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.
Soluble methane monooxygenase (MMO) is shown to be capable of catalyzing desaturation reactions in addition to the usual hydroxylation and epoxidation reactions. Dehydrogenated products are generated from MMO-catalyzed oxidation of certain substrates including ethylbenzene and cyclohexadienes. In the reaction of ethylbenzene, desaturation of ethyl C-H occurred along with the conventional hydroxvlations of ethyl and phenyl C-Hs. As a result, styrene is formed together with ethylphenols and phenylethanols. Similarly, when 1,3- and 1,4-cyclohexadienes were used as substrates, benzene was detected as a product in addition to the corresponding alcohols and epoxides. In all cases, reaction conditions were found to significantly affect the distribution among the different products. This new activity of MMO is postulated to be associated with the chemical properties of the substrates rather than fundamental changes in the nature of the oxygen and C-H activation chemistries. The formation of the desaturated products is rationalized by formation of a substrate cationic intermediate, possibly via a radical precursor. The cationic species is then proposed to partition between recombination (alcohol formation) and elimination (alkene production) pathways. This novel function of MMO indicates close mechanistic kinship between the hydroxylation and desaturation reactions catalyzed by the nonheme diiron clusters.  相似文献   

2.
The soluble methane monooxygenase (MMO) system isolated from Methylosinus trichosporium OB3b catalyzes the adventitious oxidation of alkyl substituted methylcyclopropanes. If the chemical mechanism of C-H activation by MMO involves formation of a radical or carbocation intermediate at the methyl C-H of these 'radical clock' substrates, then cyclopropyl ring opened alcohols may appear in the product mixture due to rearrangement of the intermediate. The lifetime of radical intermediates can be determined from known rearrangement rate constants, k(r). Rearrangement was observed during the oxidation of 1,1,2,2-tetramethylcyclopropane (k(r)=1.7-17. 5x10(8) s(-1), 30 degrees C) but not for cis- or trans-1, 2-dimethylcyclopropane (k(r)=1.2-6.4x10(8) s(-1), 30 degrees C) or the very fast radical clock, trans-2-phenylmethylcyclopropane (k(r)=3.4x10(11) s(-1), 30 degrees C). The results show that the occurrence of rearranged products fails to correlate with either the chemical nature of the C-H bond being broken, which is very similar for all of the methylcyclopropanes studied here, or the magnitude of the radical k(r) value. This study suggests that the steric properties of the substrate play an important role in determining the outcome of the reaction. Substrates with bulky substituents near the C-H bond that is attacked appear to yield intermediates with sufficient lifetimes to rearrange. In contrast, substrates with less steric bulk are postulated to be able to approach the reactive oxygen species in the MMO active site more closely so that intermediates are either rapidly quenched or undergo subsequent interaction with the dinuclear iron cluster of MMO that prevents rearrangement.  相似文献   

3.
Alpha-thujone (1alpha) and beta-thujone (1beta) were used to investigate the mechanism of hydrocarbon hydroxylation by cytochromes P-450(cam) (CYP101) and P-450(BM3) (CYP102). The thujones are hydroxylated by these enzymes at various positions, but oxidation at C-4 gives rise to both rearranged and unrearranged hydroxylation products. Rearranged products result from the formation of a radical intermediate that can undergo either inversion of stereochemistry or ring opening of the adjacent cyclopropane ring. Both of these rearrangements, as well as a C-4 desaturation reaction, are observed. The ring opening clock gives oxygen rebound rates that range from 0.2 x 10(10) to 2.8 x 10(10) s(-1) for the different substrate and enzyme combinations. The C-4 inversion reaction provides independent confirmation of a radical intermediate. The phenol product expected if a C-4 cationic rather than radical intermediate is formed is not detected. The results are consistent with a two-state process and provide support for a radical rebound but not a hydroperoxide insertion mechanism for cytochrome P-450 hydroxylation.  相似文献   

4.
Hydrocarbon oxidations catalyzed by methane monooxygenase purified to high specific activity from the type II methanotroph Methylosinus trichosporium OB3b were compared to the same reactions catalyzed by methane monooxygenase from the type I methanotroph Methylococcus capsulatus Bath and liver microsomal cytochrome P-450. The two methane monooxygenases produced nearly identical product distributions, in accord with physical studies of the enzymes which have shown them to be very similar. The products obtained from the oxidation of a series of deuterated substrates by the M. trichosporium methane monooxygenase were very similar to those reported for the same reaction catalyzed by liver microsomal cytochrome P-450, suggesting that the enzymes use similar mechanisms. However, differences in the product distributions and other aspects of the reactions indicated the mechanisms are not identical. Methane monooxygenase epoxidized propene in D2O and d6-propene in H2O without exchange of substrate protons or deuterons with solvent, in contrast to cytochrome P-450 (Groves, J. T., Avaria-Neisser, G. E., Fish, K. M., Imachi, M., and Kuczkowski, R. L. (1986) J. Am. Chem. Soc. 108, 3837-3838), suggesting that the mechanism of epoxidation of olefins by methane monooxygenase differs at least in part from that of cytochrome P-450. Hydroxylation of alkanes by methane monooxygenase revealed close similarities to hydroxylations by cytochrome P-450. Allylic hydroxylation of 3,3,6,6-d4-cyclohexene occurred with approximately 20% allylic rearrangement in the case of methane monooxygenase, whereas 33% was reported for this reaction catalyzed by cytochrome P-450 (Groves, J. T., and Subramanian, D. V. (1984) J. Am. Chem. Soc. 106, 2177-2181). Similarly, hydroxylation of exo,exo,exo,exo-2,3,5,6-d4-norbornane by methane monooxygenase occurred with epimerization, but to a lesser extent than reported for cytochrome P-450 (Groves, J. T., McClusky, G. A., White, R. E., and Coon, M. J. (1978) Biochem. Biophys. Res. Commun. 81, 154-160). A large intramolecular isotope effect, kH,exo/kD,exo greater than or equal to 5.5, was calculated for this reaction. However, the intermolecular kinetic isotope effect on Vm for methane oxidation was small, suggesting that steps other than C-H bond breakage were rate limiting in the overall enzymatic reaction. Similar isotope effects have been observed for cytochrome P-450. These observations indicate a stepwise mechanism of hydroxylation for methane monooxygenase analogous to that proposed for cytochrome P-450.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

5.
Brazeau BJ  Lipscomb JD 《Biochemistry》2000,39(44):13503-13515
The transient kinetics of formation and decay of the reaction cycle intermediates of the Methylosinus trichosporium OB3b methane monooxygenase (MMO) catalytic cycle are studied as a function of temperature and substrate type and deuteration. Kinetic evidence is presented for the existence of three intermediates termed compounds O, P, and P forming after the addition of O(2) to diferrous MMO hydroxylase (H(r)) and before the formation of the reactive intermediate compound Q. The Arrhenius plots for these reactions are linear and independent of substrate concentration and type, showing that substrate does not participate directly in the oxygen activation phase of the catalytic cycle. Analysis of the transient kinetic data revealed only small changes relative to the weak optical spectrum of H(r) for any of these intermediates. In contrast, large changes in the 430 nm spectral region are associated with the formation of Q. The decay reaction of Q exhibits an apparent first-order concentration dependence for all substrates tested, and the observed rate constant depends on the substrate type. The kinetics of the decay reaction of Q yield a nonlinear Arrhenius plot when methane is the substrate, and the rates in both segments of the plot increase linearly with methane concentration. Together these observations suggest that at least two reactions with a methane concentration dependence, and perhaps two methane molecules, are involved in the decay process. When CD(4) is used as the substrate, a large isotope effect and a linear Arrhenius plot are observed. Analogous plots for all other MMO substrates tested (e.g., ethane) are linear, and no isotope effect for deuterated analogues is observed. This demonstrates that a step other than C-H bond breaking is rate limiting for alternative MMO substrates. A two step Q decay mechanism is proposed that provides an explanation for the lack of an isotope effect for alternative MMO substrates and the fact that rate of oxidation of methane by Q exceeds that of many other hydrocarbons with weaker C-H bonds.  相似文献   

6.
Cytochromes P450cam and P450BM3 oxidize alpha- and beta-thujone into multiple products, including 7-hydroxy-alpha-(or beta-)thujone, 7,8-dehydro-alpha-(or beta-)thujone, 4-hydroxy-alpha-(or beta-)thujone, 2-hydroxy-alpha-(or beta-)thujone, 5-hydroxy-5-isopropyl-2-methyl-2-cyclohexen-1-one, 4,10-dehydrothujone, and carvacrol. Quantitative analysis of the 4-hydroxylated isomers and the ring-opened product indicates that the hydroxylation proceeds via a radical mechanism with a radical recombination rate ranging from 0.7 +/- 0.3 x 10(10) s(-1) to 12.5 +/- 3 x 10(10) s(-1) for the trapping of the carbon radical by the iron-bound hydroxyl radical equivalent. 7-[2H]-alpha-Thujone has been synthesized and used to amplify C-4 hydroxylation in situations where uninformative C-7 hydroxylation is the dominant reaction. The involvement of a carbon radical intermediate is confirmed by the observation of inversion of stereochemistry of the methyl-substituted C-4 carbon during the hydroxylation. With an L244A mutation that slightly increases the P450(cam) active-site volume, this inversion is observed in up to 40% of the C-4 hydroxylated products. The oxidation of alpha-thujone by human CYP1A2, CYP2C9, CYP2C19, CYP2D6, CYP2E1, and CYP3A4 occurs with up to 80% C-4 methyl inversion, in agreement with a dominant radical hydroxylation mechanism. Three minor desaturation products are produced, with at least one of them via a cationic pathway. The cation involved is proposed to form by electron abstraction from a radical intermediate. The absence of a solvent deuterium isotope effect on product distribution in the P450cam reaction precludes a significant role for the P450 ferric hydroperoxide intermediate in substrate hydroxylation. The results indicate that carbon hydroxylation is catalyzed exclusively by a P450 ferryl species via radical intermediates whose detailed properties are substrate- and enzyme-dependent.  相似文献   

7.
EPR spin-trapping experiments were carried out using the three-component soluble methane monooxygenase (MMO). Spin-traps 5,5-dimethyl-1-pyrroline N-oxide (DMPO), alpha-4-pyridyl-1-oxide N-tert-butylnitrone (POBN), and nitrosobenzene (NOB) were used to investigate the possible formation of substrate radical intermediates during catalysis. In contrast to a previous report, the NADH-coupled oxidations of various substrates did not produce any trapped radical species when DMPO or POBN was present. However, radicals were detected by these traps when only the MMO reductase component and NADH were present. DMPO and POBN were found to be weak inhibitors of the MMO reaction. In contrast, NOB is a strong inhibitor for the MMO-catalyzed nitrobenzene oxidation reaction. When NOB was used as a spin-trap in the complete MMO system with or without substrate, EPR signals from an NOB radical were detected. We propose that a molecule of NOB acts simultaneously as a substrate and a spin-trap for MMO, yielding the long-lived radical and supporting a stepwise mechanism for MMO.  相似文献   

8.
The diagnostic substrate tetramethylcyclopropane (TMCP) has been reexamined as a substrate with three drug- and xenobiotic-metabolizing cytochrome P450 enzymes, human CYP2E1, CYP3A4 and rat CYP2B1. The major hydroxylation product in all cases was the unrearranged primary alcohol along with smaller amounts of a rearranged tertiary alcohol. Significantly, another ring-opened product, diacetone alcohol, was also observed. With CYP2E1 this product accounted for 20% of the total turnover. Diacetone alcohol also was detected as a product from TMCP with a biomimetic model catalyst, FeTMPyP, but not with a ruthenium porphyrin catalyst. Lifetimes of the intermediate radicals were determined from the ratios of rearranged and unrearranged products to be 120, 13 and 1 ps for CYP2E1, CYP3A4 and CYP2B1, respectively, corresponding to rebound rates of 0.9 × 1010 s−1, 7.2 × 1010 s−1 and 1.0 × 1012 s−1. For the model iron porphyrin, FeTMPyP, a radical lifetime of 81 ps and a rebound rate of 1.2 × 1010 s−1 were determined. These apparent radical lifetimes are consistent with earlier reports with a variety of CYP enzymes and radical clock substrates, however, the large amounts of diacetone alcohol with CYP2E1 and the iron porphyrin suggest that for these systems a considerable amount of the intermediate carbon radical is trapped by molecular oxygen. These results add to the view that cage escape of the intermediate carbon radical in [FeIV–OH R] can compete with cage collapse to form a C–O bond. The results could be significant with regard to our understanding of iron-catalyzed C–H hydroxylation, the observation of P450-dependent peroxidation and the development of oxidative stress, especially for CYP2E1.  相似文献   

9.
In the presence of phenylalanine and molecular oxygen, activated phenylalanine hydroxylase catalyzes the oxidation of tetrahydrobiopterin. The oxidation of this tetrahydropterin cofactor also proceeds if the substrate, phenylalanine, is replaced by its product, tyrosine, in the initial reaction mixture. These two reactions have been defined as coupled and uncoupled, respectively, because in the former reaction 1 mol of phenylalanine is hydroxylated for every mole of tetrahydrobiopterin oxidized, whereas in the latter reaction there is no net hydroxylation of tyrosine during the oxidation of the tetrahydropterin. During the course of the coupled oxidation of tetrahydrobiopterin, a pterin 4a-carbinolamine intermediate can be detected by ultraviolet spectroscopy (Kaufman, S. (1976) in Iron and Copper Proteins (Yasunobu, K. T., Mower, H. F., and Hayaishi, O., eds) pp. 91-102, Plenum Publishing Corp., New York). Dix and Benkovic (Dix, T. A., and Benkovic, S. J. (1985) Biochemistry 24, 5839-5846) have postulated that the formation of this intermediate only occurs when the oxidation of the tetrahydropteridine is tightly coupled to the concomitant hydroxylation of the aromatic amino acid. However, during the tyrosine-dependent uncoupled oxidation of tetrahydrobiopterin by phenylalanine hydroxylase, we have detected the formation of a spectral intermediate with ultraviolet absorbance that is essentially identical to that of the carbinolamine. Furthermore, this absorbance can be eliminated by the addition of 4a-carbinolamine dehydratase, an enzyme which catalyzes the dehydration of the 4a-carbinolamine. Quantitation of this intermediate suggests that there are two pathways for the tyrosine-dependent uncoupled oxidation of tetrahydrobiopterin by phenylalanine hydroxylase because only about 0.3 mol of the intermediate is formed per mol of the cofactor oxidized.  相似文献   

10.
Myoglobin (Mb) is used as a model system for other heme proteins and the reactions they catalyze. The latest novel function to be proposed for myoglobin is a P450 type hydroxylation activity of aromatic carbons (Watanabe, Y., and Ueno, T. (2003) Bull. Chem. Soc. Jpn. 76, 1309-1322). Because Mb does not contain a specific substrate binding site for aromatic compounds near the heme, an engineered tryptophan in the heme pocket was used to model P450 hydroxylation of aromatic compounds. The monooxygenation product was not previously isolated because of rapid subsequent oxidation steps (Hara, I., Ueno, T., Ozaki, S., Itoh, S., Lee, K., Ueyama, N., and Watanabe, Y. (2001) J. Biol. Chem. 276, 36067-36070). In this work, a Mb variant (F43W/H64D/V68I) is used to characterize the monooxygenated intermediate. A modified (+16 Da) species forms upon the addition of 1 eq of H2O2. This product was digested with chymotrypsin, and the modified peptide fragments were isolated and characterized as 6-hydroxytryptophan using matrix-assisted laser desorption ionization time-of-flight tandem mass spectroscopy and 1H NMR. This engineered Mb variant represents the first enzyme to preferentially hydroxylate the indole side chain of Trp at the C6 position. Finally, heme extraction was used to demonstrate that both the formation of the 6-hydroxytryptophan intermediate (+16 Da) and subsequent oxidation to form the +30 Da final product are catalyzed by the heme cofactor, most probably via the compound I intermediate. These results provide insight into the mechanism of hydroxylation of aromatic carbons by heme proteins, demonstrating that non-thiolate-ligated heme enzymes can perform this function. This establishes Mb compound I as a model for P450 type aromatic hydroxylation chemistry.  相似文献   

11.
A mechanism for beta-chlorophenethylamine inhibition of dopamine beta-monooxygenase has been postulated in which enzyme-bound alpha-aminoacetophenone is generated, followed by an intramolecular redox reaction to yield a ketone-derived radical cation as the enzyme inhibitory species (Mangold, J. B., and Klinman, J. P. (1984) J. Biol. Chem. 259, 7772-7779). If correct, additional compounds capable of producing enzyme-bound (formula; see text) reductant should inhibit dopamine beta-monooxygenase. Phenylacetaldehyde was chosen to test this model, since beta-hydroxyphenylacetaldehyde is expected to function as a reductant in a manner analogous to alpha-aminoacetophenone. Phenylacetaldehyde exhibits the properties of a mechanism-based inhibitor. Kinetic parameters are comparable to beta-chlorophenethylamine under both initial velocity and inactivation conditions. Since phenylacetaldehyde bears little resemblance to beta-chlorophenethylamine, its analogous inhibitory action provides support for an intramolecular redox reaction (via beta-hydroxyphenylacetaldehyde oxidation to a radical cation) in dopamine beta-monooxygenase inactivation. beta-Hydroxyphenylacetaldehyde was identified as the enzymatic product of phenylacetaldehyde turnover. As predicted, this product behaves both as a time-dependent inhibitor of dopamine beta-monooxygenase and as an electron donor in enzyme-catalyzed hydroxylation of tyramine to octopamine. Phenylacetamide and p-hydroxyphenylacetamide are also found to be mechanism-based inhibitors of dopamine beta-monooxygenase. In this case the product of hydroxylation (beta-hydroxyphenylacetamide) is redox inactive and, therefore, is unable to function as either a reductant or an inhibitor. Thus, mechanism-based inhibitors are divided into two types: type I, which undergoes hydroxylation prior to inactivation, and type II, which only requires hydrogen atom abstraction. A general mechanism for dopamine beta-monooxygenase inactivation is described, in which a common mechanistic radical intermediate is formed from both pathways.  相似文献   

12.
Wallar BJ  Lipscomb JD 《Biochemistry》2001,40(7):2220-2233
Component interactions play important roles in the regulation of catalysis by methane monooxygenase (MMO). The binding of component B (MMOB) to the hydroxylase component (MMOH) has been shown in previous studies to cause structural changes in MMOH that result in altered thermodynamic and kinetic properties during the reduction and oxygen binding steps of the catalytic cycle. Here, specific amino acid residues of MMOB that play important roles in the interconversion of several intermediates of the MMO cycle have been identified. Both of the histidine residues in Methylosinus trichosporium OB3b MMOB (H5 and H33) were chemically modified by diethylpyrocarbonate (DEPC). Although the DEPC--MMOB species exhibited only minor changes relative to unmodified MMOB in steady-state MMO turnover, large decreases in the formation rate constants of the reaction cycle intermediates, compound P and compound Q, were observed. The site specific mutants H5A, H33A, and H5A/H33A were made and characterized. H5A and wild type MMOB elicited similar steady-state and transient kinetics, although the mutant caused a slightly lower rate constant for Q formation. Conversely, H33A exhibited a >50-fold decrease in the P formation rate constant, which resulted in slower formation of Q. The kinetics of the double mutant (H5A/H33A) were similar to those of H33A, suggesting that the highly conserved residue, H33, has the most significant effect on the efficient progress of the cycle. Ongoing NMR investigations of residues perturbed by formation of the MMOH-MMOB complex suggested construction of the MMOB N107G/S109A/S110A/T111A quadruple mutant. This mutant was found to elicit a nearly 2-fold increase in specific activity for steady-state MMO turnover of large substrates such as furan and nitrobenzene but caused no similar increase for the physiological substrate, methane. While the quadruple mutant did not have a significant effect on P and Q formation, it caused an almost 3-fold increase in the decay rate constant of Q for furan oxidation and a 2-fold faster product release rate constant for p-nitrophenol resulting from nitrobenzene oxidation. Conversely, this mutant caused the Q decay rate constant to decrease 7-fold for methane oxidation but left the product release step unaffected. These results show for the first time that MMOB exerts influence at late as well as early steps in the catalytic cycle. They also suggest that MMOB plays a critical role in determining the ability of MMO to distinguish between methane and larger substrates.  相似文献   

13.
Xu D  Enroth C  Lindqvist Y  Ballou DP  Massey V 《Biochemistry》2002,41(46):13627-13636
An active site residue in phenol hydroxylase (PHHY), Pro364, was mutated to serine to investigate its role in enzymatic catalysis. In the presence of phenol, the reaction between the reduced flavin of P364S and oxygen is very fast, but only 13% of the flavin is utilized to hydroxylate the substrate, compared to nearly 100% for the wild-type enzyme. The oxidative half-reaction of PHHY using m-cresol as a substrate is similarly affected by the mutation. Pro364 was suggested to be important in stabilizing the transition state of the oxygen transfer step by forming a hydrogen bond between its carbonyl oxygen and the C4a-hydroperoxyflavin [Ridder, L., Mullholland, A. J., Rietjens, I. M. C. M., and Vervoort, J. (2000) J. Am. Chem. Soc. 122, 8728-8738]. The P364S mutation may weaken this interaction by increasing the flexibility of the peptide chain; hence, the transition state would be destabilized to result in a decreased level of hydroxylation of phenol. However, when the oxidative half-reaction was studied using resorcinol as a substrate, the P364S mutant form was not significantly different from the wild-type enzyme. The rate constants for all the reaction steps as well as the hydroxylation efficiency (coupling between NADPH oxidation and resorcinol consumption) are comparable to those of the wild-type enzyme. It is suggested that the function of Pro364 in catalysis, stabilization of the transition state, is not as important in the reaction with resorcinol, possibly because the position of hydroxylation is different with resorcinol than with phenol and m-cresol.  相似文献   

14.
Dey M  Li X  Kunz RC  Ragsdale SW 《Biochemistry》2010,49(51):10902-10911
Methyl-coenzyme M reductase (MCR) from methanogenic archaea catalyzes the terminal step in methanogenesis using coenzyme B (CoBSH) as the two-electron donor to reduce methyl-coenzyme M (methyl-SCoM) to form methane and the heterodisulfide, CoBS-SCoM. The active site of MCR contains an essential redox-active nickel tetrapyrrole cofactor, coenzyme F(430), which is active in the Ni(I) state (MCR(red1)). Several catalytic mechanisms have been proposed for methane synthesis that mainly differ in whether an organometallic methyl-Ni(III) or a methyl radical is the first catalytic intermediate. A mechanism was recently proposed in which methyl-Ni(III) undergoes homolysis to generate a methyl radical (Li, X., Telser, J., Kunz, R. C., Hoffman, B. M., Gerfen, G., and Ragsdale, S. W. (2010) Biochemistry 49, 6866-6876). Discrimination among these mechanisms requires identification of the proposed intermediates, none of which have been observed with native substrates. Apparently, intermediates form and decay too rapidly to accumulate to detectible amounts during the reaction between methyl-SCoM and CoBSH. Here, we describe the reaction of methyl-SCoM with a substrate analogue (CoB(6)SH) in which the seven-carbon heptanoyl moiety of CoBSH has been replaced with a hexanoyl group. When MCR(red1) is reacted with methyl-SCoM and CoB(6)SH, methanogenesis occurs 1000-fold more slowly than with CoBSH. By transient kinetic methods, we observe decay of the active Ni(I) state coupled to formation and subsequent decay of alkyl-Ni(III) and organic radical intermediates at catalytically competent rates. The kinetic data also revealed substrate-triggered conformational changes in active Ni(I)-MCR(red1). Electron paramagnetic resonance (EPR) studies coupled with isotope labeling experiments demonstrate that the radical intermediate is not tyrosine-based. These observations provide support for a mechanism for MCR that involves methyl-Ni(III) and an organic radical as catalytic intermediates. Thus, the present study provides important mechanistic insights into the mechanism of this key enzyme that is central to biological methane formation.  相似文献   

15.
A mechanism for beta-chlorophenethylamine inhibition of dopamine beta-monooxygenase has been postulated in which bound alpha-aminoacetophenone is generated followed by an intramolecular redox reaction to yield a ketone-derived radical cation as the inhibitory species (Mangold, J.B., and Klinman, J.P. (1984) J. Biol. Chem. 259, 7772-7779). Based on the assumption that the ketone radical is the inhibitory intermediate, an analogous system was predicted and verified (Bossard, M.J., and Klinman, J.P. (1986) J. Biol. Chem. 261, 16421-16427). In the present study, the role of alpha-aminoacetophenone as the proposed intermediate in the inactivation by beta-chlorophenethylamine was examined in greater detail. From the interdependence of tyramine and alpha-aminoacetophenone concentrations, ketone inactivation is concluded to occur at the substrate site as opposed to potential binding at the reductant-binding site. Using beta-[2-1H]- and beta-[2-2H]chlorophenethylamine, the magnitude of the deuterium isotope effect on inactivation under second-order conditions has been found to be identical to that observed under catalytic turnover, D(kappa inact/Ki) = D(kappa cat/Km) = 6-7. By contrast, the isotope effect on inactivation under conditions of substrate and oxygen saturation, D kappa inact = 2, is 3-fold smaller than that seen on catalytic turnover, D kappa cat = 6. This reduced isotope effect for inactivation is attributed to a normal isotope effect on substrate hydroxylation followed by an inverse isotope effect on the partitioning of the enol of alpha-aminoacetophenone between oxidation to a radical cation versus protonation to regenerate ketone. These findings are unusual in that two isotopically sensitive steps are present in the inactivation pathway whereas only one is observable in turnover.  相似文献   

16.
Methane gas is produced from many natural and anthropogenic sources. As such, methane gas plays a significant role in the Earth's climate, being 25 times more effective as a greenhouse gas than carbon dioxide. As with nearly all other naturally produced organic molecules on Earth, there are also micro-organisms capable of using methane as their sole source of carbon and energy. The microbes responsible (methanotrophs) are ubiquitous and, for the most part, aerobic. Although anaerobic methanotrophs are believed to exist, so far, none have been isolated in pure culture. Methanotrophs have been known to exist for over 100 years; however, it is only in the last 30 years that we have begun to understand their physiology and biochemistry. Their unique ability to use methane for growth is attributed to the presence of a multicomponent enzyme system-methane monooxygenase (MMO)-which has two distinct forms: soluble (sMMO) and membrane-associated (pMMO); however, both convert methane into the readily assimilable product, methanol. Our understanding of how bacteria are capable of effecting one of the most difficult reactions in chemistry-namely, the controlled oxidation of methane to methanol-has been made possible by the isolation, in pure form, of the enzyme components.The mechanism by which methane is activated by sMMO involves abstraction of a hydrogen atom from methane by a high-valence iron species (FeIV or possibly FeV) in the hydroxylase component of the MMO complex to form a methyl radical. The radical combines with a captive oxygen atom from dioxygen to form the reaction product, methanol, which is further metabolized by the cell to produce multicarbon intermediates. Regulation of the sMMO system relies on the remarkable properties of an effector protein, protein B. This protein is capable of facilitating component interactions in the presence of substrate, modifying the redox potential of the diiron species at the active site. These interactions permit access of substrates to the hydroxylase, coupling electron transfer by the reductase with substrate oxidation and affecting the rate and regioselectivity of the overall reaction. The membrane-associated form is less well researched than the soluble enzyme, but is known to contain copper at the active site and probably iron.From an applied perspective, methanotrophs have enjoyed variable successes. Whole cells have been used as a source of single-cell protein (SCP) since the 1970s, and although most plants have been mothballed, there is still one currently in production. Our earlier observations that sMMO was capable of inserting an oxygen atom from dioxygen into a wide variety of hydrocarbon (and some non-hydrocarbon) substrates has been exploited to either produce value added products (e.g. epoxypropane from propene), or in the bioremediation of pollutants such as chlorinated hydrocarbons. Because we have shown that it is now possible to drive the reaction using electricity instead of expensive chemicals, there is promise that the system could be exploited as a sensor for any of the substrates of the enzyme.  相似文献   

17.
A new concerted mechanism is proposed for the conversion of methane to methanol on intermediate Q of soluble methane monooxygenase (sMMO), the active site of which is considered to involve an Fe2(mu-O)2 diamond core. A hybrid density functional theory (DFT) method is used for our mechanistic study on the important reactivity of the bare FeO+ complex and a diiron model of intermediate Q. The reaction pathway for the methane hydroxylation on the diiron complex is essentially identical to that for the gas-phase reaction by the bare FeO+ complex. Methane is highly activated on the dinuclear iron model through the formation of a methane complex, in which a coordinatively unsaturated iron plays a central role in the bonding interaction between the diiron model and substrate methane. A H atom abstraction via a four-centered transition state and a recombination of the OH and CH3 groups via a three-centered transition state successively occur on the dinuclear iron-oxo species, leading to the formation of a methanol complex that corresponds to intermediate T. These electronic processes take place in a concerted manner. Our mechanism for methane hydroxylation by sMMO is different from the radical mechanism that has been widely accepted for enzymatic hydrocarbon hydroxylation, especially by cytochrome P450.  相似文献   

18.
The stereochemical course of the aliphatic hydroxylation of gamma-butyrobetaine by calf liver and by Pseudomonas sp AK1 gamma-butyrobetaine hydroxylases has been determined. With [3(RS)-3-3H]-gamma-butyrobetaine or [3(R)-3-3H]-gamma-butyrobetaine as substrate, a rapid and significant loss of tritium to the medium occurred. On the other hand, with [3(S)-3-3H]-gamma-butyrobetaine, only a negligible release of tritium to the aqueous medium was observed. Indeed, on hydroxylation of [3(S)-3-2H]-gamma-butyrobetaine by either the calf liver or bacterial hydroxylase, the isolated product L-carnitine was found to have retained all of the deuterium initially present in the 3(S) position. Since the absolute configuration of the product L-carnitine has been determined to be R, such results are only compatible with a hydroxylation reaction that proceeded with retention of configuration. With [methyl-14C,3(R)-3-3H]-gamma-butyrobetaine as substrate for the calf liver hydroxylase, the percentage of tritium retained in the [methyl-14C]-L-carnitine product was determined as a function of percent reaction. The results of these studies indicated that pro-R hydrogen atom abstraction exceeded 99.9%. Experiments using racemic [methyl-14C,3(RS)-3-3H]-gamma-butyrobetaine as substrate yielded similar results and additionally allowed us to estimate alpha-secondary tritium kinetic isotope effects of 1.10 and 1.31 for the bacterial and calf liver enzymes, respectively. These results are discussed within the context of the radical mechanism for gamma-butyrobetaine hydroxylase previously proposed [Blanchard, J. S., & Englard, S. (1983) Biochemistry 22, 5922], and the required topographical arrangement of enzymic oxidant and substrate is illustrated.  相似文献   

19.
Cytochrome P450 (P450) 2D6 was first identified as the polymorphic human debrisoquine hydroxylase and subsequently shown to catalyze the oxidation of a variety of drugs containing a basic nitrogen. Residue Asp301 has been characterized as being involved in electrostatic interactions with substrates on the basis of homology modeling and site-directed mutagenesis experiments [Ellis, S. W., Hayhurst, G. P., Smith, G., Lightfoot, T., Wong, M. M. S., Simula, A. P., Ackland, M. J., Sternberg, M. J. E., Lennard, M. S., Tucker, G. T., and Wolf, C. R. (1995) J. Biol. Chem. 270, 29055-29058]. However, pharmacophore models based on the role of Asp301 in substrate binding are compromised by reports of catalytic activity toward substrates devoid of a basic nitrogen, which have generally been ignored. We characterized a high-affinity ligand for P450 2D6, also devoid of a basic nitrogen atom, spirosulfonamide [4-[3-(4-fluorophenyl)-2-oxo-1-oxaspiro[4.4]non-3-en-4-yl]benzenesulfonamide], with K(s) 1.6 microM. Spirosulfonamide is a substrate for P450 2D6 (k(cat) 6.5 min(-)(1) for the formation of a syn spiromethylene carbinol, K(m) 7 microM). Mutation of Asp301 to neutral residues (Asn, Ser, Gly) did not substantially affect the binding of spirosulfonamide (K(s) 2.5-3.5 microM). However, the hydroxylation of spirosulfonamide was attenuated in these mutants to the same extent (90%) as for the classic nitrogenous substrate bufuralol, and the effect of the D301N substitution was manifested on k(cat) but not K(m). Analogues of spirosulfonamide were also evaluated as ligands and substrates. Analogues in which the sulfonamide moiety was modified to an amide, thioamide, methyl sulfone, or hydrogen were ligands with K(s) values of 1.7-32 microM. All were substrates, and the methyl sulfone analogue was oxidized to the syn spiromethylene carbinol analogue of the major spirosulfonamide product. The D301N mutation produced varying changes in the oxidation patterns of the spirosulfonamide analogues. The peptidometic ritonavir and the steroids progesterone and testosterone had been reported to be substrates for P450 2D6, but the affinities (K(s)) were unknown; these were estimated to be 1.2, 1.5, and 15 microM, respectively (cf. 6 microM for the classic substrate bufuralol). The results are consistent with a role of Asp301 other than electrostatic interaction with a positively charged ligand. H-Bonding or electrostatic interactions probably enhance binding of some substrates, but our results show that it is not required for all substrates and explain why predictive models fail to recognize the proclivity for many substrates, especially those containing no basic nitrogen.  相似文献   

20.
K Wimalasena  K R Alliston 《Biochemistry》1999,38(45):14916-14926
Previous studies have shown that the dopamine beta-monooxygenase (DbetaM; E.C. 1.14.17.1)/1-(2-aminoethyl)-1,4-cyclohexadiene (CHDEA) reaction partitions between side chain and ring H-abstraction to produce the side-chain-hydroxylated product, 2-amino-1-(1, 4-cyclohexadienyl)ethanol, and the aromatized product, phenylethylamine, and that the two pathways do not crossover. [Wimalasena, K., and May, S. W. (1989) J. Am. Chem. Soc. 111, 2729-2731; Wimalasena, K., and Alliston, K. R. (1995) J. Am. Chem. Soc. 117, 1220-1224]. We now report that the ring H-abstraction pathway of the reaction further partitions to produce the ring hydroxylated product, CHDEA-6OH, and the aromatized product, PEA, at the carbon-oxygen bond formation step. The ring hydroxylation is shown to be stereospecific, exclusively producing the (S) product. The absolute stereospecificity of the ring and side-chain hydroxylations of the DbetaM/CHDEA reaction suggests that the side-chain pro-R hydrogen of the enzyme-bound substrate is close to perpendicular to the aromatic ring of the phenylethylamine substrate or cyclohexadiene ring of CHDEA. The relative activation energy parameters suggest that the partitioning of the ring H abstraction pathway between aromatized and ring hydroxylated products is due to the partitioning of the high-energy intermediates, the cyclohexadienyl radical and the Cu(II)-O(*) species, between carbon-oxygen bond formation and direct electron transfer. The relatively high activation enthalpic favorability and entropic unfavorability for the carbon-oxygen bond formation strongly suggest that the critical balancing of these two opposing forces is mandatory for the desired product formation.  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