scispace - formally typeset
Search or ask a question

Showing papers on "Reaction rate published in 2003"


Journal ArticleDOI
TL;DR: A review of the important area of cellulose degradation under alkaline conditions is presented in this paper, focusing on its relevance to the possible disposal of radioactive wastes in an underground repository in which cement-based waste encapsulation grouts and backfill may be employed.

429 citations


Journal ArticleDOI
TL;DR: In this paper, the kinetics of CO oxidation in excess hydrogen over a nanostructured Cu 0.1 Ce 0.9 O 2− y catalyst prepared by a sol-gel method was studied under simulated preferential oxidation (PROX) reactor conditions.

301 citations


Journal ArticleDOI
TL;DR: In this paper, the authors measured the kinetic parameters for water-gas shift reaction on Cu-based catalysts under fuel reformer conditions for fuel cell applications (7% CO, 8.5% CO2, 22% H2O, 37% H 2, and 25% Ar) at 1 atm total pressure and temperature in the range of 200°C.

242 citations


Journal ArticleDOI
TL;DR: In this paper, steady state infrared (IR) measurements for adsorption of only CO and under water-gas shift (WGS) reaction conditions indicate that formates are present on the surface of reduced ceria, and that their concentrations vary with surface area of partially reduced Ceria.
Abstract: Steady state infrared (IR) measurements for adsorption of only CO and under water–gas shift (WGS) reaction conditions indicate that formates are present on the surface of reduced ceria, and that their concentrations vary with surface area of partially reduced ceria. Under steady state WGS, the concentrations of surface formates are strongly limited at high CO conversions. However, at low temperatures and conversions, the formates are close to the equilibrium adsorption/desorption coverages obtained from only CO adsorption. Comparisons at constant temperature indicate that formate bands from IR may provide an indication of the number of active sites present on the catalyst surface, as the rates varied accordingly. The IR results favor a formate intermediate mechanism to explain WGS. However, more kinetic studies are required, and over a broad range of temperatures, to verify this conclusion. Previous low temperature kinetic studies at a relatively high CO/H2O ratios have produced a zero-order dependency for CO and the authors related this to a mechanistic scheme involving reaction of Pt-CO with CeO2 to yield CO2, followed by reoxidation of Ce2O3 by H2O, with liberation of H2. The zero-order was suggested to be due to saturation of noble metal surface with CO during WGS. Saturation of ceria with carbonates was also reported. In this study, a high H2O/CO ratio was used where the CO rate dependency was first-order. This criteria requires that the surface coverage of the adsorbed CO intermediate should be reaction rate limited. Therefore, the formates are suggested to be the intermediates.

237 citations


Journal ArticleDOI
TL;DR: Kinetic isotope effects established that the catalyst activation process occurs through a beta-hydride elimination pathway.
Abstract: A comparative kinetic examination of catalyst systems based on several monophosphinobiaryl ligands is reported. The bulk of the phosphine ligand controls the catalytic activity and the rate of catalyst activation with the catalyst based on 2-dicyclohexylphosphino-2‘,4‘,6‘-triisopropylbiphenyl providing the greatest activity and fastest activation. In the case where catalyst activation is slow (i.e., use of the smaller ligands such as 2-dicyclohexylphosphino-2‘-methylbiphenyl in combination with Pd(OAc)2) stirring the amine with the catalyst/base mixture prior to the commencement of the reaction increases the reaction rate along with the rate of catalyst activation. Kinetic isotope effects established that the catalyst activation process occurs through a β-hydride elimination pathway.

218 citations


Journal ArticleDOI
TL;DR: In this article, α-cellulose when treated with a varying level of H2SO4 underwent an abrupt change in physical structure (fibrous form to gelatinous form) at about 65% H 2SO4.
Abstract: Hydrolysis of α-cellulose by H2SO4 is a heterogeneous reaction. As such the reaction is influenced by physical factors. The hydrolysis reaction is therefore controlled not only by the reaction conditions (acid concentration and temperature) but also by the physical state of the cellulose. As evidence of this, the reaction rates measured at the high-temperature region (above 200°C) exhibited a sudden change in apparent activation energy at a certain temperature, deviating from Arrhenius law. Furthermore, α-cellulose, once it was dissolved into concentrated H2SO4 and reprecipitated, showed a reaction rate two orders of magnitude higher than that of untreated cellulose, about the same magnitude as cornstarch. The α-cellulose when treated with a varying level of H2SO4 underwent an abrupt change in physical structure (fibrous form to gelatinous form) at about 65% H2SO4. The sudden shift of physical structure and reaction pattern in response to acid concentration and temperature indicates that the main factor causing the change in cellulose structure is disruption of hydrogen bonding. Finding effective means of disrupting hydrogen bonding before or during the hydrolysis reaction may lead to a novel biomass saccharification process.

211 citations


Patent
18 Nov 2003
TL;DR: In this article, a process for the production of acetic acid by carbonylation of methanol, and reactive derivatives thereof, in a reaction mixture using a rhodium-based catalyst in low water conditions is described.
Abstract: The invention relates to a process for the production of acetic acid by carbonylation of methanol, and reactive derivatives thereof, in a reaction mixture using a rhodium-based catalyst in low water conditions. The process is used to achieve reaction rates of at least 15 g mol/l/hr. The high rate reactions proceed at water concentrations of less than 2.0 wt. %. Under certain conditions, the water concentration in the reaction mixture of the process is maintained at a desired concentration by at least one process step including adding a compound such as methyl acetate, dimethyl ether, acetic anhydride, or mixtures of these compounds to the reaction system. The process step of adding the components to the reaction mixture may be combined with other process steps for controlling water concentrations in reaction mixtures for the carbonylation of methanol.

198 citations


Journal ArticleDOI
TL;DR: In this article, the role of the catalysts are still unknown as well as their disposition within the NaAlH4 compound and the extent catalytic dehydrogenation reaction rate of a number of transition metal additions to NaAl H4 at 120°C.

195 citations


Journal ArticleDOI
07 Feb 2003-Science
TL;DR: Calculations indicated that at this temperature the reaction proceeds from a single quantum state of the reactant so that the computed rate constant has achieved a temperature-independent limit.
Abstract: We observed ring expansion of 1-methylcyclobutylfluorocarbene at 8 kelvin, a reaction that involves carbon tunneling. The measured rate constants were 4.0 × 10 −6 per second in nitrogen and 4 × 10 −5 per second in argon. Calculations indicated that at this temperature the reaction proceeds from a single quantum state of the reactant so that the computed rate constant has achieved a temperature-independent limit. According to calculations, the tunneling contribution to the rate is 152 orders of magnitude greater than the contribution from passage over the barrier. We discuss environmental effects of the solid-state inert-gas matrix on the reaction rate.

184 citations


Journal ArticleDOI
TL;DR: In this paper, an active gold catalyst for CO oxidation, supported on silica, made by a novel solution technique is reported, achieving a reaction rate of 2.7×10 −4 mmol g cat − 1 s −1 in CO oxidation.
Abstract: The preparation of an active gold catalyst for CO oxidation, supported on silica, made by a novel solution technique is reported. The surface of SBA-15 was functionalized with positively charged groups, and [AuCl4]−-species were subsequently incorporated into the channel system via ion exchange. Upon reduction with NaBH4, highly dispersed nanoparticles of gold were formed in the channels of the mesoporous host. A reaction rate of 2.7×10 −4 mmol g cat − 1 s −1 in CO oxidation was found for this composite material. Compared with a previously reported material obtained by a chemical vapor deposition (CVD) method, this reaction rate is still lower by about one order of magnitude. However, the straightforward preparation method reported here still yields more active catalysts than any other Au/SiO2 system made by a solution technique, which show almost no activity at room temperature. Since the gold particles are essentially isolated from the support by the organic coating of the channels, a support–metal interaction is highly improbable as the source of the catalytic activity of the gold. TEM images of Au/SBA-15 reveal that gold particles start sintering at temperatures higher than 100 °C. Due to this effect, a significant drop in catalytic activity is observed.

182 citations


Journal ArticleDOI
TL;DR: In this article, the reaction between synthetic ferrihydrite and dissolved sulfide was studied in artificial seawater and 0.42 M NaCl at 25 °C over the pH range 4.0-8.2.

Journal ArticleDOI
TL;DR: In this paper, the Eley-Rideal model has been used for the esterification of acetic acid with isobutanol in the presence of catalysts.
Abstract: Kinetic data on the esterification of acetic acid with isobutanol have been obtained from both the uncatalyzed and heterogeneously catalyzed reactions using a stirred batch reactor in dioxane. The equilibrium constant, which is independent of temperature ranging from 318 to 368 K, was found to be 4. The uncatalyzed reaction was proved to be second-order reversible. In the presence of the catalyst, on the other hand, the reaction has been found to occur between an adsorbed alcohol molecule and a molecule of acid in the bulk fluid (Eley–Rideal model). It has also been observed that the initial reaction rate decreases with alcohol and water concentrations and it linearly increases with that of acid. The temperature dependency of the constants appearing in the rate expression were also determined.

Journal ArticleDOI
TL;DR: In this article, the authors studied the catalytic hydrogenation of d-glucose to d-sorbitol over a 5% Ru/C catalyst in a semi-batch slurry autoclave operating at 373-403 K and 4.0-7.5 MPa hydrogen pressure.
Abstract: The catalytic hydrogenation of d -glucose to d -sorbitol over a 5% Ru/C catalyst was studied in a semi-batch slurry autoclave operating at 373–403 K and 4.0–7.5 MPa hydrogen pressure. The d -glucose concentration was varied between 0.56 and 1.39 mol/l. The kinetic experiments were carried out in the absence of mass transport limitations, which was verified by using measured gas–liquid mass transfer coefficients and estimated diffusion and liquid–solid mass transfer coefficients. Many literature reports suffer from transport limitations. In the operating regime studied the reaction rate showed a first order dependency with respect to hydrogen. A shift in the order of d -glucose was observed. At low d -glucose concentrations (up to ca. 0.3 mol/l) the reaction showed a first order dependency, while at higher concentrations this changed to zero order behavior. No inhibition by sorbitol or mannitol was observed. The kinetic data were modeled using three plausible rate models based on Langmuir–Hinshelwood–Hougen–Watson (LHHW) kinetics assuming that the surface reaction is rate-determining. Model 1 involves non-competitive adsorption of hydrogen and d -glucose. Hydrogen adsorption is either molecular or dissociative, but due to the weak adsorption it results in both cases in a linear hydrogen pressure dependency, i.e. the same rate expression; Model 2 is based on competitive adsorption of molecular hydrogen and d -glucose; and Model 3 assumes competitive adsorption of dissociatively chemisorbed hydrogen and d -glucose. All three models described the data satisfactorily and further statistic discrimination between these models was not possible.

Journal ArticleDOI
TL;DR: In this paper, the formation rate and reaction rate of ammonium sulfate salts on V 2 O 5 /AC catalyst during selective catalytic reduction (SCR) of NO with NH 3 at low temperatures were studied using elemental analysis, transient response, and TPR methods.

Journal ArticleDOI
TL;DR: The solubility of hydrogen and the corresponding Henry coefficients for 11 ionic liquids have been determined in situ at 100 atm H2 pressure and are much lower than expected and correlate with the rate of reaction for the hydrogenation of benzene to cyclohexane in these solvents.

Journal ArticleDOI
TL;DR: In this paper, the authors describe the mathematical relationship between observed reaction rates and the loading of the supported phase and show that the reaction rate shows a sharp maximum with increasing loading and at high loadings the rate is reduced to zero, due to the lack of active sites (the support is completely covered).

Journal ArticleDOI
TL;DR: In this article, the water-soluble Ru(II)-phosphine complex [RuCl2(mTPPMS)(2)}(2)] was found a suitable catalyst for the hydrogenation of NaHCO3 to NaH CO2 in aqueous solution under mild conditions with catalyst turnover frequencies (TOFs) in the range of 35-50 h(-1) at 50degreesC and 10 bar total pressure.
Abstract: The water-soluble Ru(II)-phosphine complex, [{RuCl2(mTPPMS)(2)}(2)] was found a suitable catalyst for the hydrogenation of NaHCO3 to NaHCO2 in aqueous solution under mild conditions with catalyst turnover frequencies (TOFs) in the range of 35-50 h(-1) at 50degreesC and 10 bar total pressure. The suggested reaction mechanism involves the formation of Ru(II)-hydrides of the general formula [RuHX(mTPPMS)(4)] where X = H-, HCO3- or HCO2-. At 80degreesC and 95 bar total pressure, the reduction of NaHCO3 proceeded with high reaction rate (9600 h(-1)) hitherto unobserved in purely aqueous solutions. The reactions do not require the presence of organic amine additives, however, the addition of quinoline increased the rate considerably. Aqueous suspensions of calcium carbonate could also be hydrogenated with CO2/H-2 gas mixtures. (C) 2003 Elsevier B.V. All rights reserved.

Journal ArticleDOI
TL;DR: In this paper, a fractional reaction-diffusion equation is derived from a continuous time random walk model when the transport is dispersive, and the recombination is shown to depend on the intrinsic reaction rate.
Abstract: A fractional reaction-diffusion equation is derived from a continuous time random walk model when the transport is dispersive. The exit from the encounter distance, which is described by the algebraic waiting time distribution of jump motion, interferes with the reaction at the encounter distance. Therefore, the reaction term has a memory effect. The derived equation is applied to the geminate recombination problem. The recombination is shown to depend on the intrinsic reaction rate, in contrast with the results of Sung et al. [J. Chem. Phys. 116, 2338 (2002)], which were obtained from the fractional reaction-diffusion equation where the diffusion term has a memory effect but the reaction term does not. The reactivity dependence of the recombination probability is confirmed by numerical simulations.

Journal ArticleDOI
TL;DR: Results show that the combination of H( 2)O(2) and granular activated carbon (GAC) did increase the total removal of 4-CP than that by single GAC adsorption and the dechlorination efficiency was independent of the 4- CP concentration in aqueous phase.

Book ChapterDOI
TL;DR: The use of micro-emulsions as media for organic reactions is a way to overcome the reagent incompatibility problems that are frequently encountered in organic synthesis as discussed by the authors, which can also be regarded as an alternative to phase transfer catalysis.
Abstract: The use of microemulsions as media for organic reactions is a way to overcome the reagent incompatibility problems that are frequently encountered in organic synthesis. In this sense, microemulsions can be regarded as an alternative to phase transfer catalysis. The microemulsion approach and the phase transfer approaches can also be combined, i.e. the reaction can be carried out in a microemulsion in the presence of a small amount of phase transfer agent. A very high reaction rate may then be obtained. The reaction rate in a microemulsion is often influenced by the charge at the interface and this charge depends on the type of surfactant used. For instance, reactions involving anionic reactants may be accelerated by cationic surfactants. The surfactant counterion also plays a major role for the reaction rate. The highest reactivity is obtained with small counterions, such as acetate, that are only weakly polarizable. Large polarizable anions, such as iodide, bind strongly to the interface and may prevent other anionic species to reach the reaction zone.

Journal ArticleDOI
TL;DR: In this article, the kinetics of the oxidative degradation of indigo carmine (IC) dye with hydrogen peroxide catalyzed with the supported metal complexes have been investigated, and a reaction mechanism was proposed with the formation of free radicals as reactive intermediates.
Abstract: The kinetics of the oxidative degradation of the indigo carmine (IC) dye (disodium salt of 3,3-dioxobi-indolin-2,2-ylidine-5,5-disulfonate) with hydrogen peroxide catalyzed with the supported metal complexes have been investigated. The complexes used are [Cu(amm)4]2+, [Co(amm)6]2+, [Ni(amm)6]2+, [Cu(en)2]2+, and [Cu(ma)4]2+ (amm=ammonia, en=ethylenediamine, and ma=methylamine). Silica, alumina, silica-alumina (25% Al2O3), and cation-exchange resins (Dowex-50W, 2 and 8% DVB) are used as supports. The reaction is first order with respect to [IC] while the order with respect to [H2O2] was dependent on the initial concentration and the type of the catalyst used. At lower [H2O2]0 the order was first, which then decreases with increasing [H2O2]0, finally reaching zero. This aspect is consistent with the formation of a colored peroxo-complex on the catalysts surface. The reactivity of catalysts is dependent on the redox potential of the metal ions, the amount of complex loaded per gram of dry catalyst, the type of ligand, and the support. Moreover, the reaction rate was strongly dependent on the pH of the medium, the cationic and anionic surfactants, and the irradiation with UV-light. The reaction is enthalpy controlled as confirmed from the isokinetic relationship. A reaction mechanism was proposed with the formation of free radicals as reactive intermediates.

Journal ArticleDOI
TL;DR: The observed reaction rates are too slow for immediate use in remediation system design, but the findings may provide a basis for future development of cost-effective abiotic perchlorate removal techniques.
Abstract: The rate and extent of perchlorate reduction on several types of iron metal was studied in batch and column reactors. Mass balances performed on the batch experiments indicate that perchlorate is initially sorbed to the iron surface, followed by a reduction to chloride. Perchlorate removal was proportional to the iron dosage in the batch reactors, with up to 66% removal in 336 h in the highest dosage system (1.25 g mL(-1)). Surface-normalized reaction rates among three commercial sources of iron filings were similar for acid-washed samples. The most significant perchlorate removal occurred in solutions with slightly acidic or near-neutral initial pH values. Surface mediation of the reaction is supported by the absence of reduction in batch experiments with soluble Fe2+ and also by the similarity in specific reaction rate constants (kSA) determined for three different iron types. Elevated soluble chloride concentrations significantly inhibited perchlorate reduction, and lower removal rates were observed for iron samples with higher amounts of background chloride contamination. Perchlorate reduction was not observed on electrolytic sources of iron or on a mixed-phase oxide (Fe3O4), suggesting that the reactive iron phase is neither pure zerovalent iron nor the mixed oxide alone. A mixed valence iron hydr(oxide) coating or a sorbed Fe2+ surface complex represent the most likely sites for the reaction. The observed reaction rates are too slow for immediate use in remediation system design, but the findings may provide a basis for future development of cost-effective abiotic perchlorate removal techniques.

Journal ArticleDOI
TL;DR: In this article, the detailed kinetics of the Fischer−Tropsch synthesis over an industrial Fe−Mn catalyst were studied in a continuous integral fixed-bed reactor under the conditions relevant to industrial operations.
Abstract: The detailed kinetics of the Fischer−Tropsch synthesis over an industrial Fe−Mn catalyst was studied in a continuous integral fixed-bed reactor under the conditions relevant to industrial operations [temperature, 540−600 K; pressure, 1.0−3.0 MPa; H2/CO feed ratio, 1.0−3.0; space velocity, (1.6−4.2) × 10-3 Nm3 kg of catalyst-1 s-1]. Reaction rate equations were derived on the basis of the Langmuir−Hinshelwood−Hougen−Watson type models for the Fischer−Tropsch reactions and the water-gas-shift reaction. Kinetic model candidates were evaluated by the global optimization of kinetic parameters, which were realized by first minimization of multiresponse objective functions with a genetic algorithm approach and second optimization with the conventional Levenberg−Marquardt method. It was found that an alkylidene mechanism based model could produce a good fit of the experimental data. This model shows that the desorption of the products and the insertion of methylene into the metal−alkylidene bond are the rate-dete...

Journal ArticleDOI
TL;DR: In this article, a mathematical model based on reaction rate expressions to describe the displacement of methane conversion in the steam reforming was presented, and the effect of several parameters including weight hourly space velocity (WHSV), load-to-surface ratio, reaction pressure, hydrogen partial pressure in permeate side and reaction temperature were investigated.

Journal ArticleDOI
TL;DR: In this article, the reaction rate of coke formation and its influence over catalyst deactivation was studied over a Cr 2 O 3 /Al 2 O O 3 catalyst over the temperature range of 525-575°C at atmospheric pressure, and a Langmuir-Hinshelwood mechanism provided the best fit for the reaction.
Abstract: The kinetics of propane dehydrogenation to produce propene over a Cr 2 O 3 /Al 2 O 3 catalyst has been investigated over the temperature range of 525–575 °C at atmospheric pressure. The reaction rate of coke formation and its influence over catalyst deactivation has been studied. A Langmuir–Hinshelwood mechanism provides the best fit for the reaction, while a monolayer–multilayer mechanism is proposed to model the coke growth. Furthermore, this model was able to predict coke formation under conditions far from those employed in the experiments used to obtain the kinetics.

Journal ArticleDOI
TL;DR: In this paper, the first order kinetics in hexacyanoferrate(III) and alkali concentrations and an order of less than unity in sulfanilic acid concentration (SAA) were studied spectrophotometrically.

Journal ArticleDOI
TL;DR: In this article, a multistep methodology was applied to construct a C1 surface reaction mechanism for methane oxidation on platinum, which is capable of capturing the physics of methane oxidation over a wide range of operating conditions.

Journal ArticleDOI
TL;DR: Results suggest that any calcium present in fly ash can react with arsenic vapor and capture the metal in water-insoluble forms of the less hazardous As(V) oxidation state.

Journal ArticleDOI
TL;DR: A computational model was developed using these kinetic parameters to predict which LDL targets are oxidized with varying molar excesses of HOCl, in both the absence and the presence of added ascorbate.
Abstract: Oxidation of low-density lipoproteins (LDL) is believed to contribute to the increased uptake of LDL by macrophages, which is an early event in atherosclerosis. Hypochlorous acid (HOCl) has been implicated as one of the major oxidants involved in these processes. In a previous study, the rates of reaction of HOCl with the reactive sites in proteins were investigated (Pattison, D. I., and Davies, M. J. (2001) Chem. Res. Toxicol. 14, 1453−1464). The work presented here expands on those studies to determine absolute second-order rate constants for the reactions of HOCl with various lipid components and antioxidants in aqueous solution (pH 7.4). The reactions of HOCl with phosphoryl-serine and phosphoryl-ethanolamine are rapid (k ∼ 105 M-1 s-1) and of comparable reactivity to many of the protein sites. The major products formed in these reactions are chloramines, which decay to give both nitrogen- and carbon-centered radicals. Subsequent reactions of these species may induce oxidation of the LDL lipid compone...

Journal ArticleDOI
TL;DR: In this paper, an increase of the adsorbed methanolic residue stripping charge is observed with the increase of catalyst dispersion, and the stripping peak potential shifts more negatively accounting for a lower activation barrier for the reaction.