共查询到20条相似文献,搜索用时 15 毫秒
1.
Oxidative dehydrogenation of cyclohexane to benzene, catalyzed by copperchromia catalyst, has been carried out. Data was analyzed
according to the first order rate law. It is concluded that the rate is affected by changes in the volume flow rate due to
changes, in the value of the contact time as well as the distribution of the molecules transformed over a larger or smaller volume. 相似文献
2.
The master equation of a lattice gas reaction tracks the probability of visiting all spatial configurations. The large number of unique spatial configurations on a lattice renders master equation simulations infeasible for even small lattices. In this work, a reduced master equation is derived for the probability distribution of the coverages in the infinite diffusion limit. This derivation justifies the widely used assumption that the adlayer is in equilibrium for the current coverages and temperature when all reactants are highly mobile. Given the reduced master equation, two novel and efficient simulation methods of lattice gas reactions in the infinite diffusion limit are derived. The first method involves solving the reduced master equation directly for small lattices, which is intractable in configuration space. The second method involves reducing the master equation further in the large lattice limit to a set of differential equations that tracks only the species coverages. Solution of the reduced master equation and differential equations requires information that can be obtained through short, diffusion-only kinetic Monte Carlo simulation runs at each coverage. These simulations need to be run only once because the data can be stored and used for simulations with any set of kinetic parameters, gas-phase concentrations, and initial conditions. An idealized CO oxidation reaction mechanism with strong lateral interactions is used as an example system for demonstrating the reduced master equation and deterministic simulation techniques. 相似文献
3.
Wolfgang Preis 《Monatshefte für Chemie / Chemical Monthly》2009,15(9):1059-1068
Abstract Surface exchange reactions and chemical diffusion in composites, consisting of a dilute distribution of inclusions in a matrix,
and polycrystalline materials have been modelled by application of both a square grain and a spherical grain model. The diffusion
equations have been solved numerically by employing a finite element approach in the case of the square grain model and the
Laplace transform method involving numerical Laplace inversion with respect to the spherical grain model. The boundary conditions
refer to oxygen exchange reactions between a gas phase and a mixed ionically–electronically conducting ceramic sample within
the linear response regime, i.e. small variations of the oxygen partial pressure. Diffusion profiles as well as the time dependence
of the total amount of exchanged oxygen (relaxation curves) have been calculated. A necessary requirement for effective medium
diffusion is proposed, and appropriate relations for the effective chemical surface exchange coefficient and the effective
chemical diffusion coefficient are derived. On the contrary, when the time constant for diffusion from the matrix into the
inclusions of a composite exceeds considerably the relaxation time for effective medium diffusion, relaxation curves with
two separate time constants are observed. Analogously, in the case of polycrystalline materials the overall transport process
is determined by slow (rate-limiting) bulk diffusion from the grain boundaries into the grains. Adequate formulae for the
relaxation times are given based on analytical approximations of the solution functions to the diffusion equations. In addition,
the spherical grain model is applied to interpret the re-oxidation kinetics of the positive temperature coefficient of resistivity
(PTC) ceramics based on conductivity relaxation experiments.
Graphical abstract
相似文献
4.
Preis W 《Physical chemistry chemical physics : PCCP》2006,8(22):2629-2634
Analytical solutions to the diffusion equations for fast grain boundary diffusion and surface exchange reactions at active surface sites are derived. The microstructure of the polycrystalline sample of finite thickness is modelled by parallel grain boundaries. The ratio between the surface exchange coefficient and the diffusion coefficient for the grain boundaries is assumed to be greater than or equal to that for the bulk. The analytical solutions allow the calculation of diffusion profiles for thin films. Special emphasis is laid on a detailed analysis of the time dependence of the total amount of diffusant exchanged between the constant diffusion source (e.g. gas phase) and the polycrystalline sample (e.g. oxide ceramics) which corresponds to relaxation curves obtained from, e.g., oxygen exchange measurements. The calculated relaxation curves refer to Harrison's type-A kinetics where homogeneous medium solutions are satisfactorily applicable, introducing effective kinetic parameters. Apart from expressions for the effective diffusion coefficient analogous relations for the effective surface exchange coefficient are proposed, depending on the microstructure of the polycrystalline material and the corresponding kinetic parameters of bulk and grain boundary regions, respectively. 相似文献
5.
The roles of benzoic acid and water on the Michael reaction of pentanal and nitrostyrene catalyzed by diarylprolinol silyl ether are revealed by density functional theory calculations. The calculations demonstrate that the benzoic acid is ready to attack the catalysts and form a hydrogen bond between the hydrogen atom of the COOH of benzoic acid and one of the N atoms of the catalyst. The complex formed from pentanal, catalyst and benzoic acid attacks nitroalkene and forms transition states. Finally, the transition states hydrolyze and the products are formed. The calculations demonstrate that the stereoselectivity is dominated by the steric hindrance of the 2-substituent groups, and the benzoic acid can increase the reaction rate evidently by decreasing the activation energies; however, H(3)O(+) or strong acid may prevent the formation of the transition states between enamines and nitroalkenes. The employed solvent can decrease the activation energies and promote the proton transfer from benzoic acid onto the catalyst 2. The calculated enantiomeric excess values are in good agreement with the experimental results. These calculations also reveal that the role of benzoic acid is dependent on the sophisticated structures of the catalysts and provide a valuable index for the structural design of new catalysts and selection of additives or co-catalysts. 相似文献
6.
7.
A. K. Chekalov N. G. Gvazava V. B. Vol'eva T. I. Prokof'eva A. I. Prokof'ev V. V. Ershov 《Russian Chemical Bulletin》1990,39(5):991-993
Tetrafluorosilane- catalyzed redox reactions of 3,6-di-tert-butyl-ortho-benzoquinone have been examined. From the EPR spectroscopic data and a study of the reaction products, a mechanism for these reactions is proposed involving the formation of a quinone- tetrafluorosilane complex which has strong oxidizing properties, and which enables the original quinone to undergo further reactions.Translated from Izvestiya Akademii Nauk SSSR, Seriya Khimicheskaya, No. 5, pp. 1105–1108, May, 1990. 相似文献
8.
Tuba Dedeoglu Hakan Durmaz Gurkan Hizal Umit Tunca 《Journal of polymer science. Part A, Polymer chemistry》2012,50(10):1917-1925
We designed a trifunctional initiator ( 3 ) containing anthracene, bromide, and OH functionalities and subsequently used as an initiator in atom transfer radical Polymerization (ATRP) of styrene to yield linear polystyrene (PS) with α‐anthracene, OH, and ω‐bromide terminal groups, of which bromide is later transformed into azide to result in the linear anthracene‐, OH‐, and azide‐terminated PS (l‐α‐anthracene‐OH‐ω‐azide‐PS). The copper‐catalyzed azide–alkyne cycloaddition reaction between l‐α‐anthracene‐OH‐ω‐azide‐PS and α‐furan‐protected‐maleimide‐ω‐alkyne linkage, 4 afforded the linear anthracene‐, OH‐, and maleimide‐terminated PS. The cyclization via intramolecular Diels–Alder click reaction of this linear PS and the subsequent conversion of the hydroxyl into bromide resulted in the cyclic PS with one bromide located on the ring, (c‐PS)‐Br. Finally, the c‐PS‐Br was clicked with either well‐defined tetramethylpiperidine‐1‐oxyl‐terminated poly(ethylene glycol) (PEG) or poly(ε‐caprolactone) (PCL) yielding the tadpole polymer, (c‐PS)‐b‐PEG or (c‐PS)‐b‐PCL. © 2012 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem, 2012 相似文献
9.
Mora-Seró I Dittrich T Belaidi A Garcia-Belmonte G Bisquert J 《The journal of physical chemistry. B》2005,109(31):14932-14938
Surface photovoltage transients were used to monitor both the short time dynamics (>10 ns) and the spatial distribution of electrons photoinjected in thin (2-20 nm) TiO2 layers from dye molecules adsorbed at the surface. At low temperatures (100-250 K), the dynamics are governed exclusively by spatially dependent tunneling recombination, with a rate that varies with the distance from the surface x as exp(-2x/a), and an initial exponential distribution of photoinjected electrons, n0 exp(-x/b). This model is confirmed by the observation of power law decay in time t(-a/2b) with a ratio a/b = 0.28 +/- 0.04. The stability of cis-di(isothiocyanato)-N-bis(2,2'-bipyridine-4,4'-dicarboxy) ruthenium(II) (N3) dye molecules on TiO2 during treatment in a vacuum at high temperatures was proven. For high temperatures (250-540 K), the thickness dependence of the decays indicates that the dynamics of surface recombination are retarded by the diffusion of electrons toward the interior of the film. The implications for thin layer coating in dye-sensitized solar cells are discussed. 相似文献
10.
《Tetrahedron letters》1986,27(1):83-86
Zinc homoenolate of alkyl propionate undergoes clean CC bond forming reactions with aryl, vinyl, and acid halides in the presence of a palladium catalyst. 相似文献
11.
M. Emre Hanhan 《应用有机金属化学》2008,22(5):270-275
A range of sterically hindered diimine ligands and their palladium (II) complexes were synthesized. These compounds were fully characterized by elemental analysis, 1H and 13C‐NMR spectroscopy. The use of the palladium complexes as catalysts for Suzuki and Heck coupling has been studied in an attempt to demonstrate the effect of side groups on catalytic activity. It was clearly seen that the location of side ? CH3 groups which bound to benzene ring had little effect on catalytic activity. Interestingly when we changed these ? CH3 groups with ? Cl groups the activity of the complexes increased. On the other hand, side groups which bound to imine nitrogen also had a large effect on catalytic activity. Copyright © 2008 John Wiley & Sons, Ltd. 相似文献
12.
本文采用飞行时间质谱技术测定了在紫外(355nm), 可见(560nm)和近代红外(1064nm)脉冲激光作用下, 氯分子束与Ti表面反应产物的质量分布和速度分布。所得结果表明, 不同波长激光诱导反应的主要产物相同, 有Ti, TiCl, TiCl3和TiCl4。在高能量密度的紫外激光作用下, 首次测得具有很高动能的原生Ti+。各种含Ti氯化物的飞行时间谱, 能满意地用单组分或多组分Maxwell-Boltzmann公式拟合和分析。上述激光诱导气-固表面反应的机理主要由氯分子在Ti表面上的解离吸附,吸附态氯原子在表面上生成TiClx(X=1~4)的连串反应以及激光诱导脱附所组成。近红外激光主要引起热脱附, 而紫外激光的作用还原可能有非热脱附过程。 相似文献
13.
The possibility to afford by a biomimetic reaction ‘unnatural’ products, which could offer a better bioactivity profile than natural analogues, is outlined and the first applications to the synthesis of lignans and related compounds have been reported. Here we describe the synthesis of new heterodimers, having a phenylcoumaran skeleton, by horseradish peroxidase catalyzed cross-coupling reactions of methyl esters of substituted hydroxycinnamic acids. 相似文献
14.
The cyclopentadienyl molybdenum hydride compounds, CpRMo(PMe3)3–x(CO)xH (CpR = Cp, Cp*; x = 0, 1, 2 or 3), are catalysts for the dehydrogenation of formic acid, with the most active catalysts having the composition CpRMo(PMe3)2(CO)H. The mechanism of the catalytic cycle is proposed to involve (i) protonation of the molybdenum hydride complex, (ii) elimination of H2 and coordination of formate, and (iii) decarboxylation of the formate ligand to regenerate the hydride species. NMR spectroscopy indicates that the nature of the resting state depends on the composition of the catalyst. For example, (i) the resting states for the CpMo(CO)3H and CpMo(PMe3)(CO)2H systems are the hydride complexes themselves, (ii) the resting state for the CpMo(PMe3)3H system is the protonated species [CpMo(PMe3)3H2]+, and (iii) the resting state for the CpMo(PMe3)2(CO)H system is the formate complex, CpMo(PMe3)2(CO)(κ1-O2CH), in the presence of a high concentration of formic acid, but CpMo(PMe3)2(CO)H when the concentration of acid is low. While CO2 and H2 are the principal products of the catalytic reaction induced by CpRMo(PMe3)3–x(CO)xH, methanol and methyl formate are also observed. The generation of methanol is a consequence of disproportionation of formic acid, while methyl formate is a product of subsequent esterification. The disproportionation of formic acid is a manifestation of a transfer hydrogenation reaction, which may also be applied to the reduction of aldehydes and ketones. Thus, CpMo(CO)3H also catalyzes the reduction of a variety of ketones and aldehydes to alcohols by formic acid, via a mechanism that involves ionic hydrogenation. 相似文献
15.
Bryan Branstetter 《Tetrahedron letters》2006,47(2):221-223
The formation of cyclopropanes from α,β-unsaturated aldehydes and diazo compounds has been a rather challenging goal due to the extremely reactive aldehyde starter. Herein, our group reports the first formation of ethyl 2-formyl-1-cyclopropanecarboxylate in 100% yield from the acid catalyzed reaction between acrolein and ethyl diazoacetate (EDA). 相似文献
16.
Ufuk Saim Gunay Hakan Durmaz Eda Gungor Aydan Dag Gurkan Hizal Umit Tunca 《Journal of polymer science. Part A, Polymer chemistry》2012,50(4):729-735
Well‐defined linear furan‐protected maleimide‐terminated poly(ethylene glycol) (PEG‐MI), tetramethylpiperidine‐1‐oxyl‐terminated poly(ε‐caprolactone) (PCL‐TEMPO), and azide‐terminated polystyrene (PS‐N3) or ‐poly(N‐butyl oxanorbornene imide) (PONB‐N3) were ligated to an orthogonally functionalized core ( 1 ) in a two‐step reaction mode through triple click reactions. In a first step, Diels–Alder click reaction of PEG‐MI with 1 was performed in toluene at 110 °C for 24 h to afford α‐alkyne‐α‐bromide‐terminated PEG (PEG‐alkyne/Br). As a second step, this precursor was subsequently ligated with the PCL‐TEMPO and PS‐N3 or PONB‐N3 in N,N‐dimethylformamide at room temperature for 12 h catalyzed by Cu(0)/Cu(I) through copper‐catalyzed azide‐alkyne cycloaddition and nitroxide radical coupling click reactions, yield resulting ABC miktoarm star polymers in a one‐pot mode. © 2011 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem, 2012 相似文献
17.
Alcohol oxidation reactions catalyzed by ruthenium–carbonyl complexes of thioarylazoimidazoles 下载免费PDF全文
Shyamal Kumar Sarkar Mahendra Sekhar Jana Tapan Kumar Mondal Chittaranjan Sinha 《应用有机金属化学》2014,28(8):641-651
Alcohols are oxidized by N‐methylmorpholine‐N‐oxide (NMO), ButOOH and H2O2 to the corresponding aldehydes or ketones in the presence of catalyst, [RuH(CO)(PPh3)2(SRaaiNR′)]PF6 ( 2 ) and [RuCl(CO)(PPh3)(SκRaaiNR′)]PF6 ( 3 ) (SRaaiNR′ ( 1 ) = 1‐alkyl‐2‐{(o‐thioalkyl)phenylazo}imidazole, a bidentate N(imidazolyl) (N), N(azo) (N′) chelator and SκRaaiNR′ is a tridentate N(imidazolyl) (N), N(azo) (N′), Sκ‐R is tridentate chelator; R and R′ are Me and Et). The single‐crystal X‐ray structures of [RuH(CO)(PPh3)2(SMeaaiNMe)]PF6 ( 2a ) (SMeaaiNMe = 1‐methyl‐2‐{(o‐thioethyl)phenylazo}imidazole) and [RuH(CO)(PPh3)2(SEtaaiNEt)]PF6 ( 2b ) (SEtaaiNEt = 1‐ethyl‐2‐{(o‐thioethyl)phenylazo}imidazole) show bidentate N,N′ chelation, while in [RuCl(CO)(PPh3)(SκEtaaiNEt)]PF6 ( 3b ) the ligand SκEtaaiNEt serves as tridentate N,N′,S chelator. The cyclic voltammogram shows RuIII/RuII (~1.1 V) and RuIV/RuIII (~1.7 V) couples of the complexes 2 while RuIII/RuII (1.26 V) couple is observed only in 3 along with azo reductions in the potential window +2.0 to ?2.0 V. DFT computation has been used to explain the spectra and redox properties of the complexes. In the oxidation reaction NMO acts as best oxidant and [RuCl(CO)(PPh3)(SκRaaiNR′)](PF6) ( 3 ) is the best catalyst. The formation of high‐valent RuIV=O species as a catalytic intermediate is proposed for the oxidation process. Copyright © 2014 John Wiley & Sons, Ltd. 相似文献
18.
The first order intramolecular rate constant for the reaction between the terminal groups of flexible macromolecules is calculated in the partial draining case, following the theory formulated by Wilemski and Fixman for diffusion controlled reactions. Substantial differences with respect to the nondraining and free-draining limits are evidenced, and it is shown that the rate constant in the partial-draining case has relevant contributions from all the modes of the bead-spring chain. The effects of chain flexibility and hydrodynamic interaction in macromolecules of increasing molecular weight are examined. The effective diffusion constant of the end groups increases with both the flexibility and the length of the chain. Numerical results for polystyrene (PS) and polydimethylsiloxane (PDMS) are presented and employed to calculate the relative quantum yield for fluorescence quenching. For highly flexible chains, like PDMS, quenching effects are expected in a range of molecular weight well above the limit of validity of the bead-spring model. On the contrary, for more rigid polymers, like PS, the quenching can be observed only at molecular weights lower than this limit. The calculated behavior is compared with some experimental results recently obtained by the authors. 相似文献
19.
20.
Pascal Vermeeren Trevor A. Hamlin Israel Fernndez F. Matthias Bickelhaupt 《Chemical science》2020,11(31):8105
The Diels–Alder reactions between cyclopentadiene and various α,β-unsaturated aldehyde, imine, and iminium dienophiles were quantum chemically studied using a combined density functional theory and coupled-cluster theory approach. Simple iminium catalysts accelerate the Diels–Alder reactions by lowering the reaction barrier up to 20 kcal mol−1 compared to the parent aldehyde and imine reactions. Our detailed activation strain and Kohn–Sham molecular orbital analyses reveal that the iminium catalysts enhance the reactivity by reducing the steric (Pauli) repulsion between the diene and dienophile, which originates from both a more asynchronous reaction mode and a more significant polarization of the π-system away from the incoming diene compared to aldehyde and imine analogs. Notably, we establish that the driving force behind the asynchronicity of the herein studied Diels–Alder reactions is the relief of destabilizing steric (Pauli) repulsion and not the orbital interaction between the terminal carbon of the dienophile and the diene, which is the widely accepted rationale.Quantum chemical activation strain analyses revealed that iminium catalysts accelerate Diels–Alder reactions by reducing the Pauli repulsion between reactants. 相似文献