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 62 毫秒
1.
Stratton JR  Pelton JG  Kirsch JF 《Biochemistry》2001,40(35):10411-10416
The low-barrier hydrogen bond (LBHB) between the Asp and His residues of the catalytic triad in a serine protease was perturbed via the D32C mutation in subtilisin BPN' (Bacillus protease N'). This mutant enzyme catalyzes the hydrolysis of N-Suc-Ala-Ala-Pro-Phe-SBzl with a k(cat)/K(m) value that is only 8-fold reduced from that of the wild-type (WT) enzyme. The value of k(cat)/K(m) for the corresponding p-nitroanilide (pNA) substrate is only 50-fold lower than that of the WT enzyme (DeltaDeltaG++ = 2.2 kcal/mol). The pK(a) controlling the ascending limb of the pH versus k(cat)/K(m) profile is lowered from 7.01 (WT) to 6.53 (D32C), implying that any hydrogen bond replacing that between Asp32 and His64 of the WT enzyme most likely involves the neutral thiol rather than the thiolate form of Cys32. It is shown by viscosity variation that the reaction of WT subtilisin with N-Suc-Ala-Ala-Pro-Phe-SBzl is 50% (sucrose) to 100% (glycerol) diffusion-controlled, while that of the D32C construct is 29% (sucrose) to 76% (glycerol) diffusion-controlled. The low-field NMR resonance of 18 ppm that has been assigned to a proton shared by Asp32 and His64, and is considered diagnostic of a LBHB in the WT enzyme, is not present in D32C subtilisin. Thus, the LBHB is not an inherent requirement for substantial rate enhancement for subtilisin.  相似文献   

2.
Thiols represent preferential targets of peroxynitrite in biological systems. In this work, we investigated the mechanisms and kinetics of the reaction of peroxynitrite with the dithiol dihydrolipoic acid (DHLA) and its oxidized form, lipoic acid (LA). Peroxynitrite reacted with DHLA being oxidation yields higher at alkaline pH. The stoichiometry for the reaction was two thiols oxidized per peroxynitrite. LA formation accounted for approximately 50% DHLA consumption at pH 7.4, probably reflecting secondary reactions between LA and peroxynitrite. Indeed, peroxynitrous acid reacted with LA with an apparent second-order rate constant (k(2app)) of 1400 M(-1) s(-1) at pH 7.4 and 37 degrees C. Nitrite and LA-thiosufinate were formed as reaction products. Surprisingly, the k(2app) for peroxynitrite-dependent DHLA oxidation was only 250 M(-1) s(-1) per thiol, at pH 7.4 and 37 degrees C. Testing various low-molecular-weight thiols, we found that an increase in the thiol pK (pK(SH)) value correlated with a decrease of k(2app) for the reaction with peroxynitrite at pH 7.4. The pK(SH) for DHLA is 10.7, in agreement with its modest reactivity with peroxynitrite.  相似文献   

3.
Flagg SC  Giri N  Pektas S  Maroney MJ  Knapp MJ 《Biochemistry》2012,51(33):6654-6666
Prolyl hydroxylase domain 2 (PHD2) is deemed a primary oxygen sensor in humans, yet many details of its underlying mechanism are still not fully understood. (Fe(2+) + αKG)PHD2 is 6-coordinate, with a 2His/1Asp facial triad occupying three coordination sites, a bidentate α-ketoglutarate occupying two sites, and an aquo ligand in the final site. Turnover is thought to be initiated upon release of the aquo ligand, creating a site for O(2) to bind at the iron. Herein we show that steady-state turnover is faster under acidic conditions, with k(cat) exhibiting a kinetic pK(a) = 7.22. A variety of spectroscopic probes were employed to identify the active-site acid, through comparison of (Fe(2+) + αKG)PHD2 at pH 6.50 with pH 8.50. The near-UV circular dichroism spectrum was virtually unchanged at elevated pH, indicating that the secondary structure did not change as a function of pH. UV-visible and Fe X-ray absorption spectroscopy indicated that the primary coordination sphere of Fe(2+) changed upon increasing the pH; extended X-ray absorption fine structure analysis found a short Fe-(O/N) bond length of 1.96 ? at pH 8.50, strongly suggesting that the aquo ligand was deprotonated at this pH. Solvent isotope effects were measured during steady-sate turnover over a wide pH-range, with an inverse solvent isotope effect (SIE) of k(cat) observed ((D(2)O)k(cat) = 0.91 ± 0.03) for the acid form; a similar SIE was observed for the basic form of the enzyme ((D(2)O)k(cat) = 0.9 ± 0.1), with an acid equilibrium offset of ΔpK(a) = 0.67 ± 0.04. The inverse SIE indicated that aquo release from the active site Fe(2+) immediately precedes a rate-limiting step, suggesting that turnover in this enzyme may be partially limited by the rate of O(2) binding or activation, and suggesting that aquo release is relatively slow. The unusual kinetic pK(a) further suggested that PHD2 might function physiologically to sense both intracellular pO(2) as well as pH, which could provide for feedback between anaerobic metabolism and hypoxia sensing.  相似文献   

4.
Wu SP  Wu G  Surerus KK  Cowan JA 《Biochemistry》2002,41(28):8876-8885
ISU-type proteins mediate cluster transfer to apo protein targets. Rate constants have been determined for cluster transfer from ISU to apo Fd for both Homo sapiens and Schizosaccharomyces pombe proteins, and cross reactions have also been examined. Substitution of a key aspartate residue of ISU is found to decrease the rate of cluster transfer by at least an order of magnitude (for wild-type Hs ISU cluster transfer to Hs apo Fd, k(2) approximately 540 M(-1) min(-1), relative 56 M(-1) min(-1) for D37A ISU). This change in rate constant does not reflect any change in binding affinity of the ISU and Fd proteins. The pH dependencies of cluster transfer rates are similar for WT and D37A ISU, arguing against a role for Asp37 as a catalytic base, although evidence for general base catalysis mediating deprotonation of Cys from the apo target is supported by an observed pK(a) of 6.9 determined from the pH profiles for both WT and D37A ISU. Such a pK(a) value is at the lower limit for Cys and is common for solvent-accessible Cys thiols. The temperature dependence of the rate constant defining the cluster transfer reaction for wild type versus the aspartate derivative is distinct. Thermal activation parameters (DeltaH and DeltaS) are consistent with a solvent-accessible ISU-bound cluster, with desolvation as a principle barrier to cluster transfer. Experiments to determine the dependence of reaction rate constants on viscosity indicate cluster transfer to be rate-limiting. Fully oxidized cluster appears to be the natural state for transfer to target proteins. Reduced Fd does not readily reduce ISU-bound [2Fe-2S](2+) and does not promote cluster transfer to an apo Fd target.  相似文献   

5.
A soil metagenomic library was constructed and screened for clones that conferred fosfomycin resistance. A novel protein with 46 % identity to UDP-N-acetylglucosamine enolpyruvyl transferase (MurA) from Desulfuromonas acetoxidans DSM 684 (GenBank accession number: ZP_01311756) was identified. Multiple sequence alignment revealed that the novel protein was a natural MurA, in which an aspartic acid instead of a cysteine was located in the active site. An Asp120Cys mutant of Escherichia coli was constructed from the subclone through site-specific mutagenesis, and minimum inhibitory concentration of fosfomycin for the resistant subclone and its mutant were determined. These results showed that fosfomycin resistance was a result of the aspartic acid in the active site. Analysis of all existing MurA sequences revealed that MurAs with an active site aspartic acid that can confer fosfomycin resistance occur in ~14 % of bacteria.  相似文献   

6.
Escherichia coli uracil DNA glycosylase (UDG) catalyzes the hydrolysis of premutagenic uracil bases in DNA by flipping the deoxyuridine from the DNA helix [Stivers, J. T., et al. (1999) Biochemistry 38, 952]. A general acid-base mechanism has been proposed whereby His187 facilitates leaving group departure by protonating the O2 of uracil and Asp64 activates a water molecule for nucleophilic attack at C1' of the deoxyribose. Detailed kinetic studies on the H187Q, H187A, and D64N mutant enzymes indicate that Asp64 and His187 stabilize the chemical transition state by 5.3 and 4.8 kcal/mol, respectively, with little effect on substrate or product binding. The pH dependence of k(cat) for wild-type and H187Q UDG indicates that an unprotonated group in the enzyme-substrate complex (pK(a) = 6.2 +/- 0.2) is required for catalysis. This unprotonated group has a small DeltaH of ionization (-0.4 +/- 1.7 kcal/mol) and is absent in the pH profile for D64N UDG, suggesting that it corresponds to the general base Asp64. The pH dependence of k(cat) for wild-type, H187Q, and D64N UDG shows no evidence for an essential protonated group over the pH range of 5.5-10. Hence, the pK(a) of His187 must be outside this pH range if it serves as an electrophilic catalyst. These results support a mechanism in which Asp64 serves as the general base and His187 acts as a neutral electrophile, stabilizing a developing negative charge on uracil O2 in the transition state. In the following paper of this issue we establish by crystallography and heteronuclear NMR spectroscopy that the imidazole of His187 is neutral during the catalytic cycle of UDG.  相似文献   

7.
Human fibroblast activation protein (FAP), an integral membrane serine protease, was produced in insect cells as a hexa-His-tagged protein using a recombinant baculovirus expression system. Two isoforms of FAP, glycosylated and nonglycosylated, were identified by Western blotting using an anti-His-tag antibody and separated by lectin chromatography. The glycosylated FAP was purified to near homogeneity using immobilized metal affinity chromatography and was shown to have both postprolyl dipeptidyl peptidase and postgelatinase activities. In contrast, the nonglycosylated isoform demonstrated no detectable gelatinase activity by either zymography or a fluorescence-based gelatinase activity assay. The kinetic parameters of the dipeptidyl peptidase activity for glycosylated FAP were determined using dipeptide Ala-Pro-7-amino-trifluoromethyl-coumarin as the substrate. The k(cat) is 2.0 s(-1) and k(cat)/K(m) is 1.0 x 10(4) M(-1) s(-1) at pH 8.5. The pH dependence of k(cat) reveals two ionization groups with pK(a1) of 7.0 and pK(a2) of 11.0. The pH profile of k(cat)/K(m) yields similar results with pK(a1) 6.2 and pK(a2) 11.0. The neutral pK(a1) is associated with His at the active site. The basic pK(a2) might be contributed from an ionization group that is not involved directly in catalysis, instead associated with the stability of the active site structure.  相似文献   

8.
Crane EJ  Yeh JI  Luba J  Claiborne A 《Biochemistry》2000,39(34):10353-10364
The crystal structure of the flavoprotein NADH peroxidase shows that the Arg303 side chain forms a hydrogen bond with the active-site His10 imidazole and is therefore likely to influence the catalytic mechanism. Dithionite titration of an R303M mutant [E(FAD, Cys42-sulfenic acid)] yields a two-electron reduced intermediate (EH(2)) with enhanced flavin fluorescence and almost no charge-transfer absorbance at pH 7.0; the pK(a) for the nascent Cys42-SH is increased by over 3.5 units in comparison with the wild-type EH(2) pK(a) of Cys42-SOH. The crystal structure of the R303M peroxidase has been refined at 2.45 A resolution. In addition to eliminating the Arg303 interactions with His10 and Glu14, the mutant exhibits a significant change in the conformation of the Cys42-SOH side chain relative to FAD and His10 in particular. These and other results provide a detailed understanding of Arg303 and its role in the structure and mechanism of this unique flavoprotein peroxidase.  相似文献   

9.
PhnP is a phosphodiesterase that plays an important role within the bacterial carbon-phosphorus lyase (CP-lyase) pathway by recycling a "dead-end" intermediate, 5-phospho-α-d-ribosyl 1,2-cyclic phosphate, that is formed during organophosphonate catabolism. As a member of the metallo-β-lactamase superfamily, PhnP is most homologous in sequence and structure to tRNase Z phosphodiesterases. X-ray structural analysis of PhnP complexed with orthovanadate to 1.5 ? resolution revealed this inhibitor bound in a tetrahedral geometry by the two catalytic manganese ions and the putative general acid residue H200. Guided by this structure, we probed the contributions of first- and second-sphere active site residues to catalysis and metal ion binding by site-directed mutagenesis, kinetic analysis, and ICP-MS. Alteration of H200 to alanine resulted in a 6-33-fold decrease in k(cat)/K(M) with substituted methyl phenylphosphate diesters with leaving group pK(a) values ranging from 4 to 8.4. With bis(p-nitrophenyl)phosphate as a substrate, there was a 10-fold decrease in k(cat)/K(M), primarily the result of a large increase in K(M). Moreover, the nickel ion-activated H200A PhnP displayed a bell-shaped pH dependence for k(cat)/K(M) with pK(a) values (pK(a1) = 6.3; pK(a2) = 7.8) that were comparable to those of the wild-type enzyme (pK(a1) = 6.5; pK(a2) = 7.8). Such modest effects are counter to what is expected for a general acid catalyst and suggest an alternate role for H200 in this enzyme. A Br?nsted analysis of the PhnP reaction with a series of substituted phenyl methyl phosphate esters yielded a linear correlation, a β(lg) of -1.06 ± 0.1, and a Leffler α value of 0.61, consistent with a synchronous transition state for phosphoryl transfer. On the basis of these data, we propose a mechanism for PhnP.  相似文献   

10.
Shen B  English AM 《Biochemistry》2005,44(42):14030-14044
Although biologically active, nitroxyl (HNO) remains one of the most poorly studied NO(x). Protein-based thiols are suspected targets of HNO, forming either a disulfide or sulfinamide (RSONH2) through an N-hydroxysulfenamide (RSNHOH) addition product. Electrospray ionization mass spectrometry (ESI-MS) is used here to examine the products formed during incubation of thiol proteins with the HNO donor, Angeli's salt (AS; Na2N2O3). Only the disulfide, cystine, was formed in incubates of 15 mM free Cys with equimolar AS at pH 7.0-7.4. In contrast, the thiol proteins (120-180 microM), human calbindin D(28k) (HCalB), glyceraldehyde-3-phosphate dehydrogenase (GAPDH), and bovine serum albumin (BSA) gave four distinct types of derivatives in incubates containing 0.9-2.5 mM AS. Ions at M + n x 31 units were detected in the ESI mass spectra of intact HCalB (n = 1-5) and GAPDH (n = 2), indicating conversion of thiol groups on these proteins to RSONH2 (+31 units). An ion at M + 14 dominated the mass spectrum of BSA, and intramolecular sulfinamide cross-linking of Cys34 to one of its neighboring Lys or Arg residues would account for this mass increase. Low abundant M + 14 adducts were observed for HCalB, which additionally formed mixed disulfides when free Cys was present in the AS incubates. Cys149 and Cys153 formed an intramolecular disulfide in the AS/GAPDH incubates. Since AS also produces nitrite above pH 5 (HN2O3(-) --> HNO + NO2(-)), incubation with NaNO2 served to confirm that protein modification was HNO-mediated, and prior blocking with the thiol-specific reagent, N-ethylmaleimide, demonstrated that thiols are the targets of HNO. The results provide the first systematic characterization of HNO-mediated derivatization of protein thiols.  相似文献   

11.
JM Robbins  HR Ellis 《Biochemistry》2012,51(32):6378-6387
The alkanesulfonate monooxygenase enzyme (SsuD) catalyzes the oxygenolytic cleavage of a carbon-sulfur bond from sulfonated substrates. A mechanism involving acid-base catalysis has been proposed for the desulfonation mechanism by SsuD. In the proposed mechanism, base catalysis is involved in abstracting a proton from the alkane peroxyflavin intermediate, while acid catalysis is needed for the protonation of the FMNO(-) intermediate. The pH profiles of k(cat) indicate that catalysis by SsuD requires a group with a pK(a) of 6.6 ± 0.2 to be deprotonated and a second group with a pK(a) of 9.5 ± 0.1 to be protonated. The upper pK(a) value was not present in the pH profiles of k(cat)/K(m). Several conserved amino acid residues (His228, His11, His333, Cys54, and Arg226) have been identified as having potential catalytic importance due to the similar spatial arrangements with close structural and functional relatives of SsuD. Substitutions to these amino acid residues were generated, and the pH dependencies were evaluated and compared to wild-type SsuD. Although a histidine residue was previously proposed to be the active site base, the His variants possessed similar steady-state kinetic parameters as wild-type SsuD. Interestingly, R226A and R226K SsuD variants possessed undetectable activity, and there was no detectable formation of the C4a-(hydro)peroxyflavin intermediate for the Arg226 SsuD variants. Guanidinium rescue with the R226A SsuD variant resulted in the recovery of 1.5% of the wild-type SsuD k(cat) value. These results implicate Arg226 playing a critical role in catalysis and provide essential insights into the mechanistic steps that guide the SsuD desulfonation process.  相似文献   

12.
The periplasmic C-terminal domain of the Escherichia coli DsbD protein (cDsbD) has a thioredoxin fold. The two cysteine residues in the CXXC motif serve as the reductant for the disulfide bond of the N-terminal domain which can in turn act as a reductant for various periplasmic partners. The resulting disulfide bond in cDsbD is reduced via an unknown mechanism by the transmembrane helical domain of the protein. We show by NMR analysis of (13)C, (15)N-labelled cDsbD that the protein is rigid, is stable to extremes of pH and undergoes only localized conformational changes in the vicinity of the CXXC motif, and in adjacent regions of secondary structure, upon undergoing the reduced/oxidized transition. pK(a) values have been determined, using 2D NMR, for the N-terminal cysteine of the CXXC motif, Cys461, as well as for other active-site residues. It is demonstrated using site-directed mutagenesis that the negative charges of the side-chains of Asp455 and Glu468 in the active site contribute to the unusually high pK(a) value, 10.5, of Cys461. This value is higher than expected from knowledge of the reduction potential of cDsbD. In a double mutant of cDsbD, D455N/E468Q, the pK(a) value of Cys461 is lowered to 8.6, a value close to that expected for an unperturbed cysteine residue. The pK(a) value of the second cysteine in wild-type cDsbD, Cys464, is significantly higher than the maximum pH value that was studied (pH 12.2).  相似文献   

13.
IMP dehydrogenase (IMPDH) catalyzes the oxidation of IMP to XMP with the concomitant reduction of NAD(+). This reaction involves the formation of a covalent adduct with an active site Cys. This intermediate, E-XMP, hydrolyzes to produce XMP. The mutation of Asp338 to Ala severely impairs the activity of Escherichia coli IMPDH, decreasing the value of k(cat) by 650-fold. No (D)V(m) or (D)V/K(m) isotope effects are observed when 2-(2)H-IMP is the substrate for wild-type IMPDH. Values of (D)V(m) = 2.6 and (D)V/K(m) (IMP) = 3.4 are observed for Asp338Ala. Moreover, while a burst of NADH production is observed for wild-type IMPDH, no burst is observed for Asp338Ala. These observations indicate that the mutation has decreased the rate of hydride transfer by at least 5 x 10(3)-fold. In contrast, k(cat) for the hydrolysis of 2-chloroinosine-5'-monophosphate is decreased by only 8-fold. In addition, the rate constant for inactivation by 6-chloropurine riboside 5'-monophosphate is increased by 3-fold. These observations suggest that the mutation has little effect on the nucleophilicity of the active site Cys residue. These results are consistent with a recent crystal structure that shows a hydrogen bonding network between Asp338, the 2'-OH of IMP, and the amide group of NAD(+) [Colby, T. D., Vanderveen, K., Strickler, M. D., Markham, G. D., and Goldstein, B. M. (1999) Proc. Natl. Acad. Sci. U.S.A. 96, 3531-3536].  相似文献   

14.
The Staphylococcus aureus transpeptidase SrtA catalyzes the covalent attachment of LPXTG-containing virulence and colonization-associated proteins to cell-wall peptidoglycan in Gram-positive bacteria. Recent structural characterizations of staphylococcal SrtA, and related transpeptidases SrtB from S. aureus and Bacillus anthracis, provide many details regarding the active site environment, yet raise questions with regard to the nature of catalysis and active site cysteine thiol activation. Here we re-evaluate the kinetic mechanism of SrtA and shed light on aspects of its catalytic mechanism. Using steady-state, pre-steady-state, bisubstrate kinetic studies, and high-resolution electrospray mass spectrometry, revised steady-state kinetic parameters and a ping-pong hydrolytic shunt kinetic mechanism were determined for recombinant SrtA. The pH dependencies of kinetic parameters k(cat)/K(m) and k(cat) for the substrate Abz-LPETG-Dap(Dnp)-NH(2) were bell-shaped with pK(a) values of 6.3 +/- 0.2 and 9.4 +/- 0.2 for k(cat) and 6.2 +/- 0.2 and 9.4 +/- 0.2 for k(cat)/K(m). Solvent isotope effect (SIE) measurements revealed inverse behavior, with a (D)2(O)k(cat) of 0.89 +/- 0.01 and a (D)2(O)(k(cat)/K(m)) of 0.57 +/- 0.03 reflecting an equilibrium SIE. In addition, SIE measurements strongly implicated Cys184 participation in the isotope-sensitive rate-determining chemical step when considered in conjunction with an inverse linear proton inventory for k(cat). Last, the pH dependence of SrtA inactivation by iodoacetamide revealed a single ionization for inactivation. These studies collectively provide compelling evidence for a reverse protonation mechanism where a small fraction (ca. 0.06%) of SrtA is competent for catalysis at physiological pH, yet is highly active with an estimated k(cat)/K(m) of >10(5) M(-)(1) s(-)(1).  相似文献   

15.
16.
The two most buried carboxyl groups in ribonuclease Sa (RNase Sa) are Asp33 (99% buried; pK 2.4) and Asp79 (85% buried; pK 7.4). Above these pK values, the stability of the D33A variant is 6kcal/mol less than wild-type RNase Sa, and the stability of the D79A variant is 3.3kcal/mol greater than wild-type RNase Sa. The key structural difference between the carboxyl groups is that Asp33 forms three intramolecular hydrogen bonds, and Asp79 forms no intramolecular hydrogen bond. Here, we focus on Asp79 and describe studies of 11 Asp79 variants. Most of the variants were at least 2kcal/mol more stable than wild-type RNase Sa, and the most interesting was D79F. At pH 3, below the pK of Asp79, RNase Sa is 0.3kcal/mol more stable than the D79F variant. At pH 8.5, above the pK of Asp79, RNase Sa is 3.7kcal/mol less stable than the D79F variant. The unfavorable contribution of Asp79 to the stability appears to result from the Born self-energy of burying the charge and, more importantly, from unfavorable charge-charge interactions. To counteract the effect of the negative charge on Asp79, we prepared the Q94K variant and the crystal structure showed that the amino group of the Lys formed a hydrogen-bonded ion pair (distance, 2.71A; angle, 100 degrees ) with the carboxyl group of Asp79. The stability of the Q94K variant was about the same as the wild-type at pH 3, where Asp79 is uncharged, but 1kcal/mol greater than that of wild-type RNase Sa at pH 8.5, where Asp79 is charged. Differences in hydrophobicity, steric strain, Born self-energy, and electrostatic interactions all appear to contribute to the range of stabilities observed in the variants. When it is possible, replacing buried, non-hydrogen bonded, ionizable side-chains with non-polar side-chains is an excellent means of increasing protein stability.  相似文献   

17.
VanXY(C), a bifunctional enzyme from VanC-phenotype Enterococcus gallinarum BM4174 that catalyses D,D-peptidase and D,D-carboxypeptidase activities, was purified as the native protein, as a maltose-binding protein fusion and with an N-terminal tag containing six histidine residues. The kinetic parameters of His(6)-VanXY(C) were measured for a variety of precursors of peptidoglycan synthesis involved in resistance: for D-Ala-D-Ala, the K(m) was 3.6 mm and k(cat), 2.5 s(-1); for UDP-MurNAc-L-Ala-D-Glu-L-Lys-DAla-D-Ala (UDP-MurNAc-pentapeptide[Ala]), K(m) was 18.8 mm and k(cat) 6.2 s(-1); for D-Ala-D-Ser, K(m) was 15.5 mm and k(cat) 0.35 s(-1). His(6)-VanXYC was inactive against the peptidoglycan precursor UDP-MurNAc-L-Ala-D-Glu-L-Lys-D-Ala-D-Ser (UDP-MurNAc-pentapeptide[Ser]). The rate of hydrolysis of the terminal D-Ala of UDP-MurNAc-pentapeptide[Ala] was inhibited 30% by 2 mm D-Ala-D-Ser or UDP-MurNAc-pentapeptide[Ser]. Therefore preferential hydrolysis of substrates terminating in D-Ala would occur during peptidoglycan synthesis in E. gallinarum BM4174, leaving precursors ending in D-Ser with a lower affinity for glycopeptides to be incorporated into peptidoglycan.Mutation of an aspartate residue (Asp59) of His-tagged VanXY(C) corresponding to Asp68 in VanX to Ser or Ala, resulted in a 50% increase and 73% decrease, respectively, of the specificity constant (k(cat)/K(m)) for D-Ala-D-Ala. This situation is in contrast to VanX in which mutation of Asp68-->Ala produced a greater than 200,000-fold decrease in the substrate specificity constant. This suggests that Asp59, unlike Asp68 in VanX, does not have a pivotal role in catalysis.  相似文献   

18.
Hardy LW  Kirsch JF 《Biochemistry》1984,23(6):1282-1287
The solvent kinetic isotope effects (SKIE's) on k(cat) (D(V)) and on k(cat/Km[D(V/K)] were determined for the Bacillus cereus beta-lactamase I catalyzed hydrolysis of five substrates that have values of k(cat)/K(m) varying over the range (0.014-46.3) X 10(6)M(-1) s(-1) and of k(cat) between 0.5 and 2019 s(-1). The variation of D(V/K) was only from 1.06 to 1.25 among these compounds and that in D(V) was from 1.50 to 2.16. These results require that Dk(1), the SKIE on the enzyme-substrate association rate constant, and D(k-1/k2), that on the partition ratio of the ES complex, both be near 1. The larger SKIE observed on D(V) requires that an exchangeable proton be in flight for either or both the acylation and the deacylation reaction. The pH dependence of the values k(cat)/K(m) for three substrates shows identical pK(a)s of 5.5. and 8.4. This identity combined with the fact that only one of these three substrates is kinetically "sticky" proves that the substrates can combine productively with only one protonic form of the enzyme. There is considerable substrate variation in the pK(a) values of k(cat) observed vs. pH profiles; the inflection points for all substrates studied are at pH values more extreme than are observed in the pH profiles for k(cat)/K(m).  相似文献   

19.
Malate synthase catalyzes the Claisen-like condensation of acetyl-coenzyme A (AcCoA) and glyoxylate in the glyoxylate shunt of the citric acid cycle. The Mycobacterium tuberculosis malate synthase G gene, glcB, was cloned, and the N-terminal His(6)-tagged 80 kDa protein was expressed in soluble form and purified by metal affinity chromatography. A chromogenic 4,4'-dithiodipyridine assay did not yield linear kinetics, but the generation of an active site-directed mutant, C619S, gave an active enzyme and linear kinetics. The resulting mutant exhibited kinetics comparable to those of the wild type and was used for the full kinetic analysis. Initial velocity studies were intersecting, suggesting a sequential mechanism, which was confirmed by product and dead-end inhibition. The inhibition studies delineated the ordered binding of glyoxylate followed by AcCoA and the ordered release of CoA followed by malate. The pH dependencies of k(cat) and k(cat)/K(gly) are both bell-shaped, and catalysis depends on a general base (pK = 5.3) and a general acid (pK = 9.2). Primary kinetic isotope effects determined using [C(2)H(3)-methyl]acetyl-CoA suggested that proton removal and carbon-carbon bond formation were partially rate-limiting. Solvent kinetic isotope effects on k(cat) suggested the hydrolysis of the malyl-CoA intermediate was also partially rate-limiting. Multiple kinetic isotope effects, utilizing D(2)O and [C(2)H(3)-methyl]acetyl-CoA, confirmed a stepwise mechanism in which the step exhibiting primary kinetic isotope effects precedes the step exhibiting the solvent isotope effects. We combined the kinetic data and the pH dependence of the kinetic parameters with existing structural and mutagenesis data to propose a chemical mechanism for malate synthase from M. tuberculosis.  相似文献   

20.
Wang J  Edmondson DE 《Biochemistry》2011,50(35):7710-7717
Monoamine oxidase A (MAO A) is a mitochondrial outer membrane-bound flavoenzyme important in the regulation of serotonin and dopamine levels. Because the rat is extensively used as an animal model in drug studies, it is important to understand how rat MAO A behaves in comparison with the more extensively studied human enzyme. For many reversible inhibitors, rat MAO A exhibits K(i) values similar to those of human MAO A. The pH profile of k(cat) for rat MAO A shows a pK(a) of 8.2 ± 0.1 for the benzylamine ES complex and pK(a) values of 7.5 ± 0.1 and 7.6 ± 0.1 for the ES complexes with p-CF(3)-(1)H- and p-CF(3)-(2)H-benzylamine, respectively. In contrast to the human enzyme, the rat enzyme exhibits a single pK(a) value (8.3 ± 0.1) with k(cat)/K(m) for benzylamine versus pH and pK(a) values of 7.8 ± 0.1 and 8.1 ± 0.2 for the ascending limbs, respectively, of k(cat)/K(m) versus pH profiles for p-CF(3)-(1)H- and p-CF(3)-(2)H-benzylamine and 9.3 ± 0.1 and 9.1 ± 0.2 for the descending limbs, respectively. The oxidation of para-substituted benzylamine substrate analogues by rat MAO A has large deuterium kinetic isotope effects on k(cat) and on k(cat)/K(m). These effects are pH-independent and range from 7 to 14, demonstrating a rate-limiting α-C-H bond cleavage step in catalysis. Quantitative structure-activity correlations of log k(cat) with the electronic substituent parameter (σ) at pH 7.5 and 9.0 show a dominant contribution with positive ρ values (1.2-1.3) and a pH-independent negative contribution from the steric term. Quantitative structure-activity relationship analysis of the binding affinities of the para-substituted benzylamine analogues for rat MAO A shows an increased van der Waals volume (V(w)) increases the affinity of the deprotonated amine for the enzyme. These results demonstrate that rat MAO A exhibits functional properties similar but not identical with those of the human enzyme and provide additional support for C-H bond cleavage via a polar nucleophilic mechanism.  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