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.
Hevel JM  Mills SA  Klinman JP 《Biochemistry》1999,38(12):3683-3693
The copper amine oxidases (CAOs) catalyze both the single-turnover modification of a peptidyl tyrosine to form the active-site cofactor 2,4,5-trihydroxyphenylalanine quinone (TPQ) and the oxidative deamination of primary amines using TPQ. The function of a strictly conserved tyrosine located within hydrogen-bonding distance to TPQ has been explored by employing site-directed mutagenesis on the enzyme from H. polymorpha to form the mutants Y305A, Y305C, and Y305F. Both Y305A and Y305C behave similarly with regard to aliphatic amine oxidase activity, showing 3-7-fold decreases in kinetic parameters relative to WT, while the more conservative substitution of Y305F results in a >100-fold decrease in kcat and >500-fold decrease in kcat/Km relative to WT for the reductive half-reaction. The oxidation of benzylamine by all three mutants is severely impaired, with very significant effects seen in the oxidative half-reaction. CAO activity was studied as a function of pH for WT and Y305A proteins. Profiles for WT-catalyzed methylamine oxidation and Y305A-catalyzed ethylamine oxidation are comparable, while profiles of Y305A-catalyzed methylamine oxidation suggest the pH-dependent build-up of an inhibitory intermediate, which was subsequently observed spectrophotometrically and is attributed to the product Schiff base. The relative effects of mutations at Y305 on catalytic turnover are, thus, concluded to be dependent on the nature of the amino acid which substitutes for tyrosine and the substrate used in amine oxidase assays. TPQ biogenesis experiments demonstrate a approximately 800-fold decrease in kobs for apo-Y305A compared to WT. Despite the strict conservation of Tyr305 in all CAOs, neither biogenesis nor catalytic turnover is abolished upon mutation of this residue. We propose an important, but nonessential, role for Tyr305 in the positioning of the TPQ precursor for biogenesis, and in the maintenance of the correct conformation for TPQ-derived intermediates during catalytic turnover.  相似文献   

2.
This study presents the first detailed examination by resonance Raman (RR) spectroscopy of the rates of solvent exchange for the C5 and C3 positions of the TPQ cofactor in several wild-type copper-containing amine oxidases and mutants of the amine oxidase from Hansenula polymorpha (HPAO). On the basis of crystal structure analysis and differing rates of C5 [double bond] O and C3 [bond] H exchange within the enzyme systems, but equally rapid rates of C5 [double bond] O and C3 [bond] H exchange in a TPQ model compound, it is proposed that these data can be used to determine the TPQ cofactor orientation within the active site of the resting enzyme. A rapid rate of C5 [double bond] O exchange (t(1/2) < 30 min) and a slow (t(1/2) = 6 h) to nonexistent rate of C3 [bond] H exchange was observed for wild-type HPAO, the amine oxidase from Arthrobacter globiformis, pea seedling amine oxidase at pH 7.1, and the E406Q mutant of HPAO. This pattern is ascribed to a productive TPQ orientation, with the C5 [double bond] O near the substrate-binding site and the C3 [bond] H near the Cu. In contrast, a slow rate of C5 [double bond] O exchange (t(1/2) = 1.6-3.3 h) coupled with a fast rate of C3 [bond] H exchange (t(1/2) < 30 min) was observed for the D319E and D319N catalytic base mutants of HPAO and for PSAO at pH 4.6 (t(1/2) = 4.5 h for C5 [double bond] O exchange). This pattern identifies a flipped orientation, involving 180 degrees rotation about the C alpha-C beta bond, which locates the C3 [bond] H near the substrate-binding site and the C5 double bond] O near the Cu. Finally, fast rates of both C5 [double bond] O and C3 [bond] H exchange (t(1/2) < 30 min) were observed for the amine oxidase from Escherichia coli and the N404A mutant of HPAO, suggesting a mobile cofactor, with multiple TPQ orientations between productive and flipped. These results demonstrate that opposing sides of the TPQ ring possess different degrees of solvent accessibility and that the rates of C5 [double bond] O and C3 [bond] H exchange can be used to predict the TPQ cofactor orientation in the resting forms of these enzymes.  相似文献   

3.
DuBois JL  Klinman JP 《Biochemistry》2005,44(34):11381-11388
The copper amine oxidases (CAOs) catalyze the O(2)-dependent, two-electron oxidation of amines to aldehydes at an active site that contains Cu(II) and topaquinone (TPQ) cofactor. TPQ arises from the autocatalytic, post-translational oxidation of a tyrosine side chain in the active site. Monooxygenation within the ring of tyrosine at a single Cu(II) site is unique in biology and occurs as an early step in the formation of TPQ. The mechanism of this reaction has been further examined in the CAO from Hansenula polymorpha (HPAO). When a Clark electrode fitted to a custom-made, gastight apparatus over a range of initial concentrations of O(2) was used, rates of O(2) consumption at levels greater than air are seen to be reduced relative to earlier results, yielding K(D)(apparent) = 216 microM for O(2). This is consistent with a mechanism in which O(2) binds reversibly to the active site, triggering a conformational change that promotes ligation of tyrosinate to Cu(II). The activated Cu(II)-tyrosinate species has been proposed to react with O(2) in a rate-limiting step, although it was also possible that breakdown of a putative peroxy-intermediate controlled TPQ formation. To test the latter hypothesis, Cu(II)-free HPAO was prepared with 3,5-ring-[(2)H(2)]-tyrosine incorporated throughout the primary sequence. The absence of an isotope effect on the rate of TPQ formation eliminates cleavage of this C-H bond in a proposed Cu(II)-aryl-peroxide intermediate as a rate limiting step. The role of methionine 634, previously found to moderate O(2) binding during the catalytic cycle, is shown here to serve a similar function in TPQ formation. As with catalysis, the rate of TPQ formation correlates with the volume of the hydrophobic side chain at position 634, implicating similar binding sites for O(2) during catalysis and cofactor biogenesis.  相似文献   

4.
Goto Y  Klinman JP 《Biochemistry》2002,41(46):13637-13643
Copper amine oxidases (CAOs) contain 2,4,5-trihydroxyphenylalanyl quinone (TPQ) and a copper ion in their active sites, catalyzing amine oxidation to aldehyde and ammonia concomitant with the reduction of molecular oxygen to hydrogen peroxide. Kinetic studies on the CAO from bovine serum (BSAO) [Su and Klinman (1999) Biochemistry 37, 12513-12525] and the recent reports on the cobalt substituted form of the enzyme from Hansenula polymorpha (HPAO) [Mills and Klinman (2000) J. Am. Chem. Soc. 122, 9897-9904, and Mills et al. (2002) Biochemistry, 41, 10577-10584] support pre-binding of molecular oxygen prior to a rate-limiting electron transfer from the reduced form of TPQ (p-aminohydroquinone form) to dioxygen. Although there is significant sequence homology between BSAO and HPAO, k(cat)/K(m)(O2) for BSAO under the optimal condition is one order of magnitude lower than that for HPAO. From a comparison of amino acid sequences for BSAO and HPAO, together with the X-ray crystal structure of HPAO, a plausible dioxygen pre-binding site has been identified that involves Y407, L425, and M634 in HPAO; the latter two residues are altered in BSAO to A490 and T695. To determine which of these residues plays a greater role in dioxygen chemistry, k(cat)/K(m)(O2) was determined in HPAO for the M634 --> T and L425 --> A mutants. The L425 --> A mutation does not alter k(cat)/K(m)(O2) to a large extent, whereas the M634 --> T decreased k(cat)/K(m)(O2) by one order of a magnitude, creating a catalyst that is similar to BSAO. A series of mutants at M634 (to F, L, and Q) were, therefore, prepared in HPAO and characterized with regard to k(cat)/K(m)(O2) as a function of pH. Structure reactivity correlations show a linear relationship of rate with side chain volume, rather than hydrophobicity, indicating that dioxygen reactivity increases with the bulk of the residue at position 634. This site also shows specificity for O2, in relation to the co-gas N2, since substitution of the inert gas N2 by either Ar or He has no effect on measured rates. In particular, He gas is expected to have little affinity for protein at 1 atmospheric pressure, implying little or no binding by N2 as well.  相似文献   

5.
Copper amine oxidases are homodimeric enzymes that catalyze two reactions: first, a self-processing reaction to generate the 2,4,5-trihydroxyphenylalanine (TPQ) cofactor from an active site tyrosine by a single turnover mechanism; second, the oxidative deamination of primary amine substrates with the production of aldehyde, hydrogen peroxide, and ammonia catalyzed by the mature enzyme. The importance of active site residues in both of these processes has been investigated by structural studies and site-directed mutagenesis in enzymes from various organisms. One conserved residue is a tyrosine, Tyr369 in the Escherichia coli enzyme, whose hydroxyl is hydrogen bonded to the O4 of TPQ. To explore the importance of this site, we have studied a mutant enzyme in which Tyr369 has been mutated to a phenylalanine. We have determined the X-ray crystal structure of this variant enzyme to 2.1 A resolution, which reveals that TPQ adopts a predominant nonproductive conformation in the resting enzyme. Reaction of the enzyme with the irreversible inhibitor 2-hydrazinopyridine (2-HP) reveals differences in the reactivity of Y369F compared with wild type with more efficient formation of an adduct (lambda(max) = 525 nm) perhaps reflecting increased mobility of the TPQ adduct within the active site of Y369F. Titration with 2-HP also reveals that both wild type and Y369F contain one TPQ per monomer, indicating that Tyr369 is not essential for TPQ formation, although we have not measured the rate of TPQ biogenesis. The UV-vis spectrum of the Y369F protein shows a broader peak and red-shifted lambda(max) at 496 nm compared with wild type (480 nm), consistent with an altered electronic structure of TPQ. Steady-state kinetic measurements reveal that Y369F has decreased catalytic activity particularly below pH 6.5 while the K(M) for substrate beta-phenethylamine increases significantly, apparently due to an elevated pK(a) (5.75-6.5) for the catalytic base, Asp383, that should be deprotonated for efficient binding of protonated substrate. At pH 7.0, the K(M) for wild type and Y369F are similar at 1.2 and 1.5 microM, respectively, while k(cat) is decreased from 15 s(-1) in wild type to 0.38 s(-1), resulting in a 50-fold decrease in k(cat)/K(M) for Y369F. Transient kinetics experiments indicate that while the initial stages of enzyme reduction are slower in the variant, these do not represent the rate-limiting step. Previous structural and solution studies have implicated Tyr369 as a component of a proton shuttle from TPQ to dioxygen. The moderate changes in kinetic parameters observed for the Y369F variant indicate that if this is the case, then the absence of the Tyr369 hydroxyl can be compensated for efficiently within the active site.  相似文献   

6.
Mills SA  Goto Y  Su Q  Plastino J  Klinman JP 《Biochemistry》2002,41(34):10577-10584
A recent report by Mills and Klinman [Mills, S. A., and Klinman, J. P. (2000) J. Am. Chem. Soc. 122, 9897-9904] described the preparation and initial characterization of a cobalt-substituted form of the copper amine oxidase from Hansenula polymorpha (HPAO). This enzyme was found to be fully catalytically active at saturating substrate concentrations, but with a K(m) for O(2) approximately 70-fold higher than that of the copper-containing, wild-type enzyme. Herein, we report a detailed analysis of the mechanism of catalysis for the wild-type and the cobalt-substituted forms of HPAO. Both forms of enzyme are concluded to utilize the same mechanism for oxygen reduction, involving initial, rate-limiting electron transfer from the reduced cofactor of the enzyme to prebound dioxygen. Superoxide formed in this manner is stabilized by the active site metal, facilitating the transfer of a second electron and two protons to form the product hydrogen peroxide. The elevated K(m) for O(2) at the dioxygen binding site in Co-substituted HPAO, relative to that of wild-type HPAO, is proposed to be due to a change in the net charge at the adjacent metal site from +1 (cupric hydroxide) in wild-type enzyme to +2 (cobaltous H(2)O) in cobalt-substituted HPAO.  相似文献   

7.
Adduct I (lambda(max) at approximately 430 nm) formed in the reaction of 2-hydrazinopyridine (2HP) and the TPQ cofactor of wild-type Escherichia coli copper amine oxidase (WT-ECAO) is stable at neutral pH, 25 degrees C, but slowly converts to another spectroscopically distinct species with a lambda(max) at approximately 530 nm (adduct II) at pH 9.1. The conversion was accelerated either by incubation of the reaction mixture at 60 degrees C or by increasing the pH (>13). The active site base mutant forms of ECAO (D383N and D383E) showed spectral changes similar to WT when incubated at 60 degrees C. By contrast, in the Y369F mutant adduct I was not stable at pH 7, 25 degrees C, and gradually converted to adduct II, and this rate of conversion was faster at pH 9. To identify the nature of adduct II, we have studied the effects of pH and divalent cations on the UV-vis and resonance Raman spectroscopic properties of the model compound of adduct I (2). Strikingly, it was found that addition of Cu2+ to 2 at pH 7 gave a product (3) that exhibited almost identical spectroscopic signatures to adduct II. The X-ray crystal structure of 3 shows that it is the copper-coordinated form of 2, where the +2 charge of copper is neutralized by a double deprotonation of 2. These results led to the proposal that adduct II in the enzyme is TPQ-2HP that has migrated onto the active site Cu2+. The X-ray crystal structure of Y369F adduct II confirmed this assignment. Resonance Raman and EPR spectroscopy showed that adduct II in WT-ECAO is identical to that seen in Y369F. This study clearly demonstrates that the hydrogen-bonding interaction between O4 of TPQ and the conserved Tyr (Y369) is important in controlling the position and orientation of TPQ in the catalytic cycle, including optimal orientation for reactivity with substrate amines.  相似文献   

8.
Samuels NM  Klinman JP 《Biochemistry》2005,44(43):14308-14317
Copper amine oxidase (CAO) is a dual-functioning enzyme that catalyzes the biosynthesis of a self-derived coenzyme and subsequent oxidative deamination of primary amines. The organic cofactor, 2,4,5-trihydroxyphenylalanine quinone (TPQ), is generated from the post-translational modification of an active site tyrosine (Y405) in a reaction shown to be dependent on both molecular oxygen and a mononuclear copper center. Previous investigations of Cu(II)-dependent cofactor formation in the Hansenula polymorpha amine oxidase (HPAO) provided evidence for the coordination of the precursor tyrosine in forming a ligand-to-metal charge transfer complex as a means of activating the tyrosyl ring for direct attack by triplet-state dioxygen. To further delineate the role of the metal in facilitating this complex series of reactions, apo-HPAO was reconstituted with alternate metals of varying reduction potentials and Lewis acidities [Ni(II), Co(II), Mn(II), Fe(II), and Fe(III)] and the consequence of each substitution on TPQ biogenesis examined. Ni(II) was found to support the transformation of the precursor tyrosine to the quinone cofactor to yield a mature enzyme competent for methylamine oxidation. Detailed kinetic analysis of the mechanism of TPQ biogenesis for the Ni(II)-substituted enzyme has led to the proposal of a direct electron transfer from the metal-coordinated tyrosinate to dioxygen as the dominant rate-limiting step.  相似文献   

9.
Copper amine oxidases (CAOs) catalyze the two-electron oxidation of primary amines to aldehydes, utilizing molecular oxygen as a terminal electron acceptor. To accomplish this transformation, CAOs utilize two cofactors: a mononuclear copper, and a unique redox cofactor, 2,4,5-trihydroxyphenylalanine quinone (TPQ or TOPA quinone). TPQ is derived via posttranslational modification of a specific tyrosine residue within the protein itself. In this study, the structure of an amine oxidase from Hansenula polymorpha has been solved to 2.5 A resolution, in which the precursor tyrosine is unprocessed to TPQ, and the copper site is occupied by zinc. Significantly, the precursor tyrosine directly ligands the metal, thus providing the closest analogue to date of an intermediate in TPQ production. Besides this result, the rearrangement of other active site residues (relative to the mature enzyme) proposed to be involved in the binding of molecular oxygen may shed light on how CAOs efficiently use their active site to carry out both cofactor formation and catalysis.  相似文献   

10.
2-Hydrazinopyridine (2HP) is an irreversible inhibitor of copper amine oxidases (CAOs). 2HP reacts directly at the C5 position of the TPQ cofactor, yielding an intense chromophore with lambda(max) approximately 430 nm (adduct I) in Escherichia coli amine oxidase (ECAO). The adduct I form of wild type (WT-ECAO) was assigned as a hydrazone on the basis of the X-ray crystal structure. The hydrazone adduct appears to be stabilized by two key hydrogen-bonding interactions between the TPQ-2HP moiety and two active site residues: the catalytic base (D383) and the conserved tyrosine residue (Y369). In this work, we have synthesized a model compound (2) for adduct I from the reaction of a TPQ model compound (1) and 2HP. NMR spectroscopy and X-ray crystallography show that 2 exists predominantly as the azo form (lambda(max) at 414 nm). Comparison of the UV-vis and resonance Raman spectra of 2 with adduct I in WT, D383E, D383N, and Y369F forms of ECAO revealed that adduct I in WT and D383N is a tautomeric mixture where the hydrazone form is favored. In D383E adduct I, the equilibrium is further shifted in favor of the hydrazone form. UV-vis spectroscopic pH titrations of adduct I in WT, D383N, D383E, and 2 confirmed that D383 in WT adduct I is protonated at pH 7 and stabilizes the hydrazone tautomer by a short hydrogen-bonding interaction. The deprotonation of D383 (pKa approximately 9.7) in adduct I resulted in conversion of adduct I to the azo tautomer with a blue shift of the lambda(max) to 420 nm, close to that of 2. In contrast, adduct I in D383N and D383E is stable and did not show any pH-dependent spectral changes. In Y369F, adduct I was not stable and gradually converted into a new species with lambda(max) at approximately 530 nm (adduct II). A detailed mechanism for the adduct I formation in WT has been proposed that is consistent with the mechanism proposed for the oxidation of substrate by CAOs but addresses some key differences in the active site chemistry of hydrazine inhibitors and substrate amines.  相似文献   

11.
Copper amine oxidases possess the unusual ability to generate autocatalytically their organic cofactor, which is subsequently utilized in turnover. This cofactor, 2,4,5-trihydroxyphenylalanine quinone (TPQ), is formed within the active site of these enzymes by the oxidation of a single tyrosine residue. In vitro, copper(II) and oxygen are both necessary and sufficient for the conversion of tyrosine to TPQ. In this study, the biogenesis of TPQ has been characterized in an amine oxidase from Hansenula polymorpha expressed as the apo-enzyme in Escherichia coli. With the WT enzyme, optical absorbances which are copper or oxygen dependent are observed and characterized. Active-site mutants are used to investigate further the nature of these spectral species. Evidence is presented which suggests that tyrosine is activated for reaction with oxygen by liganding to Cu(II). In the following paper in this issue [Schwartz, B., Dove, J. E., and Klinman, J. P. (2000) Biochemistry 39, 3699-3707], the initial reaction of precursor protein with oxygen is characterized kinetically. Taken together, the available data suggest a mechanism for the oxidation of tyrosine to TPQ where the role of the copper is to activate substrate.  相似文献   

12.
Copper amine oxidases (CAOs) are a family of redox active enzymes containing a 2,4,5-trihydroxyphenylalanine quinone (TPQ) cofactor generated from post translational modification of an active site tyrosine residue. The Arthrobacter globiformis amine oxidase (AGAO) has been widely used as a model to guide the design and development of selective inhibitors of CAOs. In this study, two aryl 2,3-butadienamine analogs, racemic 5-phenoxy-2,3-pentadienylamine (POPDA) and racemic 6-phenyl-2,3-hexadienylamine (PHDA), were synthesized and evaluated as mechanism-based inactivators of AGAO. Crystal structures show that both compounds form a covalent adduct with the amino group of the substrate-reduced TPQ, and that the chemical structures of the rac-PHDA and rac-POPDA modified TPQ differ by the allenic carbon that is attached to the cofactor. A chemical mechanism accounting for the formation of the respective TPQ derivative is proposed. Under steady-state conditions, no recovery of enzyme activity is detected when AGAO pre-treated with rac-PHDA or rac-POPDA is diluted with excess amount of the benzylamine substrate (100-fold K(m)). Comparing the IC(50) values further reveals that the phenoxy substituent in POPDA offers an approximately 4-fold increase in inhibition potency, which can be attributed to a favourable binding interaction between the oxygen atom in the phenoxy group and the active site of AGAO as revealed by crystallographic studies. This hypothesis is corroborated by the observed >3-fold higher partition ratio of PHDA compared to POPDA. Taken together, the results presented in this study reveal the mechanism by which aryl 2,3-butadienamines act as mechanism-based inhibitors of AGAO, and the potency of enzyme inactivation could be fine-tuned by optimizing binding interaction between the aryl substituent and the enzyme active site.  相似文献   

13.
GDP-mannose hydrolase (GDPMH) catalyzes the hydrolysis of GDP-alpha-d-sugars by nucleophilic substitution with inversion at the anomeric C1 atom of the sugar, with general base catalysis by H124. Three lines of evidence indicate a mechanism with dissociative character. First, in the 1.3 A X-ray structure of the GDPMH-Mg(2+)-GDP.Tris(+) complex [Gabelli, S. B., et al. (2004) Structure 12, 927-935], the GDP leaving group interacts with five catalytic components: R37, Y103, R52, R65, and the essential Mg(2+). As determined by the effects of site-specific mutants on k(cat), these components contribute factors of 24-, 100-, 309-, 24-, and >/=10(5)-fold, respectively, to catalysis. Both R37 and Y103 bind the beta-phosphate of GDP and are only 5.0 A apart. Accordingly, the R37Q/Y103F double mutant exhibits partially additive effects of the two single mutants on k(cat), indicating cooperativity of R37 and Y103 in promoting catalysis, and antagonistic effects on K(m). Second, the conserved residue, D22, is positioned to accept a hydrogen bond from the C2-OH group of the sugar undergoing substitution at C1, as was shown by modeling an alpha-d-mannosyl group into the sugar binding site. The D22A and D22N mutations decreased k(cat) by factors of 10(2.1) and 10(2.6), respectively, for the hydrolysis of GDP-alpha-d-mannose, and showed smaller effects on K(m), suggesting that the D22 anion stabilizes a cationic oxocarbenium transition state. Third, the fluorinated substrate, GDP-2F-alpha-d-mannose, for which a cationic oxocarbenium transition state would be destabilized by electron withdrawal, exhibited a 16-fold decrease in k(cat) and a smaller, 2.5-fold increase in K(m). The D22A and D22N mutations further decreased the k(cat) with GDP-2F-alpha-d-mannose to values similar to those found with GDP-alpha-d-mannose, and decreased the K(m) of the fluorinated substrate. The choice of histidine as the general base over glutamate, the preferred base in other Nudix enzymes, is not due to the greater basicity of histidine, since the pK(a) of E124 in the active complex (7.7) exceeded that of H124 (6.7), and the H124E mutation showed a 10(2.2)-fold decrease in k(cat) and a 4.0-fold increase in K(m) at pH 9.3. Similarly, the catalytic triad detected in the X-ray structure (H124- - -Y127- - -P120) is unnecessary for orienting H124, since the Y127F mutation had only 2-fold effects on k(cat) and K(m) with either H124 or E124 as the general base. Hence, a neutral histidine rather than an anionic glutamate may be necessary to preserve electroneutrality in the active complex.  相似文献   

14.
We have demonstrated that amino acids E (323), Y (324), E (330), and V (331) from the factor Va heavy chain are required for the interaction of the cofactor with factor Xa and optimum rates of prothrombin cleavage. We have also shown that amino acid region 332-336 contains residues that are important for cofactor function. Using overlapping peptides, we identified amino acids D (334) and Y (335) as contributors to cofactor activity. We constructed recombinant factor V molecules with the mutations D (334) --> K and Y (335) --> F (factor V (KF)) and D (334) --> A and Y (335) --> A (factor V (AA)). Kinetic studies showed that while factor Va (KF) and factor Va (AA) had a K D for factor Xa similar to the K D observed for wild-type factor Va (factor Va (WT)), the clotting activities of the mutant molecules were impaired and the k cat of prothrombinase assembled with factor Va (KF) and factor Va (AA) was reduced. The second-order rate constant of prothrombinase assembled with factor Va (KF) or factor Va (AA) for prothrombin activation was approximately 10-fold lower than the second-order rate constant for the same reaction catalyzed by prothrombinase assembled with factor Va (WT). We also created quadruple mutants combining mutations in the amino acid region 334-335 with mutations at the previously identified amino acids that are important for factor Xa binding (i.e., E (323)Y (324) and E (330)V (331)). Prothrombinase assembled with the quadruple mutant molecules displayed a second-order rate constant up to 400-fold lower than the values obtained with prothrombinase assembled with factor Va (WT). The data demonstrate that amino acid region 334-335 is required for the rearrangement of enzyme and substrate necessary for efficient catalysis of prothrombin by prothrombinase.  相似文献   

15.
R67 dihydrofolate reductase (DHFR) is a novel protein that possesses 222 symmetry. A single active site pore traverses the length of the homotetramer. Although the 222 symmetry implies that four symmetry-related binding sites should exist for each substrate as well as each cofactor, isothermal titration calorimetry (ITC) studies indicate only two molecules bind. Three possible combinations include two dihydrofolate molecules, two NADPH molecules, or one substrate with one cofactor. The latter is the productive ternary complex. To evaluate the roles of A36, Y46, T51, G64, and V66 residues in binding and catalysis, a site-directed mutagenesis approach was employed. One mutation per gene produces four mutations per active site pore, which often result in large cumulative effects. Conservative mutations at these positions either eliminate the ability of the gene to confer trimethoprim resistance or have no effect on catalysis. This result, in conjunction with previous mutagenesis studies on K32, K33, S65, Q67, I68, and Y69 [Strader, M. B., et al. (2001) Biochemistry 40, 11344-11352; Hicks, S. N., et al. (2003) Biochemistry 42, 10569-10578; Park, H., et al. (1997) Protein Eng. 10, 1415-1424], allows mapping of the active site surface. Residues for which conservative mutations have large effects on binding and catalysis include K32, Q67, I68, and Y69. These residues form a stripe that establishes the ligand binding surface. Residues that accommodate conservative mutations that do not greatly affect catalysis include K33, Y46, T51, S65, and V66. Isothermal titration calorimetry studies were also conducted on many of the mutants described above to determine the enthalpy of folate binding to the R67 DHFR.NADPH complex. A linear correlation between this DeltaH value and log k(cat)/K(m) is observed. Since structural tightness appears to be correlated with the exothermicity of the binding interaction, this leads to the hypothesis that enthalpy-driven formation of the ternary complex in these R67 DHFR variants plays a strong role in catalysis. Use of the alternate cofactor, NADH, extends this correlation, indicating preorganization of the ternary complex determines the efficiency of the reaction. This hypothesis is consistent with data suggesting R67 DHFR uses an endo transition state (where the nicotinamide ring of cofactor overlaps the more bulky side of the substrate's pteridine ring).  相似文献   

16.
Amine oxidases utilize a proton abstraction mechanism following binding of the amine substrate to the C5 position of the cofactor, the quinone form of trihydroxyphenylalanine (TPQ). Previous work [Wilmot, C. M., et al. (1997) Biochemistry 36, 1608-1620] has shown that Asp383 in Escherichia coliamine oxidase (ECAO) is the catalytic base which performs the key step of proton abstraction. This paper explores in more depth this and other roles of Asp383. The crystal structures of three mutational variants are presented together with their catalytic properties, visible spectra, and binding properties for a substrate-like inhibitor, 2-hydrazinopyridine (2-HP), in comparison to those of the wild type enzyme. In wild type ECAO, the TPQ is located in a wedge-shaped pocket which allows more freedom of movement at the substrate binding position (C5) than for TPQ ring carbons C1-C4. A role of Asp383, whose carboxylate is located close to O5, is to stabilize the TPQ in its major conformation in the pocket. Replacement of Asp383 with the isostructural, but chemically distinct, Asn383 does not affect the location or dynamics of the TPQ cofactor significantly, but eliminates catalytic activity and drastically reduces the affinity for 2-HP. Removal of the side chain carboxyl moiety, as in Ala383, additionally allows the TPQ the greater conformational flexibility to coordinate to the copper, which demonstrates that Asp383 helps maintain the active site structure by preventing TPQ from migrating to the copper. Glu383 has a greatly decreased catalytic activity, as well as a decreased affinity for 2-HP relative to that of wild type ECAO. The electron density reveals that the longer side chain of Glu prevents the pivotal motion of the TPQ by hindering its movement within the wedge-shaped active site pocket. The results show that Asp383 performs multiple roles in the catalytic mechanism of ECAO, not only in acting as the active site base at different stages of the catalytic cycle but also in regulating the mobility of the TPQ that is essential to catalysis.  相似文献   

17.
This review will focus on how X-ray crystallographic studies of copper-containing amine oxidases have complemented the solution, kinetic, and spectroscopic research on this ubiquitous class of enzymes. These enzymes not only contain a copper ion at the active site, but also a unique organic cofactor, 2,4,5-trihydroxyphenylalanine quinone (TPQ), which is absolutely required for catalysis. Structural data have not only shed light on the catalytic mechanism of the enzyme, which converts primary amines, using molecular oxygen, to aldehydes, ammonia, and hydrogen peroxide, but also on biogenesis of the cofactor. The cofactor is derived from a tyrosine in the enzyme amino acid sequence and requires only the addition of copper(II) and molecular oxygen in a self-processing event.  相似文献   

18.
To gain insight into the role of the strictly conserved histidine residue, H178, in the reaction mechanism of the methionyl aminopeptidase from Escherichia coli (EcMetAP-I), the H178A mutant enzyme was prepared. Metal-reconstituted H178A binds only one equivalent of Co(II) or Fe(II) tightly with affinities that are identical to the WT enzyme based on kinetic and isothermal titration calorimetry (ITC) data. Electronic absorption spectra of Co(II)-loaded H178A EcMetAP-I indicate that the active site divalent metal ion is pentacoordinate, identical to the WT enzyme. These data indicate that the metal binding site has not been affected by altering H178. The effect of altering H178 on activity is, in general, due to a decrease in k(cat). The k(cat) value for Co(II)-loaded H178A decreased 70-fold toward MGMM and 290-fold toward MP-p-NA compared to the WT enzyme, while k(cat) decreased 50-fold toward MGMM for the Fe(II)-loaded H178A enzyme and 140-fold toward MP-p-NA. The K(m) values for MGMM remained unaffected, while those for MP-p-NA increased approximately 2-fold for Co(II)- and Fe(II)-loaded H178A. The k(cat)/K(m) values for both Co(II)- and Fe(II)-loaded H178A toward both substrates ranged from approximately 50- to 580-fold reduction. The pH dependence of log K(m), log k(cat), and log(k(cat)/K(m)) of both WT and H178A EcMetAP-I were also obtained and are identical, within error, for H178A and WT EcMetAP-I. Therefore, H178A is catalytically important but is not required for catalysis. Assignment of one of the observed pK(a) values at 8.1 for WT EcMetAP-I was obtained from plots of molar absorptivity at lambda(max(640)) vs pH for both WT and H178A EcMetAP-I. Apparent pK(a) values of 8.1 and 7.6 were obtained for WT and H178A EcMetAP-I, respectively, and were assigned to the deprotonation of a metal-bound water molecule. The data reported herein provide support for the key elements of the previously proposed mechanism and suggest that a similar mechanism can apply to the enzyme with a single metal in the active site.  相似文献   

19.
The electrochemical behavior of redox centers in the active site of amine oxidases from lentil seedlings and Euphorbia characias latex was investigated using a mercury film electrode. Tyrosine-derived 6-hydroxydopa quinone (TPQ) and copper ions in the active site are redox centers of these amine oxidases. The enzymes undergo two reduction processes at negative potentials related to the reduction of the TPQ cofactor to the corresponding hydroquinones and the reduction of copper ions, (Cu(II)-->Cu(I)). Copper depleted enzymes, prepared by reduction with dithionite followed by dialysis against cyanide, undergo only one reduction process. Nyquist diagrams, recorded at potentials corresponding to the reduction of cofactors as dc-offset, represent charge transfer impedance followed by a Warburg-type line at low frequencies, indicating the occurrence of a diffusion controlled process in the rate-limiting step of the reduction process.  相似文献   

20.
Zheng R  Blanchard JS 《Biochemistry》2000,39(51):16244-16251
Ketopantoate reductase (EC 1.1.1.169) catalyzes the NADPH-dependent reduction of alpha-ketopantoate to D-(-)-pantoate in the biosynthesis of pantothenate. The pH dependence of V and V/K for the E. coli enzyme suggests the involvement of a general acid/base in the catalytic mechanism. To identify residues involved in catalysis and substrate binding, we mutated the following six strictly conserved residues to Ala: Lys72, Lys176, Glu210, Glu240, Asp248, and Glu256. Of these, the K176A and E256A mutant enzymes showed 233- and 42-fold decreases in V(max), and 336- and 63-fold increases in the K(m) value of ketopantoate, respectively, while the other mutants exhibited WT kinetic properties. The V(max) for the K176A and E256A mutant enzymes was markedly increased, up to 25% and 75% of the wild-type level, by exogenously added primary amines and formate, respectively. The rescue efficiencies for the K176A and E256A mutant enzymes were dependent on the molecular volume of rescue agents, as anticipated for a finite active site volume. The protonated form of the amine is responsible for recovery of activity, suggesting that Lys176 functions as a general acid in catalysis of ketopantoate reduction. The rescue efficiencies for the K176A mutant by primary amines were independent of the pK(a) value of the rescue agents (Bronsted coefficient, alpha = -0.004 +/-0.008). Insensitivity to acid strength suggests that the chemical reaction is not rate-limiting, consistent with (a) the catalytic efficiency of the wild-type enzyme (k(cat)/K(m) = 2x10(6) M(-1) s(-1) and (b) the small primary deuterium kinetic isotope effects, (D)V = 1.3 and (D)V/K = 1.5, observed for the wild-type enzyme. Larger primary deuterium isotope effects on V and V/K were observed for the K176A mutant ((D)V = 3.0, (D)V/K = 3.7) but decreased nearly to WT values as the concentration of ethylamine was increased. The nearly WT activity of the E256A mutant in the presence of formate argues for an important role for this residue in substrate binding. The double mutant (K176A/E256A) has no detectable ketopantoate reductase activity. These results indicate that Lys176 and Glu256 of the E. coli ketopantoate reductase are active site residues, and we propose specific roles for each in binding ketopantoate and catalysis.  相似文献   

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