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.
We propose a new DFT-based mechanism for methane production using the full F430 cofactor of MCR (methyl-coenzyme M reductase) along with a coordinated O=CH2CH2C(H)NH2C(H)O (surrogate for glutamine) as a model of the active site for conversion of CH3SCoM(-) (CH3SCH2CH2SO3(-)) + HSCoB to methane plus the corresponding heterodisulfide. The cycle begins with the protonation of F430, either on Ni or on the C-ring nitrogen of the tetrapyrrole ring, both of which are nearly equally favorable. The C-ring protonated form is predicted to oxidatively add CH3SCoM(-) to give a 4-coordinate Ni center where the Ni moves out of the plane of the four ring nitrogens. The movement of Ni (and the attached CH3 and SCH2CH2SO3(2-) ligands) toward the SCoB(-) (deprotonated HSCoB) cofactor allows a 2c-3e interaction to form between the two sulfur atoms. The release of the heterodisulfide yields a Ni(III) center with a methyl group attached. The concerted elimination of methane, where the methyl group coordinated to Ni abstracts the proton from the C-ring nitrogen, has a very small calculated activation barrier (5.4 kcal/mol). The NPA charge on Ni for the various reaction steps indicates that the oxidation state changes occur largely on the attached ligands.  相似文献   

2.
Methyl-coenzyme M reductase (MCR) from Methanothermobacter marburgensis (Mtm), catalyses the final step in methane synthesis in all methanogenic organisms. Methane is produced by coenzyme B-dependent two-electron reduction of methyl-coenzyme M. At the active site of MCR is the corphin cofactor F(430), which provides four-coordination through the pyrrole nitrogens to a central Ni ion in all states of the enzyme. The important MCRox1 ("ready") and MCRred1 ("active") states contain six-coordinate Ni(I) and differ in their upper axial ligands; furthermore, red1 appears to be two-electrons more reduced than in ox1 and other Ni(II) states that have been studied. On the basis of the reactivity of MCRred1 and MCRox1 with a substrate analogue and inhibitor (3-bromopropanesulfonate) and other small molecules (chloroform, dichloromethane, mercaptoethanol, and nitric oxide), we present evidence that the six-coordinate Ni(I) centers in the MCRred1 and MCRox1 states exhibit markedly different inherent reactivities. MCRred1 reacts faster with chloroform (2100-fold or 35000-fold when corrected for temperature effects), nitric oxide (90-fold), and 3-bromopropanesulfonate (10(6)-fold) than MCRox1. MCRred1 reacts with chloroform and dichloromethane and, like F(430), can catalyze dehalogenation reactions and produce lower halogenated products. We conclude that the enhanced reactivity of MCRred1 is due to the replacement of a relatively exchange-inert thiol ligand in MCRox1 with a weakly coordinating upper axial ligand in red1 that can be easily replaced by incoming ligands.  相似文献   

3.
Magnetic circular dichroism (MCD) spectroscopy and variable-temperature variable-field MCD are used in combination with density functional theory (DFT) and time-dependent DFT (TD-DFT) calculations to characterize the so-called ox1-silent, red1, and ox1 forms of the Ni-containing cofactor F430 in methyl-coenzyme M reductase (MCR). Previous studies concluded that the ox1 state, which is the precursor of the key reactive red1 state of MCR, is a Ni(I) species that derives from one-electron reduction of the Ni(II)-containing ox1-silent state. However, our absorption and MCD data provide compelling evidence that ox1 is actually a Ni(II) species. In support of this proposal, our DFT and TD-DFT calculations indicate that addition of an electron to the ox1-silent state leads to formation of a hydrocorphin anion radical rather than a Ni(I) center. These results and biochemical evidence suggest that ox1 is more oxidized than red1, which prompted us to test a new model for ox1 in which the ox1-silent species is oxidized by one electron to form a thiyl radical derived from coenzyme M that couples antiferromagnetically to the Ni(II) ion. This alternative ox1 model, formally corresponding to a Ni(III)/thiolate resonance form but with predicted S = 1/2 EPR parameters reminiscent of a Ni(I) (3dx2-y2)1 species, rationalizes the requirement for reduction of ox1 to yield the red1 species and the seemingly incongruent EPR and electronic spectra of the ox1 state.  相似文献   

4.
We present the 1.2 ? resolution X-ray crystal structure of a Ni-methyl species that is a proposed catalytic intermediate in methyl-coenzyme M reductase (MCR), the enzyme that catalyzes the biological formation of methane. The methyl group is situated 2.1 ? proximal of the Ni atom of the MCR coenzyme F(430). A rearrangement of the substrate channel has been posited to bring together substrate species, but Ni(III)-methyl formation alone does not lead to any observable structural changes in the channel.  相似文献   

5.
6.
7.
Methyl-coenzyme M reductase (MCR) catalyzes the reaction of methyl-coenzyme M (CH3-S-CoM) with coenzyme B (HS-CoB) to methane and CoM-S-S-CoB. At the active site, it contains the nickel porphinoid F430, which has to be in the Ni(I) oxidation state for the enzyme to be active. How the substrates interact with the active site Ni(I) has remained elusive. We report here that coenzyme M (HS-CoM), which is a reversible competitive inhibitor to methyl-coenzyme M, interacts with its thiol group with the Ni(I) and that for interaction the simultaneous presence of coenzyme B is required. The evidence is based on X-band continuous wave EPR and Q-band hyperfine sublevel correlation spectroscopy of MCR in the red2 state induced with 33S-labeled coenzyme M and unlabeled coenzyme B.  相似文献   

8.
Methyl-coenzyme M reductase (MCR) catalyzes methane formation from methyl-coenzyme M (methyl-SCoM) and N-7-mercaptoheptanoylthreonine phosphate (CoBSH). MCR contains a nickel hydrocorphin cofactor at its active site, called cofactor F(430). Here we present evidence that the macrocyclic ligand participates in the redox chemistry involved in catalysis. The active form of MCR, the red1 state, is generated by reducing another spectroscopically distinct form called ox1 with titanium(III) citrate. Previous electron paramagnetic resonance (EPR) and (14)N electron nuclear double resonance (ENDOR) studies indicate that both the ox1 and red1 states are best described as formally Ni(I) species on the basis of the character of the orbital containing the spin in the two EPR-active species. Herein, X-ray absorption spectroscopic (XAS) and resonance Raman (RR) studies are reported for the inactive (EPR-silent) forms and the red1 and ox1 states of MCR. RR spectra are also reported for isolated cofactor F(430) in the reduced, resting, and oxidized states; selected RR data are reported for the (15)N and (64)Ni isotopomers of the cofactor, both in the intact enzyme and in solution. Small Ni K-edge energy shifts indicate that minimal electron density changes occur at the Ni center during redox cycling of the enzyme. Titrations with Ti(III) indicate a 3-electron reduction of free cofactor F(430) to generate a stable Ni(I) state and a 2-electron reduction of Ni(I)-ox1 to Ni(I)-red1. Analyses of the XANES and EXAFS data reveal that both the ox1 and red1 forms are best described as hexacoordinate and that the main difference between ox1 and red1 is the absence of an axial thiolate ligand in the red1 state. The RR data indicate that cofactor F(430) undergoes a significant conformational change when it binds to MCR. Furthermore, the vibrational characteristics of the ox1 state and red1 states are significantly different, especially in hydrocorphin ring modes with appreciable C=N stretching character. It is proposed that these differences arise from a 2-electron reduction of the hydrocorphin ring upon conversion to the red1 form. Presumably, the ring-reduction and ligand-exchange reactions reported herein underlie the enhanced activity of MCR(red1), the only form of MCR that can react productively with the methyl group of methyl-SCoM.  相似文献   

9.
Methyl-coenzyme M reductase (MCR) catalyzes the formation of methyl-coenzyme M (CH(3)S-CH(2)CH(2)SO(3)) from methane. The active site is a nickel tetrahydrocorphinoid cofactor, factor 430, which in inactive form contains EPR-silent Ni(II). Two such forms, denoted MCR(silent) and MCR(ox1)(-)(silent), were previously structurally characterized by X-ray crystallography. We describe here the cryoreduction of both of these MCR forms by gamma-irradiation at 77 K, which yields reduced protein maintaining the structure of the oxidized starting material. Cryoreduction of MCR(silent) yields an EPR signal that strongly resembles that of MCR(red1), the active form of MCR; and stepwise annealing to 260-270 K leads to formation of MCR(red1). Cryoreduction of MCR(ox1)(-)(silent) solutions shows that our preparative method for this state yields enzyme that contains two major forms. One behaves similarly to MCR(silent), as shown by the observation that both of these forms give essentially the same redlike EPR signals upon cryoreduction, both of which give MCR(red1) upon annealing. The other form is assigned to the crystallographically characterized MCR(ox1)(-)(silent) and directly gives MCR(ox1) upon cryoreduction. X-band spectra of these cryoreduced samples, and of conventionally prepared MCR(red1) and MCR(ox1), all show resolved hyperfine splitting from four equivalent nitrogen ligands with coupling constants in agreement with those determined in previous EPR studies and from (14)N ENDOR of MCR(red1) and MCR(ox1). These experiments have confirmed that all EPR-visible forms of MCR contain Ni(I) and for the first time generated in vitro the EPR-visible, enzymatically active MCR(red1) and the activate-able "ready" MCR(ox1) from "silent" precursors. Because the solution Ni(II) species we assign as MCR(ox1)(-)(silent) gives as its primary cryoreduction product the Ni(I) state MCR(ox1), previous crystallographic data on MCR(ox1)(-)(silent) allow us to identify the exogenous axial ligand in MCR(ox1) as the thiolate from CoM; the cryoreduction experiments further allow us to propose possible axial ligands in MCR(red1). The availability of model compounds for MCR(red1) and MCR(ox1) also is discussed.  相似文献   

10.
11.
Methanogenic archaea utilize a specific pathway in their metabolism, converting C1 substrates (i.e., CO2) or acetate to methane and thereby providing energy for the cell. Methyl-coenzyme M reductase (MCR) catalyzes the key step in the process, namely methyl-coenzyme M (CH3-S-CoM) plus coenzyme B (HS-CoB) to methane and CoM-S-S-CoB. The active site of MCR contains the nickel porphinoid F430. We report here on the coordinated ligands of the two paramagnetic MCR red2 states, induced when HS-CoM (a reversible competitive inhibitor) and the second substrate HS-CoB or its analogue CH3-S-CoB are added to the enzyme in the active MCR red1 state (Ni(I)F430). Continuous wave and pulse EPR spectroscopy are used to show that the MCR red2a state exhibits a very large proton hyperfine interaction with principal values A((1)H) = [-43,-42,-5] MHz and thus represents formally a Ni(III)F430 hydride complex formed by oxidative addition to Ni(I). In view of the known ability of nickel hydrides to activate methane, and the growing body of evidence for the involvement of MCR in "reverse" methanogenesis (anaerobic oxidation of methane), we believe that the nickel hydride complex reported here could play a key role in helping to understand both the mechanism of "reverse" and "forward" methanogenesis.  相似文献   

12.
The UV-visible and electron paramagnetic resonance (EPR) spectra of MCR(red1), the catalytically active state of methyl-coenzyme M reductase, are almost identical to those observed when free coenzyme F430 or its pentamethyl ester (F430M) are reduced to the Ni(I) valence state. Investigations and proposals concerning the catalytic mechanism of MCR were therefore based on MCR(red1) containing Ni(I)F430 until, in a recent report, Tang et al. (J. Am. Chem. Soc. 2002, 124, 13242) interpreted their resonance Raman data and titration experiments as indicating that, in MCR(red1), coenzyme F430 is not only reduced at the nickel center but at one of the C=N double bonds of the hydrocorphinoid macrocycle as well. To resolve this contradiction, we have investigated the stoichiometry of the reduction of coenzyme F430 pentamethyl ester (F430M) by three independent methods. Spectroelectrochemistry showed clean reduction to a single product that exhibits the UV-vis spectrum typical for MCR(red1). In three bulk electrolysis experiments, 0.96 +/- 0.1 F/mol was required to generate the reduced species. Reduction with decamethylcobaltocene in tetrahydrofuran (THF) consumed 1 mol of (Cp)(2)Co/mol of F430M, and the stoichiometry of the reoxidation of the reduced form with the two-electron oxidant methylene blue was 0.46 +/- 0.05 mol of methylene blue/mol of reduced F430M. These experiments demonstrate that the reduction of coenzyme F430M to the species having almost identical UV-vis and EPR spectra as MCR(red1) is a one-electron process and therefore inconsistent with a reduction of the macrocycle chromophore.  相似文献   

13.
Saccharopine reductase from Magnaporthe grisea, an NADPH-containing enzyme in the α-aminoadipate pathway, catalyses the formation of saccharopine, a precursor to L-lysine, from the substrates glutamate and α-aminoadipate-δ-semialdehyde. Its catalytic mechanism has been investigated using quantum mechanics/molecular mechanics (QM/MM) ONIOM-based approaches. In particular, the overall catalytic pathway has been elucidated and the effects of electron correlation and the anisotropic polar protein environment have been examined via the use of the ONIOM(HF/6-31G(d):AMBER94) and ONIOM(MP2/6-31G(d)//HF/6-31G(d):AMBER94) methods within the mechanical embedding formulism and ONIOM(MP2/6-31G(d)//HF/6-31G(d):AMBER94) and ONIOM(MP2/6-311G(d,p)//HF/6-31G(d):AMBER94) within the electronic embedding formulism. The results of the present study suggest that saccharopine reductase utilises a substrate-assisted catalytic pathway in which acid/base groups within the cosubstrates themselves facilitate the mechanistically required proton transfers. Thus, the enzyme appears to act most likely by binding the three required reactant molecules glutamate, α-aminoadipate-δ-semialdehyde and NADPH in a manner and polar environment conducive to reaction.  相似文献   

14.
Results of our studies on polymerization kinetics and tests of copolymerization statistical models of ethylene-norbornene (E-N) copolymers obtained on the basis of microstructures determined by 13C NMR analysis are reported. Ethylene-norbornene (E-N) copolymers were synthesized by catalytic systems composed of racemic isospecific metallocenes, i-Pr[(3Pri-Cp)(Flu)]ZrCl2 or a constrained geometry catalyst (CGC) and methylaluminoxane. Polymerization kinetics revealed that E-N copolymerization is quasi living under standard polymerization conditions. Calculations of the number of active sites and of chain propagation and chain transfer turnover frequencies indicate that the metal is mainly in the Mt-N* state, while the Mt-E* state contributes more to transfer and propagation rates. The first-order and the second-order Markov statistics have been tested by using the complete tetrad distribution obtained from 13C NMR analysis of copolymer microstructures. The root-mean-square deviations between experimental and calculated tetrads demonstrate that penultimate (second-order Markov) effects play a decisive role in E-N copolymerizations. Results show clues for more complex effects depending on the catalyst geometry in copolymers obtained at high N/E feed ratios. Comonomer concentration was shown to have a strong influence on copolymer microstructure and copolymer properties. The copolymer microstructure of alternating isotactic copolymers obtained with i-Pr[(3Pri-Cp)(Flu)]ZrCl2 have been described at pentad level. Second-order Markov statistics better describes also the microstrucure of these copolymers.  相似文献   

15.
The mechanisms of liquid-phase phenol ozonation are revised. A new mechanism in which a significant role is played by free-radical reactions is suggested for this process.  相似文献   

16.
17.
We identify and discuss a non-adiabatic channel for triplet state relaxation. The contribution of this channel is found to depend strongly on the spacing between triplet levels with the same spin projection. The present analysis, which requires the familiar radiative and spin-orbit contributions in addition to the new non-adiabatic channel, provides an explanation of experimental data for substituted benzenes.  相似文献   

18.
On the mechanism of multi-bubble sonoluminescence   总被引:1,自引:0,他引:1  
The thermal chemiluminescent model of multibubble sonoluminescence (MBSL) is considered, and the contradictions that follow from its basic propositions are analyzed. It is shown that, if the thermal mechanism of sonoluminescence (SL) is operative, the continuous spectrum should be emitted before the band spectrum. It is established that, if the thermal recombination mechanism of a solid is operative, the duration of SL should be at least two orders of magnitude longer than that observed experimentally (<5 ns); according to the chemiluminescent model of SL, the SL burst time must be several orders of magnitude longer. Consideration is given to the previously proposed mechanism of emission of spectral lines of metals from solutions of alkali and alkaline-earth metal salts, which correlates with results of sonolysis of salt solutions under the action of ultrasound (US) pulses. A mechanism based on the theory of local electrification of cavitation bubbles in the US field is put forward to treat MBSL. It is shown that this mechanism agrees with experimental results: the SL burst time corresponds to the characteristic time of fluorescence.  相似文献   

19.
Summary Along with detectors of the universal type, selective detectors showing responses of a variable degree to the groups of compounds containing different elements or functional groups have found wide use recently, One such detector is a thermionic detector (TID) showing high sensitivity and selectivity towards phosphorus-containing compounds. At present, a number of companies produce TID of various types which found particularly wide use in analysis of organophosporus pesticides. Despite accumulation of rather extensive facts on TID performance, the mechanism of detecting by a thermionic detector remains vague in many respects [1]. It was the purpose of this paper to carry out experiments for clearing some processes of thermionic detection.
Über den Mechanismus des thermionischen Nachweises
Zusammenfassung Neben den Detektoren mit universeller Anwendung haben in jüngster Zeit selektive Detektoren ausgedehnte Anwendung gefunden, die ein mehr oder weniger spezifisches Ansprechverhalten gegenüber Verbindungsklassen mit verschiedenen Elementen oder funktionellen Gruppen zeigen. Ein derartiger Detektor ist der Thermionische Detektor (TID), der eine hohe Empfindlichkeit und Selektivität gegenüber phosphorhaltigen Verbindungen aufweist. Zur Zeit stellt eine Anzahl von Firmen TID's verschiedener Arten her, die in besonderem Maße zur Analyse von phosphororganischen Pesticiden eigesetzt wurden. Trotz der inzwischen vorliegenden ausführlichen Angaben über die Leistung des TID blieb der Meß-Mechanismus durch einen thermionischen Detektor bisher in mancher Beziehung unklar [1]. Es war der Zweck dieser Arbeit, Untersuchungen zur Aufklärung einiger Prozesse des thermionischen Nachweises unter Mithilfe anorganischer Salze anzustellen.

Sur le mecanisme de la detection thermionique
Sommaire A coté des détecteurs du type universal, des détecteurs sélectifs sont de plus en plus employés, ils ont une réponse plus ou moins marquée pour des composés qui contiennent des éléments ou des groupes fonctionnels définis. Un exemple de détecteur sélectif est le détecteur thermionique (DTI) à haute sensibilité et sélectivité pour les composés qui contiennent du phosphore. A présent, plusieurs fabricants produisent différents types de DTI dont on se sert surtout pour l'analyse des pesticides organophosphoriques. Malgré l'accumulation de données sur la performance du DTI, le mécanisme de la détection par un détecteur thermionique reste encore mal connu à plusieurs égards [1]. Le présent travail a pour but d'effecteur des essais afin d'élucider quelques processus de la détection thermionique.
  相似文献   

20.
The catalytic mechanism of nitrate reduction by periplasmic nitrate reductases has been investigated using theoretical and computational means. We have found that the nitrate molecule binds to the active site with the Mo ion in the +6 oxidation state. Electron transfer to the active site occurs only in the proton‐electron transfer stage, where the MoV species plays an important role in catalysis. The presence of the sulfur atom in the molybdenum coordination sphere creates a pseudo‐dithiolene ligand that protects it from any direct attack from the solvent. Upon the nitrate binding there is a conformational rearrangement of this ring that allows the direct contact of the nitrate with MoVI ion. This rearrangement is stabilized by the conserved methionines Met141 and Met308. The reduction of nitrate into nitrite occurs in the second step of the mechanism where the two dimethyl‐dithiolene ligands have a key role in spreading the excess of negative charge near the Mo atom to make it available for the chemical reaction. The reaction involves the oxidation of the sulfur atoms and not of the molybdenum as previously suggested. The mechanism involves a molybdenum and sulfur‐based redox chemistry instead of the currently accepted redox chemistry based only on the Mo ion. The second part of the mechanism involves two protonation steps that are promoted by the presence of MoV species. MoVI intermediates might also be present in this stage depending on the availability of protons and electrons. Once the water molecule is generated only the MoVI species allow water molecule dissociation, and, the concomitant enzymatic turnover. © 2009 Wiley Periodicals, Inc. J Comput Chem, 2009  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