scispace - formally typeset
Search or ask a question

Showing papers on "Reaction rate constant published in 1996"


Journal ArticleDOI
TL;DR: The properties of iron metal that make it useful in remediation of chlorinated solvents may also lead to reduction of other groundwater contaminants such as nitro aromatic compounds (NACs) as mentioned in this paper.
Abstract: The properties of iron metal that make it useful in remediation of chlorinated solvents may also lead to reduction of other groundwater contaminants such as nitro aromatic compounds (NACs). Nitrobenzene is reduced by iron under anaerobic conditions to aniline with nitrosobenzene as an intermediate product. Coupling products such as azobenzene and azoxybenzene were not detected. First-order reduction rates are similar for nitrobenzene and nitrosobenzene, but aniline appearance occurs more slowly (typical pseudo-first-order rate constants 3.5 × 10-2, 3.4 × 10-2, and 8.8 × 10-3 min-1, respectively, in the presence of 33 g/L acid-washed, 18−20 mesh Fluka iron turnings). The nitro reduction rate increased linearly with concentration of iron surface area, giving a specific reaction rate constant (3.9 ± 0.2 × 10-2 min-1 m-2 L). The minimal effects of solution pH or ring substitution on nitro reduction rates, and the linear correlation between nitrobenzene reduction rate constants and the square-root of mixing ra...

709 citations


Journal ArticleDOI
TL;DR: The results support that peroxynitrite anion rapidly reacts with carbon dioxide to yield an adduct (ONOOCO2-) which can participate in oxidation and nitration processes, thus redirecting the primary reactivity of peroxysitrite.

550 citations


Journal ArticleDOI
TL;DR: In this paper, an alternative strategy to functionalize HF-etched Si surfaces involving halogenation and subsequent reaction with alkyl Grignard or alkyal lithium reagents was proposed.
Abstract: Despite the fact that H-terminated, HF-etched Si crystals are the starting point for construction of most contemporary electronic devices,1 little is known about the chemical reactions of H-terminated Si surfaces under ambient temperature and pressure.2,3 Functionalization of Si without partial oxidation and/or formation of electrical defects is potentially important in fabricating improved electronic devices4,5 as well as in measurement of charge transfer rate constants at semiconductor/ liquid contacts.6 One recently described approach involves the reaction of HF-etched Si(111) with olefins and organic diacyl peroxides, in which formation of a self-assembled (near)monolayer of Si-alkyls was hypothesized.2 We report here an alternative strategy to functionalize HF-etched Si surfaces involving halogenation and subsequent reaction with alkyl Grignard or alkyl lithium reagents. We report vibrational spectroscopic and temperature programmed desorption data which confirm that the alkyl groups are bonded covalently to the Si surface, and we demonstrate that such overlayer formation can impede the undesirable oxidation of Si in a variety of environments while providing surfaces of high electrical quality. The H-terminated Si surface7 was first exposed to PCl5 for 20-60 min at 80-100 °C, in chlorobenzene with benzoyl peroxide as the radical initiator.8,9 Upon chlorination, the XP survey spectra (Figure 1) showed peaks at 270.2 ( 0.4 binding electron volts, BeV, (Cl 2s) and 199.3 ( 0.4 BeV (Cl 2p), indicating that this procedure yielded Cl on the surface. The high-resolution XP spectrum of the Si 2p peak of this surface displayed, in addition to the substrate Si signal, an additional peak located at 0.98 ( 0.12 BeV higher in binding energy (Figure 2) whose position and intensity was consistent with the formation of a surface Si-Cl bond.10 Auger electron spectra

432 citations


Journal ArticleDOI
TL;DR: In this paper, a semitransparent TiO2 film with extraordinarily high photocatalytic activity was prepared on a glass substrate by sintering aTiO2 sol at 450 °C.
Abstract: A semitransparent TiO2 film with extraordinarily high photocatalytic activity was prepared on a glass substrate by sintering a TiO2 sol at 450 °C. Crystallographic analysis by X-ray diffraction and Raman spectroscopy showed that the film was purely anatase. The photocatalytic properties of the film were investigated by measuring the photodegradative oxidation of gaseous acetaldehyde at various concentrations under strong and weak UV light irradiation conditions. The photocatalytic activity of the film was higher than that of one of the most active commercial TiO2 powders, Degussa P-25. The kinetics of acetaldehyde degradation as catalyzed by the TiO2 film as well as by P-25 powder were analyzed in terms of the Langmuir-Hinshelwood model. It is shown that the number of adsorption sites per unit true surface area is larger with the TiO2 film, as analyzed in the powder form, than with P-25 powder. Meanwhile, the first-order reaction rate constant is also much larger with the film than with P-25 powder. Moreover, under most experimental conditions, particularly with high concentrations of acetaldehyde and weak UV illumination intensity, the quantum efficiency was found to exceed 100% on an absorbed-photon basis, assuming that only photo-generated holes play a major role in the reaction. This leads to the conclusion that the photodegradative oxidation of acetaldehyde is not mediated solely by hydroxyl radicals, generated via hole capture by surface hydroxyl ions or water molecules, but also by photocatalytically generated superoxide ion, which can be generated by the reduction of adsorbed oxygen with photogenerated electrons.

413 citations


Journal ArticleDOI
TL;DR: This study used a stopped-flow kinetic technique monitored by UV−VIS spectroscopy and an initial rate method to ascertain the rate constant and the rate dependence of each reactant, and observed that Cr(VI) removal conformed to −d[Cr(VI)]/dt = kcr[Fe(II)]0.6.
Abstract: Reduction of Cr(VI) to Cr(III) is environmentally favorable as the latter species is not toxic to most living organisms and also has a low mobility and bioavailability. Ferrous iron is one possible reductant implicated as a major contributor to the removal of Cr(VI) from suboxic and anoxic waters and soils. Despite the importance of this redox reaction, no mechanistic or kinetic information are available, which are needed to determine the rate of Cr(VI) reduction and to assess the role of oxygen in limiting this reaction. In this study we used a stopped-flow kinetic technique monitored by UV−VIS spectroscopy and an initial rate method to ascertain the rate constant and the rate dependence of each reactant. We observed that the rate of Cr(VI) removal conformed to −d[Cr(VI)]/dt = kcr[Fe(II)]0.6[Cr(VI)]1 where kcr = 56.3 (±3.7) mmol-0.6 min-1 L0.6. Based on this rate expression and that for the oxygenation of Fe(II), Cr(VI) reduction should be unaffected by oxygen except at pH values in excess of 8 even at m...

397 citations


Journal ArticleDOI
TL;DR: Experimental evidence is provided that the most stable form of singly protonated bradykinin is a salt-bridge structure, and the presence of a second charge can dramatically influence the dissociation dynamics of these ions.
Abstract: Blackbody infrared radiative dissociation (BIRD) spectra of singly and doubly protonated bradykinin and its analogues are measured in a Fourier-transform mass spectrometer. Rate constants for dissociation are measured as a function of temperature with reaction delays up to 600 s. From these data, Arrhenius activation parameters in the zero-pressure limit are obtained. The activation parameters and dissociation products for the singly protonated ions are highly sensitive to small changes in ion structure. The Arrhenius activation energy (Ea) and pre-exponential (or frequency factor, A) of the singly protonated ions investigated here range from 0.6 to 1.4 eV and 105 to 1012 s-1, respectively. For bradykinin and its analogues differing by modification of the residues between the two arginine groups on either end of the molecule, the singly and doubly protonated ions have average activation energies of 1.2 and 0.8 eV, respectively, and average A values of 108 and 1012 s-1, respectively, i.e., the presence of ...

378 citations


Journal ArticleDOI
TL;DR: In this article, the role of Fenton's reaction [oxidation of Fe(II) by hydrogen peroxide] in the catalytic cycle was examined, and a number of competing processes were observed in model systems.
Abstract: Iron redox cycling can catalyze the oxidation of humic substances and increase the rate of oxygen consumption in surface waters rich in iron and organic carbon. This study examines the role of Fenton`s reaction [oxidation of Fe(II) by hydrogen peroxide] in this catalytic cycle. A number of competing processes were observed in model systems containing dissolved Fe, hydrogen peroxide, and Suwannee River fulvic acid. First, the effective rate constant of Fenton`s reaction increased with increasing fulvic acid concentration, indicating the formation Fe(II)-fulvate complexes that react more rapidly with hydrogen peroxide than Fe(II)-aquo complexes. This effect was significant at pH 5 but negligible at pH 3. A second effect was scavenging of the HO{sup .} radical produced in Fenton`s reaction by fulvic acid, forming an organic radical. The organic radical reduced oxygen to HO{sub 2}{sup .}/O{sub 2}{sup .-}, which then regenerated hydrogen peroxide by reaction with Fe(II). Finally, Fe(III) was reduced by a dark reaction with fulvic acid, characterized by an initially fast reduction followed by slower processes. The behavior of Fe(II) and hydrogen peroxide over time in the presence of fulvic acid and oxygen could be described by a kinetic model taking all of these reactions into account. The netmore » result was an iron redox cycle in which hydrogen peroxide as well as oxygen were consumed (even though direct oxidation of Fe(II) by oxygen was not significant), and the oxidation of fulvic acid was accelerated. 56 refs., 7 figs., 1 tab.« less

354 citations


Journal ArticleDOI
TL;DR: A critical review of the basic data concerning the physics and chemistry of low pressure SiH 4 glow discharges used to deposit hydrogenated amorphous silicon films (a-Si:H) is presented in this article.
Abstract: This paper presents a critical review of the basic data concerning the physics and chemistry of low pressure SiH 4 glow discharges used to deposit hydrogenated amorphous silicon films (a-Si:H). Starting with an updated table of thermochemical data, we analyze the gas-phase elementary processes consisting of i) electron-molecule collisions, ii) ion-molecule collisions, iii) neutral-neutral collisions, iv) other electron and ion collisions involving electron-ion and ion-ion recombination, electron attachment on radicals and detachment of anions, and v) cluster growth kinetics in dusty plasmas. Experimental data or theoretical estimates are given and discussed in terms of cross-sections. collision and reaction rate constants, and transport coefficients. We also analyze the surface processes and reaction probabilities of ions, radicals and molecules.

308 citations


Journal ArticleDOI
TL;DR: Peroxynitrite ion (ONO2-) reacted rapidly with CO2 to form a short-lived intermediate provisionally identified as the ONO2CO2- adduct, which produced 3-nitrotyrosine and 3,3'-dityrosine as the major oxidation products in tyrosine oxidation.
Abstract: Peroxynitrite ion (ONO2-) reacted rapidly with CO2 to form a short-lived intermediate provisionally identified as the ONO2CO2- adduct. This adduct was more reactive in tyrosine oxidation than ONO2- itself and produced 3-nitrotyrosine and 3,3'-dityrosine as the major oxidation products. With tyrosine in excess, the rate of 3-nitrotyrosine formation was independent of the tyrosine concentration and was determined by the rate of formation of the ONO2CO2- adduct. The overall yield of oxidation products was also independent of the concentration of tyrosine and medium acidity; approximately 19% of the added ONO2- was converted to products under all reaction conditions. However, the 3-nitrotyrosine/3,3'-dityrosine product ratio depended upon the pH, tyrosine concentration, and absolute reaction rate. These data are in quantitative agreement with a reaction mechanism in which the one-electron oxidation of tyrosine by ONO2CO2- generates tyrosyl and NO2 radicals as intermediary species, but are inconsistent with mechanisms that invoke direct electrophilic attack on the tyrosine aromatic ring by the adduct. Based upon its reactivity characteristics, ONO2CO2- has a lifetime shorter than 3 ms and a redox potential in excess of 1 V, and oxidizes tyrosine with a bimolecular rate constant greater than 2 x 10(5) M-1 s-1. In comparison, in CO2-free solutions, oxidation of tyrosine by peroxynitrite was much slower and gave significantly lower yields (approximately 8%) of the same products. When tyrosine was the limiting reactant, 3,5-dinitrotyrosine was found among the reaction products of the CO2-catalyzed reaction, but this compound was not detected in the uncatalyzed reaction.

288 citations


Journal ArticleDOI
TL;DR: In this article, the rate of hydrogen sulfide by hydrous Fe(III) oxides has been determined as a function of pH (4.0-8.5), temperature (5-45 °C) and ionic strength (0-4 m) in seawater.

277 citations


Journal ArticleDOI
TL;DR: In this paper, a re-evaluation of kinetic data showed that methoxide base-catalyzed methanolysis of soybean oil at 40°C (6:1 methanol:oil molar ratio) to form methyl esters proceeds approximately 15 times more slowly than butanolysis at 30°C.
Abstract: A re-evaluation of kinetic data shows that the methoxide base-catalyzed methanolysis of soybean oil at 40°C (6:1 methanol:oil molar ratio) to form methyl esters proceeds approximately 15 times more slowly than butanolysis at 30°C. This is interpreted to be the result of a two-phase reaction in which methanolysis occurs only in the methanol phase. Low oil concentration in methanol causes the slow reaction rate; a slow dissolving rate of the oil in the methanol causes an initiation period. Intermediate mono- and diglycerides preferentially remain in the methanol, and react further, thus explaining the deviation from second-order kinetics. The same explanations apply for hydroxide ion catalyzed methanolysis. At the 6:1 methanol:oil molar ratio the addition of a cosolvent, such as 1.25 volumes of tetrahydrofuran (THF) per volume of methanol, produces an oil-dominant one-phase system in which methanolysis speeds up dramatically and occurs as fast as butanolysis. The critical separation of the glycerol-rich phase still occurs and does so faster than in the cosolvent-free system. For THF, recycle of solvent is simplified because of the similar boiling points of THF (67°C) and methanol (65°C). Possible explanations for the abnormal slowing of the methanolysis reactions are presented in terms of (1) lower rate constants for mono- and diglyceride reactions due to the formation of cyclic intermediates, (2) a fall in the polarity of the reaction mixture due to either methanol depletion or mixing of the oil, methanol and cosolvent, and (3) depletion of hydroxide ion when this is present.

Journal ArticleDOI
TL;DR: In this article, the irreversible denaturation of β-lactoglobulins A and B and α -lactalbumin were determined. But the authors focused on the temperature dependence of the denaturation process and not the initial protein concentration.
Abstract: Reconstituted whole milk was heated using pilot-scale heating equipment. Kinetic and thermodynamic parameters for the irreversible denaturation of β-lactoglobulins A and B and α-lactalbumin were determined. α-Lactalbumin denaturation was first order, whereas both β-lactoglobulin variants had a reaction order of 1.5. Arrhenius plots for all three proteins showed an abrupt change in temperature dependence. In the low-temperature range, the thermodynamic parameters were ascribed to typical denaturation processes in which the unfolding of the protein tertiary structure is the rate-determining step. At higher temperatures, these parameters were in the range expected for typical condensation reactions, suggesting that aggregation processes may be rate-determining in this temperature range. The rate constants for β-lactoglobulin denaturation were independent of the initial protein concentration at all temperatures. For α-lactalbumin at temperatures below 85 °C the rate constants may have been dependent on the in...

Journal ArticleDOI
TL;DR: In this article, a dual-level generalized transition state theory and statistical calculations based on correlated electronic structure calculations with augmented correlated basis sets are used to predict rate constants and deuterium kinetic isotope effects for the competing SN2 and E2 reactions of ClO-with C2H5Cl.
Abstract: Dual-level generalized transition state theory and statistical calculations based on correlated electronic structure calculations with augmented correlated basis sets are used to predict rate constants and deuterium kinetic isotope effects for the competing SN2 and E2 reactions of ClO- with C2H5Cl. The E2 reaction is favored by zero point effects despite its higher barrier. Furthermore, the entropic contribution of low-frequency transition state modes promotes the E2 reaction by about an order of magnitude at 300 K, despite nearly equal combined electronic/zero point propensities for the two reactions. The kinetic isotope effect for the E2 reaction is predicted to be 8.3 at 100 K and 3.1 at room temperature. The kinetic isotope effect of the SN2 reaction is predicted to be surprisingly low at low temperatures (0.06 at 100 K) and 0.60 at room temperature. The activation energy for both reactions is predicted to be negative below about 900 K and positive above 1000 K. Variational effects on the location of ...

Journal ArticleDOI
TL;DR: In this paper, reaction models for the formation and decomposition of metal hydrides with one rate determining partial reaction step are presented, based on reasonable physical assumptions for each of the partial steps.

Journal ArticleDOI
TL;DR: In this paper, Terephthalate dosimeter is used in sonochemical studies and the underlying chemistry have been elucidated using ionizing radiation techniques, and the products were studied after γ-radiolysis and the kinetics were followed by pulse radiolysis.

Journal ArticleDOI
TL;DR: The simulations suggest that unexpected deviations from ideal binding progress curves may be due to highly transport influenced binding kinetics, and the use of a thinner polymer matrix could significantly increase the range of detectable rate constants.

Journal ArticleDOI
TL;DR: It is shown here that the deuterium kinetic isotope effect (KIE) for the reaction of compound Q with CH4 is 50-100, which is one of the largest effects observed to date, and a modified mechanism introducing the possibility of hydrogen-atom reabstraction by an intermediate methyl radical is proposed.
Abstract: The reduced hydroxylase component (MMOH) of soluble methane monooxygenase (MMO) from Methylosinus trichosporium OB3b reacts with O2 and CH4 to produce CH3OH and H2O in a single-turnover reaction. Transient kinetic analysis of this reaction has revealed at least five and probably six intermediates during the turnover [Lee, S.-K., Nesheim, J. C., & Lipscomb, J. D. (1993) J. Biol. Chem. 268, 21569-21577; Liu, Y., Nesheim, J. C., Lee, S.-K., & Lipscomb, J. D. (1995) J. Biol. Chem. 270, 24662-24665]. One intermediate, termed compound Q, reacts with CH4 to yield enzyme-bound product. It is shown here that the deuterium kinetic isotope effect (KIE) for the reaction of compound Q with CH4 is 50-100, which is one of the largest effects observed to date. The rate constants for the reactions of the deuterated homologs of methane decrease monotonically as the deuterium content increases, suggesting that a large primary isotope effect dominates. The KIEs determined by analyzing the products after a single turnover have the following values: 1:1 CH4:CD4 (19); CD3H (12); CD2H2 (9); and CH3D (4). The KIE values determined by directly observing the reactive intermediate and by monitoring product ratios are all large, consistent with complete C-H bond breaking in the oxygenation step of the reaction. However, the differences in the KIE values determined by these two methods suggest that the reaction is more complex than currently proposed. A modified mechanism introducing the possibility of hydrogen-atom reabstraction by an intermediate methyl radical is proposed.

Journal ArticleDOI
TL;DR: A new kinetic analysis of biosensor data is presented, based upon a phenomenological two-compartment approximate description of transport, which permits determination of the correct values of chemical rate constants that are 10- to 100-fold greater than those that can be correctly evaluated by previous analyses.

Journal ArticleDOI
TL;DR: In this article, an analytical fit of the lowest 3A′ potential energy surface of the N2+O→NO+N reaction based on the CCI ab initio data is obtained.
Abstract: Theoretical determinations of the thermal rate constants and product energy distributions of the N2+O→NO+N reaction, which plays a crucial role in hydrocarbon air combustion and high temperature air chemistry, are carried out using a quasiclassical trajectory method. An analytical fit of the lowest 3A′ potential energy surface of this reaction based on the CCI ab initio data is obtained. The trajectory study is done on this surface and an analytical 3A″ surface proposed by Gilibert et al. [J. Chem. Phys. 97, 5542 (1992)]. The thermal rate constants computed from 3000 to 20 000 K are in good agreement with the available experimental data. In addition, the dependence of the rate constant on the N2 internal state is studied. It is found that a low vibrational excitation can reduce the rate constant of this reaction by a factor of 3. Also, we investigate the effect of the N2 vibrational state on the product NO vibrational distribution, and it is found that at low N2 vibrational states, the NO vibrational dist...

Journal ArticleDOI
TL;DR: The mechanism of the proton-coupled electron transfer reaction, + H+ → QA(QBH)- was studied in reaction centers (RCs) from the photosynthetic bacterium Rb. sphaeroides by substituting quinones with different redox potentials into the QA site as discussed by the authors.
Abstract: The mechanism of the proton-coupled electron transfer reaction, + H+ → QA(QBH)- was studied in reaction centers (RCs) from the photosynthetic bacterium Rb. sphaeroides by substituting quinones with different redox potentials into the QA site. These substitutions change the driving force for electron transfer without affecting proton transfer rates or proton binding equilibria around the QB site. The measured rate constants, increased with increasing electron driving force (by a factor of 10 per 160 meV change in redox free energy). The proton-coupled electron transfer was modeled by (i) four possible two-step mechanisms in which electron transfer can precede or follow proton transfer and can be either the rate determining or fast step in the overall reaction and (ii) a one-step mechanism involving the concerted transfer of an electron and a proton. The free energy dependencies of these possible mechanisms were predicted using Marcus theory and were compared to the observed dependence. The two stepwise mec...

Journal ArticleDOI
TL;DR: In this article, a simple model for the vertical concentration gradient of PCBs in the troposphere was developed and used to calculate the total global loss rate (8300 t yr-1) of polychlorinated biphenyls from the atmosphere due to removal by OH.
Abstract: Polychlorinated biphenyls (PCBs) are transported through the environment primarily in the atmosphere and may undergo chemical reactions, particularly with the OH radical, while in the vapor phase. Rate constants for the gas-phase reaction of 14 PCB congeners with the OH radical over the temperature range of 323−363 K were measured. The calculated temperature dependences of the reactions were used to estimate OH−PCB reaction rate constants at 298 K; see Table 4. These 298 K rate constants agree well with literature rate constants for PCBs containing zero to two chlorines and with OH−PCB reaction rate constants estimated from a structure−activity method. Calculated atmospheric lifetimes of PCBs due to OH-initiated reactions varied from 2 days for biphenyl to 34 days for a pentachlorobiphenyl. A simple model for the vertical concentration gradient of PCBs in the troposphere was developed and used to calculate the total global loss rate (8300 t yr-1) of PCBs from the atmosphere due to removal by OH. This path...

Journal ArticleDOI
TL;DR: In this article, the effect of Co2+, Ni2+, Cu2+ and Zn2+ ions as Lewis acid catalysts on the rate and endo-exo selectivity of the Diels-Alder reaction between the bidentate dienophiles 3-phenyl-1-(2-pyridyl)-2-propen-1-ones (1a-e) and cyclo-pentadiene (2) in water has been studied.
Abstract: Here we report the first detailed study of a Diels-Alder (DA) reaction that is catalyzed by Lewis acids in water. The effect of Co2+, Ni2+, Cu2+ and Zn2+ ions as Lewis acid catalysts on the rate and endo-exo selectivity of the DA reaction between the bidentate dienophiles 3-phenyl-1-(2-pyridyl)-2-propen-1-ones (1a-e) and cyclo-pentadiene (2) in water has been studied. Relative to the uncatalyzed reaction in acetonitrile, catalysis by 0.010 M CU(NO3)(2) in water accelerates the reaction by a factor of 79 300. The kinetics of the catalyzed reaction were analyzed in terms of equilibrium constants for complexation of the Lewis acid with 1a-e and rate constants for the reaction of the resulting complexes with 2. The rate enhancement imposed upon the uncatalyzed DA reaction of substrates 1 with 2 by water is much more pronounced than that for the catalyzed reaction. The increase of the endo-exo selectivity induced by water in the uncatalyzed process is completely absent for the Lewis acid catalyzed reaction. The modest solvent and substituent effects observed for the catalyzed reaction indicate that the change in charge separation during the activation process is not larger than the corresponding change For the uncatalyzed reaction.

Journal ArticleDOI
TL;DR: This novel reaction constitutes a potential cellular defense line against peroxynitrite, one of the important reactive species in inflammatory processes and reduction of the selenoxide in cells to regenerate ebselen would allow for a sustained defense againstperoxynite.
Abstract: Ebselen, 2-phenyl-1,2-benzisoselenazol-3(2H)-one, rapidly reacts with peroxynitrite, the rate constant being of the order of 10(6) M-1 s-1; the reaction yields the selenoxide of the parent molecule, 2-phenyl-1,2-benzisoselenazol-3(2H)-one 1-oxide, as the sole selenium-containing product; a stoichiometry of 1 mol of ebselen reacted and of the selenoxide formed per mole of peroxynitrite was observed. The reaction was studied in detail at neutral and alkaline pH (pH 10-11). It also proceeds at acidic pH where peroxynitrous acid (ONOOH) is predominant, the yield of the selenoxide being lower because peroxynitrous acid (pKa = 6.8) decays rapidly. Reduction of the selenoxide in cells to regenerate ebselen would allow for a sustained defense against peroxynitrite. This novel reaction constitutes a potential cellular defense line against peroxynitrite, one of the important reactive species in inflammatory processes.

Journal ArticleDOI
TL;DR: The second‐order rate constant for the reaction of ebselen with peroxynitrite (ONOO−) is (2.0±0.1) × 106 M−1s−1 at pH ≥ 8 and 25°C, 3–4 orders of magnitude higher than the rate constants observed for cysteine, ascorbate, or methionine.

Journal ArticleDOI
TL;DR: The rate constants for deprotonation of ethyl acetate by 3-substituted quinuclidines are correlated by β = 1.09 ± 0.05 as mentioned in this paper.
Abstract: The rate constants for deprotonation of ethyl acetate by 3-substituted quinuclidines are correlated by β = 1.09 ± 0.05. The limits of kBH = 2−5 × 109 M-1 s-1 for the encounter-limited reaction of the simple oxygen ester enolate with protonated quinuclidine (pKBH = 11.5) were combined with kB = 2.4 × 10-5 M-1 s-1 for deprotonation of ethyl acetate by quinuclidine, to give pKaK = 25.6 ± 0.5 for ionization of ethyl acetate as a carbon acid in aqueous solution. A rate−equilibrium correlation for proton transfer from methyl and benzylic monocarbonyl compounds to hydroxide ion has been extended by 6 pK units in the thermodynamically unfavorable direction, and it is shown that the absence of curvature of this correlation is inconsistent with a constant Marcus intrinsic barrier for the enolization of simple carbonyl compounds.

Journal ArticleDOI
TL;DR: In this paper, an enthalpy of formation of HOCO of ΔHof,0=−(205±10) kJ mol−1 was derived on the basis of energy and angular-momentum-dependent rates of HO formation.
Abstract: Thermal rate constants of the complex‐forming bimolecular reaction HO+CO■HOCO→H+CO2 were measured between 90 and 830 K in the bath gas He over the pressure range 1–700 bar. In addition, the vibrational relaxation of HO in collisions with CO was studied between 300 and 800 K. HO was generated by laser photolysis and monitored by saturated laser‐induced fluorescence. The derived second‐order rate coefficients showed a pronounced pressure and complicated non‐Arrhenius temperature dependence. Above 650 K, the disappearance of HO followed a biexponential time law, indicating thermal instability of collisionally stabilized HOCO. By analyzing the corresponding results, an enthalpy of formation of HOCO of ΔHof,0=−(205±10) kJ mol−1 was derived. On the basis of energy‐ and angular‐momentum‐dependent rates of HOCO formation, activated complex properties for the addition reaction HO+CO→HOCO were derived from the limiting high‐pressure rate constants; with the limiting low‐pressure rate constants, activated complex pr...

Journal ArticleDOI
TL;DR: In this article, the authors investigated the relationship between the particle growth rate and tetraethylorthosilicate (TEOS) hydrolysis and condensation rate with reaction condition variables, such as temperature, NH3 concentration and H2O concentration.
Abstract: Kinetic studies of the hydrolysis and condensation of tetraethylorthosilicate (TEOS) during the formation of uniform silica particles were performed through determining concentrations of TEOS and silicic acid by means of gas chromatography and a conductometer, respectively. It was shown that both hydrolysis of TEOS and condensation of Si(OH)4 are first order with TEOS and Si(OH)4, respectively, and the relationships of the hydrolysis and condensation rate constants with reaction condition variables, such as temperature, NH3 concentration, and H2O concentration, were determined. In addition, the particle growth rate was investigated with relation to the hydrolysis and condensation kinetics. Experiments showed that, during most of the reaction, the amount of formed particles is less than that of consumed TEOS, indicating that reaction intermediates exist during the process of silica formation. In the early stages of the Stober process, the reaction intermediates include silicic acid and subparticles, while ...

Journal ArticleDOI
TL;DR: In this article, the potential drop across the interface between two immiscible electrolyte solutions (ITIES), φ, can be quantitatively controlled and varied by changing the ratio of concentrations of the potential-determining ion in the two liquid phases.
Abstract: The potential drop across the interface between two immiscible electrolyte solutions (ITIES), φ, can be quantitatively controlled and varied by changing the ratio of concentrations of the potential-determining ion in the two liquid phases. This approach was used to study the potential dependence of the rate constant for electron transfer (ET) at the ITIES (kf) by scanning electrochemical microscopy (SECM) with no external potential bias applied. The Tafel plot obtained for ET between aqueous Ru(CN)64- and the oxidized form of zinc porphyrin in benzene was linear with a transfer coefficient, α = 0.5, determined from the slope of a plot of ln kf vs φ, in agreement with conventional ET theory. The observed change in the ET rate with the interfacial potential drop cannot be attributed to concentration effects and represents the potential dependence of the apparent rate constant. This result is discussed in relation to the interface thickness and structure. The SECM was also used to study solid phase formation...

Journal ArticleDOI
TL;DR: In this paper, it was shown that the suppression of the backward reaction on Pt is essential for overall water splitting into H 2 and O 2, and this is one of the important functions of NaHCO 3 and Na 2 CO 3.
Abstract: The stoichiometric photodecomposition of water into H 2 and O 2 proceeds over ZrO 2 powder suspended in aqueous solutions in the absence of loaded metals such as Pt, RuO 2 and NiO x . The activity is affected significantly by the pressure of the reaction system, the types of additive and the pH of the solution. Specifically, the addition of carbonate, such as NaHCO 3 and Na 2 CO 3 , leads to a remarkable increase in the activity and stability of the gas evolution rate vs. the reactor pressure. The surface of ZrO 2 was covered with carbonate species by the addition of carbonate salt. It is thought that carbonate species on ZrO 2 play an important role in the desorption of O 2 via the carbonate radical. In contrast, the promotion effect of platinum loading on water splitting is negligible. This behaviour can be explained in terms of the backward reaction (H 2 O formation from H 2 and O 2 ) on Pt and the barrier height of the semiconductor-metal junction at which electrons in the conduction band migrate. Furthermore, it is concluded that the suppression of the backward reaction on Pt is essential for overall water splitting into H 2 and O 2 , and this is one of the important functions of NaHCO 3 and Na 2 CO 3 . Pt on ZrO 2 in carbonate solutions exists as an oxide; therefore the effect of the backward reaction is suppressed.

Journal ArticleDOI
TL;DR: In this paper, an experimental and theoretical study on the effect of SO2 on CO oxidation with and without NO present has been carried out in an isothermal quartz flow reactor at atmospheric pressure in the temperature range 800-1300 K. The experimental data supported by model analysis suggest that SO2 primarily reacts with O atoms forming SO3, which is subsequently consumed mainly by reaction with HO2.
Abstract: An experimental and theoretical study on the effect of SO2 on moist CO oxidation with and without NO present has been carried out. The experiments were performed in an isothermal quartz flow reactor at atmospheric pressure in the temperature range 800–1300 K. Inlet concentrations of SO2 ranged from 0 to 1800 ppmv, while the NO ranged between 0, 100, or 1500 ppm. SO2 inhibits CO oxidation under the conditions investigated, shifting the fast oxidation regime 20–40 K towards higher temperatures at 1500 ppm SO2. The inhibition is most pronounced at high O atom levels. The experimental data supported by model analysis suggest that SO2 primarily reacts with O atoms forming SO3, which is subsequently consumed mainly by reaction with O and HO2. Addition of NO significantly diminishes the effect of SO2. Since NO is usually present in combustion flue gases, the impact of SO2 on CO burnout in most practical systems is projected to be small. The H/S/O thermochemistry and reaction subset has been revised based on recent experimental and theoretical results, and a chemical kinetic model has been established. The model provides a reasonable overall description of the effect of SO2 and NO on moist CO oxidation, while the SO3/SO2 ratio is well predicted over the range of conditions investigated. In order to enhance model performance further, rate constants for a number of SO2 and SO3 reactions need to be determined with higher accuracy. © 1996 John Wiley & Sons, Inc.