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.
Calculations using different quantum mechanical methods including semiempirical (MNDO,AM1 and PM3), ab initio (RHF and MP2 calculations using the 6-311G and 6-311++G** basis sets), and density functional theory (LSDA, BP, MIXBP and B3LYP, i.e., B3LYP/6-311+G**//B3LYP/6-31G*) have been performed on the thermal fragmentation of cyclopropanone to ethylene and carbon monoxide. All RHF calculations predict a concerted single step mechanism for this conversion. The estimated activation energies vary from 34.4 to 54.6 kcal·mol-1, mainly localized around 37±2 kcal·mol-1, depending on the method. Whereas the calculated RHF reaction energies also varied from 14.5 to -33.3 kcal·mol-1, the B3LYP/6-311+G**//B3LYP/6-31G* method predicts the experimental value (-17.7 kcal·mol-1) within experimental uncertainties. Remarkably, semiempirical AM1 and PM3 methods and simple DFT calculations, LSDA, predict comparable results to the more advanced methods. UHF ab initio calculations predict the same single step mechanism, whereas a multistep biradical mechanism with an unrealistically low activation energy is favored by the semiempirical methods. Structures of the activated complex of the single step mechanism, estimated by different methods, are very similar and consistent with a nonlinear cheletropic [2s + 2a] reaction, as predicted by the orbital symmetry rules and earlier EHT calculations.Electronic Supplementary Material available.  相似文献   

2.
The coordination environment of the CuB center of the quinol oxidase from Acidianus ambivalens, a type B heme–copper oxygen reductase, was investigated by Fourier transform (FT) IR and extended X-ray absorption fine structure (EXAFS) spectroscopy. The comparative structural chemistry of dinuclear Fe–Cu sites of the different types of oxygen reductases is of great interest. Fully reduced A. ambivalens quinol oxidase binds CO at the heme a 3 center, with ν(CO)=1,973 cm−1. On photolysis, the CO migrated to the CuB center, forming a CuBI–CO complex with ν(CO)=2,047 cm−1. Raising the temperature of the samples to 25°C did not result in a total loss of signal in the FTIR difference spectrum although the intensity of these signals was reduced sevenfold. This observation is consistent with a large energy barrier against the geminate rebinding of CO to the heme iron from CuB, a restricted limited access at the active-site pocket for a second binding, and a kinetically stable CuB–CO complex in A. ambivalens aa 3. The CuB center was probed in a number of different states using EXAFS spectroscopy. The oxidized state was best simulated by three histidines and a solvent O scatterer. On reduction, the site became three-coordinate, but in contrast to the bo 3 enzyme, there was no evidence for heterogeneity of binding of the coordinated histidines. The CuB centers in both the oxidized and the reduced enzymes also appeared to contain substoichiometric amounts (0.2 mol equiv) of nonlabile chloride ion. EXAFS data of the reduced carbonylated enzyme showed no difference between dark and photolyzed forms. The spectra could be well fit by 2.5 imidazoles, 0.5 Cl and 0.5 CO ligands. This arrangement of scatterers would be consistent with about half the sites remaining as unligated Cu(his)3 and half being converted to Cu(his)2ClCO, a 50/50 ratio of Cu(his)2Cl and Cu(his)3CO, or some combination of these formulations. Electronic Supplementary Material Supplementary material is available for this article at .  相似文献   

3.
The mechanism for the oxidation of catechol by catechol oxidase has been studied using B3LYP hybrid density functional theory. On the basis of the X-ray structure of the enzyme, the molecular system investigated includes the first-shell protein ligands of the two metal centers as well as the second-shell ligand Cys92. The cycle starts out with the oxidized, open-shell singlet complex with oxidation states Cu2(II,II) with a μ-η22 bridging peroxide, as suggested experimentally, which is obtained from the oxidation of Cu2(I,I) by dioxygen. The substrate of each half-reaction is a catechol molecule approaching the dicopper complex: the first half-reaction involves Cu(I) oxidation by peroxide and the second one Cu(II) reduction. The quantitative potential energy profile of the reaction is discussed in connection with experimental data. Since no protons leave or enter the active site during the catalytic cycle, no external base is required. Unlike the previous density functional theory study, the dicopper complex has a charge of +2.  相似文献   

4.
The computational prediction of gas phase enthalpy (neutral substrates) and aqueous free energy (anion substrates) changes has been evaluated for the oxygen atom transfer reaction X + 1/2O2 → XO. Several density functionals (SVWN, BP86, B3LYP) at double- and triple-ζ levels were surveyed, along with one composite ab initio method (G3(MP2)). Results are presented for extensive main group test sets for which experimental thermochemistry is available. In addition, several minimal reaction couples of the type [MIVOL2]/[MVIO2L2] (M = Mo, W) have been examined. Overall, the results suggest a computational approach to the energetics of oxo transfer is feasible, potentially affording an expanded oxo transfer reactivity scale.  相似文献   

5.
This study uses the Gaussian 03 program and density functional theory B3LYP with three basis set methods—[B3LYP/6-311+G(d,p), B3LYP/6-31+G(2d,p), and B3LYP/6-31G(d,p)]—to model the highly energetic ionic compound diguanidinium 5,5′-azotetrazolate (GZT) to research its decomposition mechanisms and thermodynamic properties. Molecular-type cracking patterns are proposed, which were initiated by heterocyclic ring opening, sequential cracking of the two five-membered rings of GZT, and simultaneous release of N2 molecules; whereas proton transfer, bond-breaking, and atomic rearrangements were performed subsequently. Finally, 15 reaction paths and five transition states were obtained. All possible decomposition species and transition states, including intermediates and products, were identified, and their corresponding enthalpy and Gibbs free energy values were obtained. The results revealed that (1) the maximum activation energy required is 187.8 kJ mol–1, and the enthalpy change (ΔH) and Gibbs free-energy change (ΔG) of the net reaction are ?525.1 kJ mol–1 and ?935.6 kJ mol–1, respectively; (2) GZT can release large amounts of energy, the main contribution being from the disintegration of the 5,5'-azotetrazolate anion (ZT2?) skeleton (ΔH?=??598.3 kJ mol–1); and (3) the final products contained major amounts of N2 gas, but remaining gas molecules such as HCN and NH3 were obtained, which are in agreement with experimental results. The detailed decomposition simulation results demonstrated the feasibility of this method to calculate the energies of the thermodynamic reactions for the highly energetic GZT and predict the most feasible pathways and the final products.  相似文献   

6.
The electronic and geometric structures of tetracyclo[5.3.0.02,6.03,10]deca-4,8-diene (hypostrophene) have been investigated by ab initio and DFT/B3LYP methods using the 6-31G* and 6-311G* basis sets. The double bonds of hypostrophene are endo-pyramidalized. The cationic intermediates and products formed in the addition reaction have been investigated using the HF/6-311G*, HF/6-311G**, and B3LYP/6-311G* methods. The bridged bromonium cation was more stable than the U-type cation. Considering that the bridged cation does not isomerize to the less stable U-type cation, it is not possible for the U-type product to be obtained in the reaction. The bridged bromonium cation transformed into the more stable N-type cation and the N-type product was obtained via this cation. The thermodynamic stability of the exo, exo and exo, endo isomers of the N-type dibromide molecule were almost identical. The N-type product was 16.6 kcal mol−1 more stable than the U-type product. Figure General energy diagram of the hypostrophene–bromine (HS–Br2) system (kcal mol−1) (MP2/6-311G*//HF/6-311G*)  相似文献   

7.
The monohydroxo-bridged dicopper(II) complex (1), its reduced dicopper(I) analogue (2) and the trans-μ-1,2-peroxo-dicopper(II) adduct (3) with the macrocyclic N-donor ligand [22]py4pz (9,22-bis(pyridin-2′-ylmethyl)-1,4,9,14,17,22,27,28,29,30- decaazapentacyclo -[22.2.114,7.111,14.117,20]triacontane-5,7(28),11(29),12,18,20(30), 24(27),25-octaene), have been prepared and characterized, including a 3D structure of 1 and 2. These compounds represent models of the three states of the catechol oxidase active site: met, deoxy (reduced) and oxy. The dicopper(II) complex 1 catalyzes the oxidation of catechol model substrates in aerobic conditions, while in the absence of dioxygen a stoichiometric oxidation takes place, leading to the formation of quinone and the respective dicopper(I) complex. The catalytic reaction follows a Michaelis–Menten behavior. The dicopper(I) complex binds molecular dioxygen at low temperature, forming a trans-μ-1,2-peroxo-dicopper adduct, which was characterized by UV–Vis and resonance Raman spectroscopy and electrochemically. This peroxo complex stoichiometrically oxidizes a second molecule of catechol in the absence of dioxygen. A catalytic mechanism of catechol oxidation by 1 has been proposed, and its relevance to the mechanisms earlier proposed for the natural enzyme and other copper complexes is discussed. Electronic Supplementary Material Supplementary material is available for this article at  相似文献   

8.
Six uridine and six deoxyuridine isomers were studied at the B3LYP and TD B3LYP theoretical level and 6–31+G(d) basis function. The stability and the excited states of the isomers were studied in order to clarify some known experimental data. It was established that the rotation of the oxo uracil ring in uridine is energetically more likely to occur in the excited state than in the ground state, driven by the bright 1 ππ* state and the dark charge transfer 1nπ* state. Very high energy barriers (on the So) were found for thermal intramolecular proton transfer processes.  相似文献   

9.
Ribulose-1,5-bisphosphate carboxylase/oxygenase (RuBisCO), the most important enzyme for the assimilation of carbon into biomass, features a well-known isotope effect with regards to the CO2 carbon atom. This kinetic isotope effect α = k 12/k 13 for the carboxylation step of the RuBisCO reaction sequence, and its microscopic origin, was investigated with the help of cluster models and quantum chemical methods [B3LYP/6-31G(d,p)]. We use a recently proposed model for the RuBisCO active site, in which a water molecule remains close to the reaction center during carboxylation of ribulose-1,5-bisphosphate [B. Kannappan, J.E. Gready, J. Am. Chem. Soc. 130 (2008), 15063]. Alternative active-site models and/or computational approaches were also tested. An isotope effect alpha for carboxylation is found, which is reasonably close to the one measured for the overall reaction, and which originates from a simple frequency shift of the bending vibration of 12CO2 compared to 13CO2. The latter is the dominant mode for the product formation at the transition state.  相似文献   

10.
We investigated the viability of using a BC2N nanotube to detect formaldehyde (H2CO) molecule by means of B3LYP and M06 density functionals. The results indicate that the molecule is weakly adsorbed on the intrinsic BC2N nanotube releasing energy of 0.8 kcal mol-1 (at B3LYP/6-31G(d)) without significant effect on the HOMO-LUMO energy gap and electrical conductivity of the tube. Thus, H2CO cannot be detected using this intrinsic nanotube. To overcome this problem, a carbon atom of the tube wall was substituted by a Si atom. It was demonstrated that the Si-doped tube cannot only strongly adsorb the H2CO molecule, but also may effectively detect its presence because of the increase in the electric conductivity of the tube.  相似文献   

11.
Carbon monoxide dehydrogenase from the bacterium Oligotropha carboxidovorans catalyzes the oxidation of CO to CO2 at a unique [CuSMoO2] cluster. In the bacteria the cluster is assembled post-translational. The integration of S, and particularly of Cu, is rate limiting in vivo, which leads to CO dehydrogenase preparations containing the mature and fully functional enzyme along with forms of the enzyme deficient in one or both of these elements. The active sites of mature and immature forms of CO dehydrogenase were converted into a [MoO3] centre by treatment with potassium cyanide. We have established a method, which rescues 50% of the CO dehydrogenase activity by in vitro reconstitution of the active site through the supply of sulphide first and subsequently of Cu(I) under reducing conditions. Immature forms of CO dehydrogenase isolated from the bacterium, which were deficient in S and/or Cu at the active site, were similarly activated. X-ray crystallography and electron paramagnetic resonance spectroscopy indicated that the [CuSMoO2] cluster was properly reconstructed. However, reconstituted CO dehydrogenase contains mature along with immature forms. The chemical reactions of the reconstitution of CO dehydrogenase are summarized in a model, which assumes resulphuration of the Mo-ion at both equatorial positions at a 1:1 molar ratio. One equatorial Mo–S group reacts with Cu(I) in a productive fashion yielding a mature, functional [CuSMoO2] cluster. The other Mo–S group reacts with Cu(I), then Cu2S is released and an oxo group is introduced from water, yielding an inactive [MoO3] centre.  相似文献   

12.
The hydrogen-activating cluster (H cluster) in [FeFe]-hydrogenases consists of two moieties. The [2Fe]H subcluster is a (L)(CO)(CN)Fe(μ-RS2)(μ-CO)Fe(CysS)(CO)(CN) centre. The Cys-bound Fe is called Fe1, the other iron Fe2. The Cys-thiol forms a bridge to a [4Fe–4S] cluster, the [4Fe–4S]H subcluster. We report that electron paramagnetic resonance (EPR) spectra of the 57Fe-enriched enzyme from Desulfovibrio desulfuricans in the Hox–CO state are consistent with a magnetic hyperfine interaction of the unpaired spin with all six Fe atoms of the H cluster. In contrast to the inactive aerobic enzyme, the active enzyme is easily destroyed by light. The [2Fe]H subcluster in some enzyme molecules loses CO by photolysis, whereupon other molecules firmly bind the released CO to form the Hox–CO state giving rise to the so-called axial 2.06 EPR signal. Though not destroyed by light, the Hox–CO state is affected by it. As demonstrated in the accompanying paper [49] two of the intrinsic COs, both bound to Fe2, can be exchanged by extrinsic 13CO during illumination at 2 °C. We found that only one of the three 13COs, the one at the extrinsic position, gives an EPR-detectable isotropic superhyperfine interaction of 0.6 mT. At 30 K both the inhibiting extrinsic CO bound to Fe2 and one more CO can be photolysed. EPR spectra of the photolysed products are consistent with a 3d 7 system of Fe with the formal oxidation state +1. The damaged enzyme shows a light-sensitive g=5 signal which is ascribed to an S=3/2 form of the [2Fe]H subcluster. The light sensitivity of the enzyme explains the occurrence of the g=5 signal and the axial 2.06 signal in published EPR spectra of nearly all preparations studied thus far.  相似文献   

13.
Spiro[2,4]hepta-4,6-diene and spiro[4,4]nona-1,3-diene react with [Mo(CO)2Cp’(NCCH3)2][BF4] (Cp’ = Cp, Ind; Ind = η5-C9H7) to afford the corresponding diene complex [Mo(diene)(CO)2Cp′]+. When Cp’ = Ind, the reaction proceeded forward leading to ring opening in the case of the small spiro ring. Although this and another product resulting from migration of the side arm to the carbonyl were detected when Cp’ = Cp, they did not form from the diene complex. A DFT/PBE1PBE study was carried out and showed a kinetically controlled reaction pathway leading from the [Mo(diene)(CO)2Ind]+ to the reaction product, with an activation barrier of 21.3 kcal mol−1. The thermodynamic preferred species was the non-observed complex (insertion), and its formation required higher barriers. In the presence of Cp, all the barriers increased significantly, explaining the inertness of the initial diene complex. The interpretation of this behaviour is associated with the ease of the η5 → η3 haptotropic rearrangement of the indenyl, which helps to lower some relevant barriers. This route is not available for the Cp analogue.  相似文献   

14.
The reaction of the air-tolerant CO dehydrogenase from Oligotropha carboxidovorans with H2 has been examined. Like the Ni-Fe CO dehydrogenase, the enzyme can be reduced by H2 with a limiting rate constant of 5.3 s−1 and a dissociation constant Kd of 525 μm; both kred and kred/Kd, reflecting the breakdown of the Michaelis complex and the reaction of free enzyme with free substrate in the low [S] regime, respectively, are largely pH-independent. During the reaction with H2, a new EPR signal arising from the Mo/Cu-containing active site of the enzyme is observed which is distinct from the signal seen when the enzyme is reduced by CO, with greater g anisotropy and larger hyperfine coupling to the active site 63,65Cu. The signal also exhibits hyperfine coupling to at least two solvent-exchangeable protons of bound substrate that are rapidly exchanged with solvent. Proton coupling is also evident in the EPR signal seen with the dithionite-reduced native enzyme, and this coupling is lost in the presence of bicarbonate. We attribute the coupled protons in the dithionite-reduced enzyme to coordinated water at the copper site in the native enzyme and conclude that bicarbonate is able to displace this water from the copper coordination sphere. On the basis of our results, a mechanism for H2 oxidation is proposed which involves initial binding of H2 to the copper of the binuclear center, displacing the bound water, followed by sequential deprotonation through a copper-hydride intermediate to reduce the binuclear center.  相似文献   

15.
Kinetic behavior of penicillin acylase immobilized on acrylic carrier   总被引:1,自引:0,他引:1  
The usefulness of Lilly's kinetic equation to describe penicillin G hydrolysis performed by immobilized penicillin acylase onto the acrylic carrier has been shown. Based on the experimental results characteristic kinetic constants have been estimated. The effect of noncompetitive inhibition of 6-amino penicillanic acid has not been found. Five components of reaction resistance have been defined. These components were also estimated for the reaction of the native enzyme as well as the Boehringer preparation.List of Symbols C E g/m3 enzyme concentration - C P,C Q mol/m3 product concentrations - C S mol/m3 substrate concentration - C SO mol/m3 initial substrate concentration - K A mol/m3 constant which defines the affinity of a substrate to the enzyme - K iS mol/m3 substrate inhibitory constant - K iP mol/m3 PhAA inhibitory constant - K iQ mol/m3 6-APA inhibitory constant - k 3 mol/g/min constant rate of dissociation of the active complex - R(1) concentrational component of reaction resistance - R(2) resistance component derived from substrate affinity - R(3) resistance component due to the inhibition of the enzyme by substrate - R(4) resistance component due to the inhibition of the enzyme by PhAA - R(5) resistance component due to inhibition of the enzyme by 6-APA - r = dCs/dt mol/m3 min rate of reaction - t min reaction time - (i) relative resistance of reaction  相似文献   

16.
A mechanism for the oxidation of 3,5-di-tert-butylcatechol (dtbc) with dioxygen to the corresponding quinone (dtbq), catalyzed by bispidine-dicopper complexes (bispidines are various mono- and dinucleating derivatives of 3,7-diazabicyclo[3.3.1]nonane with bis-tertiary-amine–bispyridyl or bis-tertiary-amine–trispyridyl donor sets), is proposed on the basis of (1) the stoichiometry of the reaction as well as the stabilities and structures [X-ray, density functional theory (B3LYP, TZV)] of the bispidine-dicopper(II)–3,4,5,6-tetrachlorcatechol intermediates, (2) formation kinetics and structures (molecular mechanics, MOMEC) of the end-on peroxo–dicopper(II) complexes and (3) kinetics of the stoichiometric (anaerobic) and catalytic (aerobic) copper-complex-assisted oxidation of dtbc. This involves (1) the oxidation of the dicopper(I) complexes with dioxygen to the corresponding end-on peroxo–dicopper(II) complexes, (2) coordination of dtbc as a bridging ligand upon liberation of H2O2 and (3) intramolecular electron transfer to produce dtbq, which is liberated, and the dicopper(I) catalyst. Although the bispidine complexes have reactivities comparable to those of recently published catalysts with macrocyclic ligands, which seem to reproduce the enzyme-catalyzed process in various reaction sequences, a strikingly different oxidation mechanism is derived from the bispidine–dicopper-catalyzed reaction. Electronic supplementary material Supplementary material is available in the online version of this article at and is accessible for authorized users.  相似文献   

17.
Two-layer ONIOM method at the ONIOM(B3LYP/6-31G(d):PM6) level of theory was applied to study the cycloaddition reaction of α-chlorocarbanions (CR2Cl, where R is H, Cl, CH3 CN, and NO2) and fullerene. The results show that the reaction pathways depend on the electron withdrawing functional groups or the electron donating functional groups contained in the α-chlorocarbanions. The energy profile analysis reveals that functionalization of fullerene by CCl3 , C(CH3)2Cl, and CH2Cl is more favorable than by C(CN)2Cl and C(NO2)2Cl in terms of the thermodynamic point of view.  相似文献   

18.
Molybdenum and tungsten complexes as models for the active sites of assimilatory or dissimilatory nitrate reductases (NR) were computed at the CPCM-B98/SDDp//B3LYP/Lanl2DZp* plus zero point energy level of density functional theory. The ligands were chosen on the basis of available experimental protein or small chemical model structures. A water molecule is found to bind to assimilatory NR models [(Me2C2S2)MO(YMe)] (−11.5 kcal mol−1 for M is Mo, Y is S) and may be replaced by nitrate (−4.5 kcal mol−1) (but a hydroxy group may not). Nature’s choice of M is Mo and Y is S for NR has the largest activation energy for protein-free models (13.3 kcal mol−1) and the least exothermic reaction energy for the nitrate reduction (−14.9 kcal mol−1) compared with M is W and Y is O or Se alternatives. Water binding to dissimilatory NR model complexes [(Me2C2S2)2M(YR)] is considerably endothermic (10.3 kcal mol−1); nitrate binding is only slightly so (1.5 kcal mol−1 for RY is MeS). The exchange of an oxo ligand (assimilatory NR) for a dithiolato ligand (dissimilatory NR model) reduces the exothermicity (−8.6 kcal mol−1 relative to the fivefold-coordinate reduced complex) and raises the barrier for oxygen atom transfer (OAT) in the nitrate complex (19.2 kcal mol−1). Not for the mono but only for the bisdithiolato complexes hydrogen bonding involving the coordinated substrate may significantly lower the OAT barrier as shown by explicitly adding water molecules. Substitution of tungsten for molybdenum generally lowers OAT activation energies and makes nitrate reduction reaction energies more negative. Bidentate carboxylato binding identified in Escherichia coli NarGHI is the preferred binding mode also for an acetato model. However, one dithiolato ligand folds when the MoVI center is bare of a good π-donor ligand, e.g., an oxo group. Computations on [(mnt)2MoIV(YR)(PPh3)] [mnt is (CN)2C2S2 2−] gave a smaller nitrate reduction activation energy for RY is Cl, compared with RY is PhS, although experimentally only the phenyl thiolato complex and not the chloro complex was found to be a functional NR model. Electronic supplementary material The online version of this article (doi:) contains supplementary material, which is available to authorized users.  相似文献   

19.
The mechanism of the enantioselective deprotonation of cyclohexene oxide with isopinocampheyl-based chiral lithium amide was studied by quantum chemical calculations. The transition states of eight molecules were fully optimized at the ab initio HF/3-21G and density functional B3LYP/3-21G levels with Gaussian 98. The activation energies were calculated at the B3LYP/6-31+G(3df,2p)//B3LYP/3-21G level. We found the theoretical evaluation to be consistent with the experimental data. At the best case, an enantiomeric excess of up to 95% for (R)-2-scyclohexen-1-ol was achieved with ()-N, N-diisopinocampheyl lithium amide. Enantioselective deprotonation of cyclohexene oxide Electronic Supplementary Material Supplementary material is available for this article at Dedicated to Professor Dr. Paul von Ragué Schleyer on the occasion of his 75th birthday.  相似文献   

20.
The catalytic mechanism of naphthalene 1,2-dioxygenase has been investigated by means of hybrid density functional theory. This Rieske-type enzyme, which contains an active site hosting a mononuclear non-heme iron(II) complex, uses dioxygen and two electrons provided by NADH to carry out the cis-dihydroxylation of naphthalene. Since a (hydro)peroxo-iron(III) moiety has been proposed to be involved in the catalytic cycle, it was probed whether and how this species is capable of cis-dihydroxylation of the aromatic substrate. Different oxidation and protonation states of the Fe–O2 complex were studied on the basis of the crystal structure of the enzyme with oxygen bound side-on to iron. It was found that feasible reaction pathways require a protonated peroxo ligand, FeIII–OOH; the deprotonated species, the peroxo-iron(III) complex, was found to be inert toward naphthalene. Among the different chemical patterns which have been explored, the most accessible one involves an epoxide intermediate, which may subsequently evolve toward an arene cation, and finally to the cis-diol. The possibility that an iron(V)-oxo species is formed prior to substrate hydroxylation was also examined, but found to implicate a rather high energy barrier. In contrast, a reasonably low barrier might lead to a high-valent iron-oxo species [i.e. iron(IV)-oxo] if a second external electron is supplied to the mononuclear iron center before dioxygenation.Electronic Supplementary Material Supplementary material is available in the online version of this article at http://dx.doi.org/10.1007/s00775-004-0537-0  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