68*:Glu265 intersubunit salt bridge that is conserved (Csb) in all known aspartate aminotransferases (AATases), except those of animal cytosolic, Ac (His68*:Glu265), and plant mitochondrial, Pm (Met68*:Gln265), origins, was evaluated in the Escherichia coli AATase. Two double-mutant cycles, to K68M/E265Q and the charge reversed K68E/E265K, were characterized with the context dependence (C) and impact (I) formalism, previously defined for functional chimeric analysis. Mutations of Lys68* with Glu265 fixed are generally more deleterious than the converse mutations of Glu265 with Lys68* fixed, showing that buried negative charges have greater effects than buried positive charges in this context. Replacement of the charged Lys68*:Glu265 with the K68M/E265Q neutral pair introduces relatively small effects on the kinetic parameters. The differential sensitivity of k(cat)/K(M, L-Asp) and k(cat)/K(M, alpha-KG) to salt bridge mutagenic replacements is shown by a linear-free energy relationship, in which the logarithms of the latter second order rate constants are generally decreased by a factor of two more than are those of the former. Thus, k(cat)/K(M, L-Asp) and k(cat)/K(M, alpha-KG) are 133 and 442 mM(-1)s(-1) for the wild-type (WT) enzyme, respectively, but their relative order is reversed in the more severely compromised mutants (14.8 and 5.3 mM(-1)s(-1) for K68E). A Venn diagram illustrates apparent forced covariances of groups of amino acids that accompany the naturally occurring salt bridge replacements in the Pm and Ac classes. The more deeply rooted tree indicates that the Csb variant was the ancestral specie.  相似文献   

15.
A direct mass spectrometric approach was used for the determination of steady-state kinetic parameters, the turnover number (k(cat)), the Michaelis constant (K(M)), and the specificity constant (k(cat)/K(M)) for an enzyme-catalyzed hydrolysis of xylooligosaccharides. Electrospray ionization mass spectrometry was performed to observe product distributions and to determine k(cat), K(M), and k(cat)/K(M) values for Trichoderma reesei endo-1,4-beta-xylanase II (TRX II) with xylohexaose (Xyl(6)), xylopentaose (Xyl(5)), xylotetraose (Xyl(4)), and xylotriose (Xyl(3)) as substrates. The determined k(cat)/K(M) values (0.93, 0.37, 0.027, and 0.00015 microM(-1) s(-1), respectively) indicated that Xyl(6) was the most preferred substrate of TRX II. In addition, the obtained K(M) value for Xyl(5) (136 microM) was roughly twice as high as that for Xyl(6) (73 microM), suggesting that at least six putative subsites contribute to the substrate binding in the active site of TRX II. Previous mass spectrometric assays for enzyme kinetics have been used mostly in the case of reactions that result in a transfer of acidic groups (e.g., phosphate) into neutral oligosaccharides giving rise to negatively charged products. Here we demonstrate that such analysis is also feasible in the case of neutral underivatized oligosaccharides. Implications of the results for the catalytic mechanism of TRX II in particular are discussed.  相似文献   

16.
Salts inhibit the activity of sweet almond beta-glucosidase. For cations (Cl(-) salts) the effectiveness follows the series: Cu(+2), Fe(+2)>Zn(+2)>Li(+)>Ca(+2)>Mg(+2)>Cs(+)>NH(4)(+)>Rb(+)>K(+)>Na(+) and for anions (Na(+) salts) the series is: I(-)>ClO(4)(-)>(-)SCN>Br(-) approximately NO(3)(-)>Cl(-) approximately (-)OAc>F(-) approximately SO(4)(-2). The activity of the enzyme, like that of most glycohydrolases, depends on a deprotonated carboxylate (nucleophile) and a protonated carboxylic acid for optimal activity. The resulting pH-profile of k(cat)/K(m) for the beta-glucosidase-catalyzed hydrolysis of p-nitrophenyl glucoside is characterized by a width at half height that is strongly sensitive to the nature and concentration of the salt. Most of the inhibition is due to a shift in the enzymic pK(a)s and not to an effect on the pH-independent second-order rate constant, (k(cat)/K(m))(lim). For example, as the NaCl concentration is increased from 0.01 M to 1.0 M the apparent pK(a1)increases (from 3.7 to 4.9) and the apparent pK(a2)decreases (from 7.2 to 5.9). With p-nitrophenyl glucoside, the value of the pH-independent (k(cat)/K(m))(lim) (=9 x 10(4) M(-1) s(-1)) is reduced by less than 4% as the NaCl concentration is increased. There is a similar shift in the pK(a)s when the LiCl concentration is increased to 1.0 M. The results of these salt-induced pK(a) shifts rule out a significant contribution of reverse protonation to the catalytic efficiency of the enzyme. At low salt concentration, the fraction of the catalytically active monoprotonated enzyme in the reverse protonated form (i.e., proton on the group with a pK(a) of 3.7 and dissociated from the group with a pK(a) of 7.2) is very small ( approximately 0.03%). At higher salt concentrations, where the two pK(a)s become closer, the fraction of the monoprotonated enzyme in the reverse protonated form increases over 300-fold. However, there is no increase in the intrinsic reactivity, (k(cat)/K(m))(lim), of the monoprotonated species. For other enzymes which may show such salt-induced pK(a) shifts, this provides a convenient test for the role of reverse protonation.  相似文献   

17.
Schultz BE  Misialek S  Wu J  Tang J  Conn MT  Tahilramani R  Wong L 《Biochemistry》2004,43(34):11083-11091
Histone deacetylase (HDAC) enzymes modulate gene expression through the deacetylation of acetylated lysine residues on histone proteins. They operate in biological systems as part of multiprotein corepressor complexes. To understand the reactivity of isolated HDACs and the contribution of cofactor binding to reactivity, the reaction kinetics of isolated, recombinant human HDACs 1, 2, 3, 6, 8, and 10 were measured using a novel, continuous protease-coupled enzyme assay. Values of k(cat) and k(cat)/K(m) and the pH dependence of these values were determined for the reactions of each isozyme with acetyl-Gly-Ala-(N(epsilon)-acetyl-Lys)-AMC. Values of k(cat) spanned the range of 0.006-2.8 s(-1), and k(cat)/K(m) values ranged from 60 to 110000 M(-1) s(-1). The pH profiles for both k(cat) and k(cat)/K(m) were bell-shaped for all of the HDAC isozymes, with pH optima at approximately pH 8. Values of K(i) for the inhibitor trichostatin A were determined for each isozyme. The inhibition constants were generally similar for all HDAC isozymes, except that the value for HDAC8 was significantly higher than that for the other isozymes. The reaction of HDAC8 with an alternative substrate was performed to assess the steric requirements of the HDAC8 active site, and the effect of phosphorylation on HDAC1 activity was examined. The results are discussed in terms of the biological roles of the HDAC enzymes and the proposed reaction mechanism of acetyllysine hydrolysis by these enzymes.  相似文献   

18.
Wolthers KR  Schimerlik MI 《Biochemistry》2001,40(15):4722-4737
Binding of Ca(2+)-activated calmodulin (Ca(2+)-CaM) to neuronal nitric-oxide synthase (nNOS) increases the rate of 2,6-dichloroindolphenol (DCIP) reduction 2-3-fold and that of cytochrome c(3+) 10-20-fold. Parallel initial velocity patterns indicated that both substrates were reduced via two-half reactions in a ping-pong mechanism. Product and dead-end inhibition data with DCIP were consistent with an iso ping-pong bi-bi mechanism; however, product and dead-end inhibition studies with cytochrome c(3+) were consistent with the (two-site) ping-pong mechanism previously described for the NADPH-cytochrome P450 reductase-catalyzed reduction of cytochrome c(3+) [Sem, D., and Kasper, C. (1994) Biochemistry 33, 12012--12021]. Dead-end inhibition by 2'-adenosine monophosphate (2'AMP) was competitive versus NADPH for both electron acceptors, although the value of the slope inhibition constant, K(is), was 25-30-fold greater with DCIP as the substrate than with cytochrome c(3+). The difference in the apparent affinity of 2'AMP is proposed to result from a rapidly equilibrating isomerization step that occurs in both mechanisms prior to the binding of NADPH. Thus, initial velocity, product, and dead-end inhibition data were consistent with a di-iso ping-pong bi-bi and an iso (two-site) ping-pong mechanism for the reduction of DCIP and cytochrome c(3+), respectively. The presence Ca(2+)-CaM did not alter the proposed kinetic mechanisms. The activated cofactor had a negligible effect on (k(cat)/K(m))(NADPH), while it increased (k(cat)/K(m))(DCIP) and (k(cat)/K(m))(cytc) 4.5- and 23-fold, respectively.  相似文献   

19.
Isopentenyl phosphate kinase (IPK) catalyzes the ATP-dependent phosphorylation of isopentenyl phosphate (IP) to form isopentenyl diphosphate (IPP) during biosynthesis of isoprenoid metabolites in Archaea. The structure of IPK from the archeaon Thermoplasma acidophilum (THA) was recently reported and guided the reconstruction of the IP binding site to accommodate the longer chain isoprenoid monophosphates geranyl phosphate (GP) and farnesyl phosphate (FP). We created four mutants of THA IPK with different combinations of alanine substitutions for Tyr70, Val73, Val130, and Ile140, amino acids with bulky side chains that limited the size of the side chain of the isoprenoid phosphate substrate that could be accommodated in the active site. The mutants had substantially increased GP kinase activity, with 20-200-fold increases in k(cat)(GP) and 30-130-fold increases in k(cat)(GP)/K(M)(GP) relative to those of wild-type THA IPK. The mutations also resulted in a 10(6)-fold decrease in k(cat)(IP)/K(M)(IP) compared to that of wild-type IPK. No significant change in the kinetic parameters for the cosubstrate ATP was observed, signifying that binding between the nucleotide binding site and the IP binding site was not cooperative. The shift in substrate selectivity from IP to GP, and to a lesser extent, FP, in the mutants could act as a starting point for the creation of more efficient GP or FP kinases whose products could be exploited for the chemoenzymatic synthesis of radiolabeled isoprenoid diphosphates.  相似文献   

20.
Neurolysin (EP24.16) and thimet oligopeptidase (EP24.15) are closely related metalloendopeptidases. Site-directed mutagenesis of Tyr(613) (EP24.16) or Tyr(612) (EP24.15) to either Phe or Ala promoted a strong reduction of k(cat)/K(M) for both enzymes. These data suggest the importance of both hydroxyl group and aromatic ring at this specific position during substrate hydrolysis by these peptidases. Furthermore, the EP24.15 A607G mutant showed a k(cat)/K(M) of 2x10(5) M(-1) s(-1) for the Abz-GFSIFRQ-EDDnp substrate, similar to that of EP24.16 (k(cat)/K(M)=3x10(5) M(-1) s(-1)) which contains Gly at the corresponding position; the wild type EP24.15 has a k(cat)/K(M) of 2.5x10(4) M(-1) s(-1) for this substrate.  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