首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
2-Methyl-3-hydroxypyridine-5-carboxylic acid (MHPC) oxygenase (MHPCO) and 5-pyridoxic acid oxygenase are flavoenzymes catalyzing an aromatic hydroxylation and a ring-cleavage reaction. Both enzymes are involved in biodegradation of vitamin B6 in bacteria. Oxygen-tracer experiments have shown that the enzymes are monooxygnases since only one atom of molecular oxygen is incorporated into the products. Kinetics of MHPCO has shown that the enzyme is similar to single-component flavoprotein hydroxylases in that the binding of MHPC is required prior to the flavin reduction by NADH, and C4a-hydroperoxy-FAD and C4a-hydroxy-FAD are found as intermediates. Investigation on the protonation status of the substrate upon binding to the enzyme has shown that only the tri-ionic form of MHPC is bound at the MHPCO active site. Using a series of FAD analogues with substituents at the 8-position of the isoalloxazine ring, the oxygenation of MHPC by the C4a-hydroperoxy-FAD was shown to occur via an electrophilic aromatic substitution mechanism. Recently, the X-ray structures of MHPCO and a complex of MHPC-MHPCO at 2.1 Å have been reported and show the presence of nine water molecules in the enzyme active site. Based on structural data, a few residues, Tyr82, Tyr223, Arg181, were suggested to be important for catalysis of MHPCO.  相似文献   

2.
Many flavoenzymes catalyze hydroxylation of aromatic compounds especially phenolic compounds have been isolated and characterized. These enzymes can be classified as either single‐component or two‐component flavin‐dependent hydroxylases (monooxygenases). The hydroxylation reactions catalyzed by the enzymes in this group are useful for modifying the biological properties of phenolic compounds. This review aims to provide an in‐depth discussion of the current mechanistic understanding of representative flavin‐dependent monooxygenases including 3‐hydroxy‐benzoate 4‐hydroxylase (PHBH, a single‐component hydroxylase), 3‐hydroxyphenylacetate 4‐hydroxylase (HPAH, a two‐component hydroxylase), and other monooxygenases which catalyze reactions in addition to hydroxylation, including 2‐methyl‐3‐hydroxypyridine‐5‐carboxylate oxygenase (MHPCO, a single‐component enzyme that catalyzes aromatic‐ring cleavage), and HadA monooxygenase (a two‐component enzyme that catalyzes additional group elimination reaction). These enzymes have different unique structural features which dictate their reactivity toward various substrates and influence their ability to stabilize flavin intermediates such as C4a‐hydroperoxyflavin. Understanding the key catalytic residues and the active site environments important for governing enzyme reactivity will undoubtedly facilitate future work in enzyme engineering or enzyme redesign for the development of biocatalytic methods for the synthesis of valuable compounds.  相似文献   

3.
Zhao G  Jorns MS 《Biochemistry》2002,41(31):9747-9750
Monomeric sarcosine oxidase (MSOX) contains covalently bound FAD and catalyzes the oxidation of sarcosine (N-methylglycine) and other secondary amino acids, including L-proline. The reductive half-reaction with L-proline proceeds via a rapidly attained equilibrium (K(d)) between free E(ox) and the E(ox).S complex, followed by a practically irreversible reduction step (E(ox).S --> E(red).P) associated with a rate constant, k(lim). The effect of pH on the reductive half-reaction shows that the K(d) for L-proline binding is pH-independent (pH 6.46-9.0). This indicates that MSOX binds the zwitterionic form of L-proline, the predominant species in solution at neutral pH (pK(a) = 10.6). Values for the limiting rate of reduction (k(lim)) are, however, strongly pH-dependent and indicate that an ionizable group in the E(ox).L-proline complex (pK(a) = 8.02) must be unprotonated for conversion to E(red).P. Charge-transfer interaction with L-proline as the donor and FAD as acceptor is possible only with the anionic form of L-proline. The ionizable group in the E(ox).L-proline complex is required for conversion of enzyme-bound L-proline from the zwitterionic to the reactive anionic form, as judged by the independently determined pK(a) for charge-transfer complex formation with the MSOX flavin (pK(a) = 7.94). The observation that the anionic form of L-proline with a neutral amino group is the reactive species in the reduction of MSOX is similar to that observed for other flavoenzymes that oxidize amines, including monoamine oxidase and trimethylamine dehydrogenase.  相似文献   

4.
1. The first metabolic step in the biodegradation of 4-hydroxypyridine by an Agrobacterium sp. was hydroxylation to form pyridine-3,4-diol. 2. Extracts required 1mol of O(2) and 1mol of NADH or NADPH for the conversion of 4-hydroxypyridine into pyridine-3,4-diol, suggesting that the enzyme responsible, 4-hydroxypyridine-3-hydroxylase, was a mixed function mono-oxygenase. 3. After treatment with acidic (NH(4))(2)SO(4) the enzyme required FAD for activity; FMN and riboflavin would not substitute for FAD. 4. The rate of anaerobic reduction of FAD by NAD(P)H was increased more than tenfold in the presence of 4-hydroxypyridine, suggesting that the mechanism of hydroxylation was similar to that of other aromatic hydroxylases which are of the mono-oxygenase type. 5. The partially purified enzyme was extremely specific for its heterocyclic substrate but would utilize either NADH or NADPH. 6. 4-Hydroxypyridine-3-hydroxylase was strongly inhibited by high substrate concentration (above 0.5mm) especially below pH7.5. 8. The inflexion at pH8.4 in a pK(m) versus pH plot, together with strong inhibition by p-chloromercuribenzoate, suggested a role for thiol groups in substrate binding.  相似文献   

5.
Enterococcus faecalis dihydroorotate dehydrogenase B is a heterodimer of 28 and 33 kDa encoded by the pyrK and pyrDb genes. Both subunits copurify during all chromatographic steps, and, as determined by HPLC, one FMN and one FAD are bound per heterodimer. The enzyme catalyzes efficient oxidation of 4-S-NADH by orotate. Isotope effect and pH data suggest that reduction of flavin by NADH at the PyrK site is only partially rate limiting with no kinetically significant proton transfer occurring in the reductive half-reaction; therefore, a group exhibiting a pK of 5.7 +/- 0.2 represents a residue involved in binding of NADH rather than in catalysis. The reducing equivalents are shuttled between the NADH-oxidizing flavin in PyrK and the orotate-reacting flavin in PyrDb, by iron-sulfur centers through flavin semiquinones as intermediates. A solvent kinetic isotope effect of 2.5 +/- 0.2 on V is indicative of rate-limiting protonation in the oxidative half-reaction and most likely reflects the interaction between the isoalloxazine N1 of the orotate-reducing flavin and Lys 168 (by analogy with L. lactis DHODase A). The oxidative half-reaction is facilitated by deprotonation of the group(s) with pK(s) of 5.8-6.3 and reflects either deprotonation of the reduced flavin or binding of orotate; this step is followed by hydride transfer to C6 and general acid-assisted protonation (pK of 9.1 +/- 0.2) at C5 of the product.  相似文献   

6.
The FAD of p-hydroxybenzoate hydroxylase (PHBH) is known to exist in two conformations. The FAD must be in the in-position for hydroxylation of p-hydroxybenzoate (pOHB), whereas the out-position is essential for reduction of the flavin by NADPH. In these investigations, we have used 8-mercapto-FAD and 8-hydroxy-FAD to probe the movement of the flavin in catalysis. Under the conditions employed, 8-mercapto-FAD (pK(a) = 3.8) and 8-hydroxy-FAD (pK(a) = 4.8) are mainly anionic. The spectral characteristics of the anionic forms of these flavins are very sensitive to their environment, making them sensitive probes for detecting movement of the flavin during catalysis. With these flavin analogues, the enzyme hydroxylates pOHB efficiently, but at a rate much slower than that of enzyme with FAD. Reaction of oxygen with reduced forms of these modified enzymes in the absence of substrate appears to proceed through the formation of the flavin-C4a-hydroperoxide intermediate, as with normal enzyme, but the decay of this intermediate is so fast compared to its formation that very little accumulates during the reaction. However, after elimination of H2O2 from the flavin-C4a-hydroperoxide, a perturbed oxidized enzyme spectrum is observed (Eox*), and this converts slowly to the spectrum of the resting oxidized form of the enzyme (Eox). In the presence of pOHB, PHBH reconstituted with 8-mercapto-FAD also shows the additional oxidized intermediate (Eox*) after the usual oxygenated C4a-intermediates have formed and decayed in the course of the hydroxylation reaction. This Eox* to Eox step is postulated to be due to flavin movement. Furthermore, binding of pOHB to resting (Eox) follows a three-step equilibrium mechanism that is also consistent with flavin movement being the rate-limiting step. The rate for the slowest step during pOHB binding is similar to that observed for the conversion of Eox* to Eox during the oxygen reaction in the absence or presence of substrate. Steady-state kinetic analysis of PHBH substituted with 8-mercapto-FAD demonstrated that the apparent k(cat) is also similar to the rate of Eox* conversion to Eox. Presumably, the protein environment surrounding the flavin in Eox* differs slightly from that of the final resting form of the enzyme (Eox).  相似文献   

7.
Kantz A  Gassner GT 《Biochemistry》2011,50(4):523-532
Styrene monooxygenase (SMO) is a two-component flavoenzyme composed of an NADH-specific flavin reductase (SMOB) and FAD-specific styrene epoxidase (NSMOA). NSMOA binds tightly to reduced FAD and catalyzes the stereospecific addition of one atom of molecular oxygen to the vinyl side chain of styrene in the enantioselective synthesis of S-styrene oxide. In this mechanism, molecular oxygen first reacts with NSMOA(FAD(red)) to yield an FAD C(4a)-peroxide intermediate. This species is nonfluorescent and has an absorbance maximum of 382 nm. Styrene then reacts with the peroxide intermediate with a second-order rate constant of (2.6 ± 0.1) × 10(6) M(-1) s(-1) to yield a fluorescent intermediate with an absorbance maximum of 368 nm. We compute an activation free energy of 8.7 kcal/mol for the oxygenation step, in good agreement with that expected for a peroxide-catalyzed epoxidation, and acid-quenched samples recovered at defined time points in the single-turnover reaction indicate that styrene oxide synthesis is coincident with the formation phase of the fluorescent intermediate. These findings support FAD C(4a)-peroxide being the oxygen atom donor and the identity of the fluorescent intermediate as an FAD C(4a)-hydroxide product of the styrene epoxidation. Overall, four pH-dependent rate constants corresponding to peroxyflavin formation (pK(a) = 7.2), styrene epoxidation (pK(a) = 7.7), styrene oxide dissociation (pK(a) = 8.3), and hydroxyflavin dehydration (pK(a) = 7.6) are needed to fit the single-turnover kinetics.  相似文献   

8.
The ornithine hydroxylase from Pseudomonas aeruginosa (PvdA) catalyzes the FAD-dependent hydroxylation of the side chain amine of ornithine, which is subsequently formylated to generate the iron-chelating hydroxamates of the siderophore pyoverdin. PvdA belongs to the class B flavoprotein monooxygenases, which catalyze the oxidation of substrates using NADPH as the electron donor and molecular oxygen. Class B enzymes include the well studied flavin-containing monooxygenases and Baeyer-Villiger monooxygenases. The first two structures of a class B N-hydroxylating monooxygenase were determined with FAD in oxidized (1.9 Å resolution) and reduced (3.03 Å resolution) states. PvdA has the two expected Rossmann-like dinucleotide-binding domains for FAD and NADPH and also a substrate-binding domain, with the active site at the interface between the three domains. The structures have NADP(H) and (hydroxy)ornithine bound in a solvent-exposed active site, providing structural evidence for substrate and co-substrate specificity and the inability of PvdA to bind FAD tightly. Structural and biochemical evidence indicates that NADP+ remains bound throughout the oxidative half-reaction, which is proposed to shelter the flavin intermediates from solvent and thereby prevent uncoupling of NADPH oxidation from hydroxylated product formation.  相似文献   

9.
Styrene monooxygenase (SMO) from Pseudomonas putida S12 is a two-component flavoenzyme composed of the NADH-specific flavin reductase, SMOB, and FAD-specific styrene epoxidase, SMOA. Here, we report the cloning, and expression of native and histidine-tagged versions of SMOA and SMOB and studies of the flavin transfer and styrene oxygenation reactions. In the reductive half-reaction, SMOB catalyzes the two-electron reduction of FAD with a turnover number of 3200 s(-1). Single turnover studies of the reaction of reduced SMOA with substrates indicate the formation of a stable oxygen intermediate with the absorbance characteristics of a flavin hydroperoxide. Based on the results of numerical simulations of the steady-state mechanism of SMO, we find that the observed coupling of NADH and styrene oxidation can be best explained by a model, which includes both the direct transfer and passive diffusion of reduced FAD from SMOB to SMOA.  相似文献   

10.
Rabeh WM  Alguindigue SS  Cook PF 《Biochemistry》2005,44(14):5541-5550
O-Acetylserine sulfhydrylase (OASS) catalyzes the last step in the cysteine biosynthetic pathway in enteric bacteria and plants, substitution of the beta-acetoxy group of O-acetyl-l-serine (OAS) with inorganic bisulfide. The first half of the sulfhydrylase reaction, formation of the alpha-aminoacrylate intermediate, limits the overall reaction rate, while in the second half-reaction, with bisulfide as the substrate, chemistry is thought to be diffusion-limited. In order to characterize the second half-reaction, the pH dependence of the pseudo-first-order rate constant for disappearance of the alpha-aminoacrylate intermediate was measured over the pH range 6.0-9.5 using the natural substrate bisulfide, and a number of nucleophilic analogues. The rate is pH-dependent for substrates with a pK(a) > 7, while the rate constant is pH-independent for substrates with a pK(a) < 7 suggesting that the pK(a)s of the substrate and an enzyme group are important in this half of the reaction. In D(2)O, at low pD values, the amino acid external Schiff base is trapped, while in H(2)O the reaction proceeds through release of the amino acid product, which is likely rate-limiting for all nucleophilic reactants. A number of new beta-substituted amino acids were produced and characterized by (1)H NMR spectroscopy.  相似文献   

11.
The 3-hydroxybenzoate hydroxylase (MHBH) from Comamonas testosteroni KH122-3s is a single-component flavoprotein monooxygenase, a member of the glutathione reductase (GR) family. It catalyzes the conversion of 3-hydroxybenzoate to 3,4-dihydroxybenzoate with concomitant requirements for equimolar amounts of NADPH and molecular oxygen. The production of dihydroxy-benzenoid derivative by hydroxylation is the first step in the aerobic degradation of various phenolic compounds in soil microorganisms. To establish the structural basis for substrate recognition, the crystal structure of MHBH in complex with its substrate was determined at 1.8 A resolution. The enzyme is shown to form a physiologically active homodimer with crystallographic 2-fold symmetry, in which each subunit consists of the first two domains comprising an active site and the C-terminal domain involved in oligomerization. The protein fold of the catalytic domains and the active-site architecture, including the FAD and substrate-binding sites, are similar to those of 4-hydroxybenzoate hydroxylase (PHBH) and phenol hydroxylase (PHHY), which are members of the GR family, providing evidence that the flavoprotein aromatic hydroxylases share similar catalytic actions for hydroxylation of the respective substrates. Structural comparison of MHBH with the homologous enzymes suggested that a large tunnel connecting the substrate-binding pocket to the protein surface serves for substrate transport in this enzyme. The internal space of the large tunnel is distinctly divided into hydrophilic and hydrophobic regions. The characteristically stratified environment in the tunnel interior and the size of the entrance would allow the enzyme to select its substrate by amphiphilic nature and molecular size. In addition, the structure of the Xe-derivative at 2.5 A resolution led to the identification of a putative oxygen-binding site adjacent to the substrate-binding pocket. The hydrophobic nature of the xenon-binding site extends to the solvent through the tunnel, suggesting that the tunnel could be involved in oxygen transport.  相似文献   

12.
Glucose oxidase from the mold Aspergillus niger (EC 1.1.3.4) oxidizes beta-D-glucose with a wide variety of oxidizing substrates. The substrates were divided into three main groups: molecular oxygen, quinones, and one-electron acceptors. The kinetic and chemical mechanism of action for each group of substrates was examined in turn with a wide variety of kinetic methods and by means of molecular modeling of enzyme-substrate complexes. There are two proposed mechanisms for the reductive half-reaction: hydride abstraction and nucleophilic attack followed by deprotonation. The former mechanism appears plausible; here, beta-D-glucose is oxidized to glucono-delta-lactone by a concerted transfer of a proton from its C1-hydroxyl to a basic group on the enzyme (His516) and a direct hydride transfer from its C1 position to the N5 position in FAD. The oxidative half-reaction proceeds via one- or two-electron transfer mechanisms, depending on the type of the oxidizing substrate. The active site of the enzyme contains, in addition to FAD, three amino acid side chains that are intimately involved in catalysis: His516 with a pK(a)=6.9, and Glu412 with pK(a)=3.4 which is hydrogen bonded to His559, with pK(a)>8. The protonation of each of these residues has a strong influence on all rate constants in the catalytic mechanism.  相似文献   

13.
The ascomycetous yeast Candida parapsilosis CBS604 catabolizes 4-hydroxybenzoate through the initial formation of hydroquinone (1, 4-dihydroxybenzene). High levels of hydroquinone hydroxylase activity are induced when the yeast is grown on either 4-hydroxybenzoate, 2,4-dihydroxybenzoate, 1,3-dihydroxybenzene or 1, 4-dihydroxybenzene as the sole carbon source. The monooxygenase constitutes up to 5% of the total amount of protein and is purified to apparent homogeneity in three chromatographic steps. Hydroquinone hydroxylase from C. parapsilosis is a homodimer of about 150 kDa with each 76-kDa subunit containing a tightly noncovalently bound FAD. The flavin prosthetic group is quantitatively resolved from the protein at neutral pH in the presence of chaotropic salts. The apoenzyme is dimeric and readily reconstituted with FAD. Hydroquinone hydroxylase from C. parapsilosis catalyzes the ortho-hydroxylation of a wide range of monocyclic phenols with the stoichiometric consumption of NADPH and oxygen. With most aromatic substrates, no uncoupling of hydroxylation occurs. Hydroxylation of monofluorinated phenols is highly regiospecific with a preference for C6 hydroxylation. Binding of phenol highly stimulates the rate of flavin reduction by NADPH. At pH 7.6, 25 degrees C, this step does not limit the rate of overall catalysis. During purification, hydroquinone hydroxylase is susceptible towards limited proteolysis. Proteolytic cleavage does not influence the enzyme dimeric nature but results in relatively stable protein fragments of 55, 43, 35 and 22 kDa. N-Terminal peptide sequence analysis revealed the presence of two nick sites and showed that hydroquinone hydroxylase from C. parapsilosis is structurally related to phenol hydroxylase from Trichosporon cutaneum. The implications of these findings for the catalytic mechanism of hydroquinone hydroxylase are discussed.  相似文献   

14.
Angucyclines are aromatic polyketides produced in Streptomycetes via complex enzymatic biosynthetic pathways. PgaE and CabE from S. sp PGA64 and S. sp. H021 are two related homo-dimeric FAD and NADPH dependent aromatic hydroxylases involved in the early steps of the angucycline core modification. Here we report the three-dimensional structures of these two enzymes determined by X-ray crystallography using multiple anomalous diffraction and molecular replacement, respectively, to resolutions of 1.8 A and 2.7 A. The enzyme subunits are built up of three domains, a FAD binding domain, a domain involved in substrate binding and a C-terminal thioredoxin-like domain of unknown function. The structure analysis identifies PgaE and CabE as members of the para-hydroxybenzoate hydroxylase (pHBH) fold family of aromatic hydroxylases. In contrast to phenol hydroxylase and 3-hydroxybenzoate hydroxylase that utilize the C-terminal domain for dimer formation, this domain is not part of the subunit-subunit interface in PgaE and CabE. Instead, dimer assembly occurs through interactions of their FAD binding domains. FAD is bound non-covalently in the "in"-conformation. The active sites in the two enzymes differ significantly from those of other aromatic hydroxylases. The volumes of the active site are significantly larger, as expected in view of the voluminous tetracyclic angucycline substrates. The structures further suggest that substrate binding and catalysis may involve dynamic rearrangements of the middle domain relative to the other two domains. Site-directed mutagenesis studies of putative catalytic groups in the active site of PgaE argue against enzyme-catalyzed substrate deprotonation as a step in catalysis. This is in contrast to pHBH, where deprotonation/protonation of the substrate has been suggested as an essential part of the enzymatic mechanism.  相似文献   

15.
Phenol hydroxylase, an FAD-containing monooxygenase, catalyzes the conversion of substituted phenols to the corresponding catechol. Use of metapyrocatechase, capable of dioxygenation of several catechols to give highly absorbing products, permitted determination of the time course of product release from phenol hydroxylase. Product dissociated prior to complete reoxidation of the enzyme, most likely concomitant with formation of the 4a-hydroxyflavin species (intermediate III). Deuterated phenol and thiophenol exhibited no kinetic isotope effect during the oxidative half-reaction. Isotope effects of 1.7 to 3.7 were found with resorcinol for the conversion of the second intermediate to intermediate III. These effects limited the possible models for phenol hydroxylation. An attempt was made to distinguish whether the spectrum of intermediate II is due entirely to that of the flavin moiety of phenol hydroxylase or whether some radical intermediate form involved in the formation of catechol makes a significant visible contribution. Reduced native and 6-hydroxy-FAD phenol hydroxylase were reacted with oxygen and resorcinol in order to provide evidence for the identity of intermediate II.  相似文献   

16.
para-Hydroxybenzoate hydroxylase is a flavoprotein monooxygenase that catalyses a reaction in two parts: reduction of the flavin adenine dinucleotide (FAD) in the enzyme by reduced nicotinamide adenine dinucleotide phosphate (NADPH) in response to binding p-hydroxybenzoate to the enzyme and oxidation of reduced FAD with oxygen to form a hydroperoxide, which then oxygenates p-hydroxybenzoate. These different reactions are coordinated through conformational rearrangements of the protein and isoalloxazine ring during catalysis. Earlier research showed that reduction of FAD occurs when the isoalloxazine of the FAD moves to the surface of the protein to allow hydride transfer from NADPH. This move is coordinated with protein rearrangements that are triggered by deprotonation of buried p-hydroxybenzoate through a H-bond network that leads to the surface of the protein. In this paper, we examine the involvement of this same H-bond network in the oxygen reactions-the initial formation of a flavin-C4a-hydroperoxide from the reaction between oxygen and reduced flavin, the electrophilic attack of the hydroperoxide upon the substrate to form product, and the elimination of water from the flavin-C4a-hydroxide to form oxidized enzyme in association with product release. These reactions were measured through absorbance and fluorescence changes in the FAD during the reactions. Results were collected over a range of pH for the reactions of wild-type enzyme and a series of mutant enzymes with the natural substrate and substrate analogues. We discovered that the rate of formation of the flavin hydroperoxide is not influenced by pH change, which indicates that the proton required for this reaction does not come from the H-bond network. The rate of the hydroxylation reaction increases with pH in a manner consistent with a pK(a) of 7.1. We conclude that the H-bond network abstracts the phenolic proton from p-hydroxybenzoate in the transition state of oxygen transfer. The rate of formation of oxidized enzyme increases with pH in a manner consistent with a pK(a) of 7.1, indicating the involvement of the H-bond network. We conclude that product deprotonation enhances the rate of a specific conformational change required for both product release and the elimination of water from C4a-OH-FAD.  相似文献   

17.
Maize polyamine oxidase (MPAO) is a flavin adenine dinucleotide (FAD)-dependent enzyme that catalyses the oxidation of spermine and spermidine at the secondary amino groups. The structure of MPAO indicates a 30-A long U-shaped tunnel that forms the catalytic site, with residues Glu62 and Glu170 located close to the enzyme-bound FAD and residue Tyr298 in close proximity to Lys300, which in turn is hydrogen-bonded to the flavin N(5) atom via a water molecule (HOH309). To provide insight into the role of these residues in the catalytic mechanism of FAD reduction, we have performed steady-state and stopped-flow studies with wild-type, Glu62Gln, Glu170Gln, Tyr298Phe, and Lys300Met MPAO enzymes. We show that the steady-state enzyme activity is governed by an ionisable group with a macroscopic pK(a) of approximately 5.8. Kinetic analysis of the Glu62Gln, Glu170Gln, and Tyr298Phe MPAO enzymes have indicated (i) only small perturbations in catalytic activity as a result of mutation and (ii) steady-state pH profiles essentially unaltered when compared to the wild-type enzyme, suggesting that these residues do not play a critical role in the reaction mechanism. These kinetic observations are consistent with computational calculations that suggest that Glu62 and Glu170 are protonated over the pH range accessible to kinetic studies. Substitution of Lys300 with Met in MPAO resulted in a 1400-fold decrease in the rate of flavin reduction and a 160-fold decrease in the equilibrium dissociation constant for the Lys300Met-spermidine complex, consistent with a major role for this residue in the mechanism of substrate oxidation. A sizable solvent isotope effect (SIE = 5) accompanies FAD reduction in the wild-type enzyme and steady-state turnover (SIE = 2.3) of MPAO, consistent with the reductive half-reaction of MPAO making a major contribution to rate limitation in steady-state turnover. Studies using the enzyme-monitored turnover method indicate that oxidized FAD is the prominent form during steady-state turnover, consistent with the reductive half-reaction being rate-limiting. Our studies indicate the importance of Lys300 and probable importance of HOH309 to the mechanism of flavin reduction in MPAO. Possible roles for Lys300 and water in the mechanism of flavin reduction are discussed.  相似文献   

18.
p-Hydroxyphenylacetate (HPA) 3-hydroxylase (HPAH) catalyzes the hydroxylation of HPA at the ortho-position to yield 3,4-dihydroxyphenylacetate. The enzyme is a flavin-dependent two-component monooxygenase that consists of a reductase component and an oxygenase component (C(2)). C(2) catalyzes the hydroxylation of HPA using oxygen and reduced FMN as co-substrates. To date, the effects of pH on the oxygenation of the two-component monooxygenases have never been reported. Here, we report the reaction kinetics of C(2)·FMNH(-) with oxygen at various pH values investigated by stopped-flow and rapid quenched-flow techniques. In the absence of HPA, the rate constant for the formation of C4a-hydroperoxy-FMN (~1.1 × 10(6) m(-1)s(-1)) was unaffected at pH 6.2-9.9, which indicated that the pK(a) of the enzyme-bound reduced FMN was less than 6.2. The rate constant for the following H(2)O(2) elimination step increased with higher pH, which is consistent with a pK(a) of >9.4. In the presence of HPA, the rate constants for the formation of C4a-hydroperoxy-FMN (~4.8 × 10(4) m(-1)s(-1)) and the ensuing hydroxylation step (15-17 s(-1)) were not significantly affected by the pH. In contrast, the following steps of C4a-hydroxy-FMN dehydration to form oxidized FMN occurred through two pathways that were dependent on the pH of the reaction. One pathway, dominant at low pH, allowed the detection of a C4a-hydroxy-FMN intermediate, whereas the pathway dominant at high pH produced oxidized FMN without an apparent accumulation of the intermediate. However, both pathways efficiently catalyzed hydroxylation without generating significant amounts of wasteful H(2)O(2) at pH 6.2-9.9. The decreased accumulation of the intermediate at higher pH was due to the greater rates of C4a-hydroxy-FMN decay caused by the abolishment of substrate inhibition in the dehydration step at high pH.  相似文献   

19.
Xu D  Ballou DP  Massey V 《Biochemistry》2001,40(41):12369-12378
Three residues in the active site of the flavoprotein phenol hydroxylase (PHHY) were independently changed by site-directed mutagenesis. One of the mutant forms of PHHY, Tyr289Phe, is reduced by NADPH much slower than is the wild-type enzyme, although it has a slightly higher redox potential than the wild-type enzyme. In the structure of the wild-type enzyme, residue Tyr289 is hydrogen-bonded with the FAD when the latter is at the "out" position but has no direct contact with the flavin when it is "in". The oxidative half-reaction of PHHY is not significantly affected by this mutation, contrary to the concept that Tyr289 is a critical residue in the hydroxylation reaction [Enroth, C., Neujahr, H., Schneider, G., and Lindqvist, Y. (1998) Structure 6, 605-617; Ridder, L., Mullholland, A. J., Rietjens, I. M. C. M., and Vervoort, J. (2000) J. Am. Chem. Soc. 122, 8728-8738]. Tyr289 may help stabilize the FAD in the out conformation where it can be reduced by NADPH. For the Asp54Asn mutant form of PHHY, the initial step of the oxidative half-reaction is significantly slower than for the wild-type enzyme. Asp54Asn utilizes less than 20% of the reduced flavin for hydroxylating the substrate with the remainder forming H(2)O(2). Similar changes are observed when Arg281, a residue between Asp54 and the solvent, is mutated to Met. These two residues are suggested to be part of the active site environment the enzyme provides for the flavin cofactor to function optimally in the oxidative half-reaction. In the construction of the mutant forms of PHHY, it was determined that 11 of the previously reported amino acid residues in the sequence of PHHY were incorrect.  相似文献   

20.
Phenol hydroxylase that catalyzes the conversion of phenol to catechol in Rhodococcus erythropolis UPV-1 was identified as a two-component flavin-dependent monooxygenase. The two proteins are encoded by the genes pheA1 and pheA2, located very closely in the genome. The sequenced pheA1 gene was composed of 1,629 bp encoding a protein of 542 amino acids, whereas the pheA2 gene consisted of 570 bp encoding a protein of 189 amino acids. The deduced amino acid sequences of both genes showed high homology with several two-component aromatic hydroxylases. The genes were cloned separately in cells of Escherichia coli M15 as hexahistidine-tagged proteins, and the recombinant proteins His6PheA1 and His6PheA2 were purified and its catalytic activity characterized. His6PheA1 exists as a homotetramer of four identical subunits of 62 kDa that has no phenol hydroxylase activity on its own. His6PheA2 is a homodimeric flavin reductase, consisting of two identical subunits of 22 kDa, that uses NAD(P)H in order to reduce flavin adenine dinucleotide (FAD), according to a random sequential kinetic mechanism. The reductase activity was strongly inhibited by thiol-blocking reagents. The hydroxylation of phenol in vitro requires the presence of both His6PheA1 and His6PheA2 components, in addition to NADH and FAD, but the physical interaction between the proteins is not necessary for the reaction.  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