Your browser doesn't support javascript.
loading
Mostrar: 20 | 50 | 100
Resultados 1 - 20 de 33
Filtrar
Mais filtros










Base de dados
Intervalo de ano de publicação
1.
Inorg Chem ; 63(3): 1625-1632, 2024 Jan 22.
Artigo em Inglês | MEDLINE | ID: mdl-38180901

RESUMO

The aqueous reaction of [IrCl6]2- with CH3SO2- is biphasic and yields a 1:1 mixture of [IrCl6]3- and [IrCl5(H2O)]2- and CH3SO2Cl in the initial rapid phase. The next slow phase corresponds to the hydrolysis of CH3SO2Cl to yield CH3SO3- and Cl-. The initial phase shows kinetic inhibition by [IrCl6]3- that can be minimized by the addition of the radical scavenger propiolic acid. A detailed analysis of the kinetics indicates a mechanism with reversible outer-sphere electron transfer from CH3SO2- to [IrCl6]2- as the first step, followed by the irreversible inner-sphere oxidation of CH3SO2• by [IrCl6]2- to yield [IrCl5(H2O)]2- and CH3SO2Cl. Analysis of the inhibition by [IrCl6]3- and the kinetic effects of propiolic acid enable the determination of the equilibrium constant for the first electron-transfer step. This equilibrium constant then yields E° (CH3SO2•/CH3SO2-) = 1.01 V vs NHE at 25 °C. This is the first report of a standard potential for an alkanesulfonyl radical.

2.
J Phys Chem A ; 127(50): 10600-10612, 2023 Dec 21.
Artigo em Inglês | MEDLINE | ID: mdl-38085654

RESUMO

The aqueous electron affinity and aqueous reduction potentials for F•, Cl•, Br•, I•, OH•, SH•, SeH•, TeH•, ClO•, BrO•, and IO• were calculated using electronic structure methods for explicit cluster models coupled with a self-consistent reaction field (SMD) to treat the aqueous solvent. Calculations were conducted using MP2 and correlated molecular orbital theory up to the CCSD(T)-F12b level for water tetramer clusters and MP2 for octamer cluster. Inclusion of explicit waters was found to be important for accurately predicting the redox potentials in a number of cases. The calculated reduction potentials for X• and ChH• were predicted to within ∼0.1 V of the reported literature values. Fluorine is anomalous due to abstraction of a hydrogen from one of the surrounding water molecules to form a hydroxyl radical and hydrogen fluoride, so its redox potential was calculated using only an implicit model. Larger deviations from experiment were predicted for ClO• and BrO•. These deviations are due to the free energy of solvation of the anion being too negative, as found in the pKa calculations, and that for the neutral being too positive with the current approach.

3.
J Phys Chem A ; 126(49): 9190-9206, 2022 Dec 15.
Artigo em Inglês | MEDLINE | ID: mdl-36469972

RESUMO

Heats of formation and gas phase acidities for the simple acids and their deprotonated anions (A- = F-, Cl-, Br-, I-, OH-, SH-, SeH-, TeH-, OCl-, OBr-, and OI-) were calculated using the Feller-Peterson-Dixon (FPD) method with large basis sets including Douglass-Kroll scalar relativistic corrections. Hydration of the neutral and anionic species was predicted using the supermolecule-continuum approach, resulting in absolute hydration free energies that, when combined with calculated gas phase acidities, produce aqueous acidities and pKa values for these simple acids that are, in general, in excellent agreement with experimental literature values. Absolute hydration free energy values converged quickly in terms of the experimental values for neutral species, requiring only four explicit H2O molecules. HI is anomalous in that it fully dissociates ionically in a water tetramer and was treated without explicit water molecules. The hydration energies of anionic species converged more slowly and were modeled with up to 16 explicit H2O molecules. Calculated values for ΔHf and ΔGgas agree with experimental values within ca. 1.2 kcal/mol, and ΔGaq and ΔΔGhyd agree with experimental values within ca. 2 kcal/mol in most cases.


Assuntos
Água , Termodinâmica , Ânions
4.
Phys Chem Chem Phys ; 24(20): 12541-12549, 2022 May 25.
Artigo em Inglês | MEDLINE | ID: mdl-35579031

RESUMO

It is widely considered that aqueous chlorine atoms (Cl˙) convert to the species HClOH˙ with a half life of about 3 µs and that this species plays an important role in the chemistry of aqueous chlorine atoms. Here it is shown that there is no firm evidence for the existence of HClOH˙ as a species distinct from Cl˙, that the chemistry attributed to HClOH˙ can be accounted for by other well-established species, and that almost all published mechanisms that include reactions of HClOH˙ violate the principle of detailed balancing. More than 100 publications are identified that violate the principle of detailed balancing with HClOH˙ reactions. Proposals for the participation of HClOH˙ in reaction mechanisms originated in studies of the photochemical peroxydisulfate/chloride reaction; here we provide a revised mechanism that omits HClOH˙, complies with the principle of detailed balancing, and has a minimal number of reaction steps.


Assuntos
Cloretos , Cloro , Halogênios
5.
Dalton Trans ; 51(6): 2135-2157, 2022 Feb 08.
Artigo em Inglês | MEDLINE | ID: mdl-35029613

RESUMO

The iron-catalyzed disproportionation of H2O2 has been investigated for over a century, as has been its ability to induce the oxidation of other species present in the system (Fenton reaction). The mechanisms of these reactions have been under consideration at least since 1932. Unfortunately, little or no attention has been paid to ensuring the conformity of the proposed mechanisms and rate constants with the constraints of the principle of detailed balancing. Here we identify more than 200 publications having mechanisms that violate the principle of detailed balancing. These violations occur through the use of incorrect values for certain rate constants, the use of incorrect forms of the rate laws for certain steps in the mechanisms, and the inclusion of illegal loops. A core mechanism for the iron-catalyzed decomposition of H2O2 is proposed that is consistent with the principle of detailed balancing and includes both the one-electron oxidation of H2O2 by Fe(III) and the Fe(II) reduction of HO2˙.

6.
J Phys Chem A ; 124(29): 6084-6095, 2020 Jul 23.
Artigo em Inglês | MEDLINE | ID: mdl-32574051

RESUMO

A supermolecule-continuum approach with water clusters up to n = 16 H2O molecules has been used to predict the absolute hydration free energies at 298 K (ΔGhyd) of both hydrogen (H•) and hydride (H-) to be 4.6 ± 1 and -78 ± 3 kcal/mol, respectively. These values are combined with a high accuracy prediction of the gas-phase electron affinity (ΔGgas,298K = -16.9 kcal/mol) to determine the aqueous electron affinity of H• of 99.5 ± 3 kcal/mol, which yields a reduction potential for H• vs SHE of -0.03 ± 0.15 V. This value is in agreement within 0.2 V with most estimates obtained using a wide variety of approaches. These results can be used to improve the absolute hydricity scale in water which provides additional insights into how a putative hydride interacts with solvent but do not change the ability to predict the relative reactivity of two species using relative hydricity scales.

7.
Environ Sci Technol ; 54(7): 4658-4663, 2020 04 07.
Artigo em Inglês | MEDLINE | ID: mdl-32126765

RESUMO

Advanced oxidation processes tend to have very complex reaction mechanisms, and models containing over 150 steps have been developed to describe the chemistry. Without the aid of automation, it is extremely difficult to avoid the development of kinetic mechanisms that violate the principle of detailed balancing. Here, we apply DETBAL, a computer application, to systematically identify many violations of the principle of detailed balancing in a model proposed for the UV/chloramine process. We then show that these violations can also be found in dozens of other proposed models for advanced oxidation processes. Suggested repairs to these violations are provided. These repairs lead to no significant changes in the model predictions because the illegal loops include steps that are unnecessary under the conditions modeled. The model omits certain steps that do have significant effects on the model predictions.


Assuntos
Cloraminas , Poluentes Químicos da Água , Peróxido de Hidrogênio , Cinética , Oxirredução , Raios Ultravioleta
8.
J Phys Chem A ; 123(47): 10240-10245, 2019 Nov 27.
Artigo em Inglês | MEDLINE | ID: mdl-31693368

RESUMO

Large-scale kinetic models containing more than 50 homogeneous reaction steps have been developed to help understand the behavior of actinide solutions in various circumstances. One specific objective is to understand the behaviors in nitrate solutions as such solutions are used in actinide separations. A challenge arises in developing these large-scale models while ensuring compliance with the principle of detailed balancing because it can be very difficult to identify all embedded reaction loops. These loops can violate the principle of detailed balancing either by consisting of reversible steps that do not meet Wegscheider's condition or by having unopposed irreversible steps that cause illegality. Here we report the development of DETBAL, which is a software code that systematically identifies a basis of reversible loops and a basis of all loops within the model, and then automatically checks for illegal loops. We apply DETBAL to two recent large-scale models of the radiation of nitrate solutions, show that these models violate the principle of detailed balancing in many ways, provide potential solutions to these violations, and demonstrate the effect of some of these modifications.

9.
Inorg Chem ; 58(21): 14752-14759, 2019 Nov 04.
Artigo em Inglês | MEDLINE | ID: mdl-31638784

RESUMO

Methanesulfonyl iodide is produced in aqueous solutions from the reaction of triiodide with methanesulfinate. Dichroic crystals of (CH3SO2I)4·KI3·2I2 are formed from KI/I2 solutions with high concentrations of CH3SO2-, while dichroic crystals of (CH3SO2I)2·RbI3 are formed from RbI/I2 solutions. X-ray crystallography of these two compounds shows that the CH3SO2I molecules coordinate through their oxygen atoms to the metal cations and that the S-I bond length is 2.44 Å. At low concentrations of CH3SO2-, the solutions remain homogeneous and the sulfonyl iodide is formed in a rapid equilibrium: CH3SO2- + I3- ⇌ CH3SO2I + 2I-, KMSI = 1.07 ± 0.01 M at 25 °C (µ = 0.1 M, NaClO4). The sulfonyl iodide solutions display an absorbance maximum at 309 nm with a molar absorptivity of 667 M-1 cm-1. Stopped-flow studies reveal that the equilibrium is established within the dead time of the instrument (∼2 ms). Solutions of CH3SO2I decompose slowly to form the sulfonate: CH3SO2I + H2O → CH3SO3- + I- + 2H+, khyd. In dilute phosphate buffer, this decomposition occurs with khyd = 2.0 × 10-4 s-1; the decomposition rate shows an inverse-squared dependence on [I-] because of the KMSI equilibrium.

10.
J Phys Chem A ; 123(26): 5436-5445, 2019 Jul 05.
Artigo em Inglês | MEDLINE | ID: mdl-31240925

RESUMO

It is not uncommon for proposed complex reaction mechanisms to violate the principle of detailed balancing. Here, we draw attention to three ways in which such violations can occur: reversible reaction loops where the rate constants do not attain closure, illegal loops, and reversible steps having rate equations in the forward and reverse directions that are inconsistent with the equilibrium expressions. We present two simple methods to test whether a proposed mechanism is consistent with the first two aspects of the principle of detailed balancing. Both methods are restricted to closed homogeneous isothermal reactions having mechanisms that consist of stoichiometrically balanced reaction steps. The first method is restricted to mechanisms in which all reaction steps are reversible; values of Δf G° are assigned to all reaction species, equilibrium constants are then computed for all steps, and all rate constants for elementary steps are constrained by the relationship Keq = kf/ kr. The second method is applicable to mechanisms that can consist of a series of reversible and/or irreversible reaction steps. One first examines the subset of reversible steps to determine whether any of these steps are stoichiometrically equivalent to a combination of any of the other steps. If so, the forward and reverse rate expressions must yield equilibrium constants that are in agreement with the stoichiometric relationships. Next, the complete set of steps is examined to look for "illegal reaction loops". Both of these procedures are performed by constructing matrices that represent the stoichiometries of the various reaction steps and then performing row reductions to identify basis sets of loops. A method based on linear programming is described that determines whether a mechanism contains any illegal loops. These methods are applied in the analysis of several published reaction mechanisms.

12.
Inorg Chem ; 57(3): 1665-1669, 2018 Feb 05.
Artigo em Inglês | MEDLINE | ID: mdl-29364656

RESUMO

Herein is a dissection of the energetic contributions to a correlation between the rates and driving forces for oxygen atom transfer from three inorganic peroxides to the halides. Experimental data are used to calculate conventional and absolute hydration enthalpies for OCl-, OBr-, and OI-. It is found that the hydration enthalpies are more exothermic for the OX- species in comparison to their X- congeners, and it is found that the hydration enthalpies are approximately constant on progressing from OCl- to OI-. Both of these trends are contrary to expectations based on simple models of ionic hydration. Similar trends are seen in the Gibbs energies of hydration. The strong decrease in E° from OCl- to OI- is seen to arise primarily from these differing trends in hydration energies rather than the gas-phase oxygen atom affinities of the halides. These effects show that the Marcus-like driving-force dependence for oxygen atom transfer from peroxides to the halides arises from the differing trends in hydration energies rather than in the intrinsic O-X- bond strengths.

13.
Inorg Chem ; 56(7): 4047-4056, 2017 Apr 03.
Artigo em Inglês | MEDLINE | ID: mdl-28290673

RESUMO

Sulfur-containing compounds are known to be susceptible to oxidation by aqueous HOCl, but the factors affecting the rates of these reactions are not well-established. Here we report on the kinetics of oxidation of thiosulfate, thiourea, thioglycolate, (methylthio)acetate, tetrathionate, dithiodiglycolate, and dithiodipropionate at 25 °C and 0.4 M ionic strength. These reactions obey the general rate law -d[OCl-]/dt = (kOCl-[OCl-] + kHOCl[HOCl])[substrate] with some exceptions: tetrathionate and the two disulfides undergo rate-limiting hydrolysis at high pH, and dithiodiglycolate has an additional term in the rate law that is second order in [substrate]. The reactions of HOCl are believed to have a Cl+ transfer mechanism, and in the case of thiosulfate the rate of hydrolysis of the ClS2O3- intermediate was determined. In the case of thiourea evidence was obtained for thiourea monoxide as a long-lived product. It is shown that sulfite and species with terminal sulfur atoms have kHOCl values in the vicinity of 1 × 109 M-1 s-1, while SCN- and thioethers react somewhat more slowly; tetrathionate, trithionate, and disulfides react much more slowly. Comparison of the rate constants with those for oxidation of these sulfur substrates by H2O2 and [Pt(CN)4Cl2]2- shows that HOCl reacts a few orders of magnitude more rapidly than [Pt(CN)4Cl2]2- and ∼9 orders of magnitude more rapidly than H2O2. Many of the kHOCl values are leveled by the high electrophilicity of HOCl. It is proposed that the kOCl- values correspond to oxygen-atom transfer mechanisms, as supported by LFERS (linear free energy relationships) relating these rate constants to those for reactions of H2O2 and [Pt(CN)4Cl2]2-.

14.
Inorg Chem ; 55(15): 7797-803, 2016 Aug 01.
Artigo em Inglês | MEDLINE | ID: mdl-27410173

RESUMO

Detailed reports on the oxidation of aqueous H2S by mild one-electron oxidants are lacking, presumably because of the susceptibility of these reactions to trace metal-ion catalysis and the formation of turbid sulfur sols. Here we report on the reaction of [IrCl6](2-) with H2S in acetate buffers. Dipicolinic acid (dipic) is shown to be effective in suppressing metal-ion catalysis. In the presence of dipic the reaction produces [IrCl6](3-) and polysulfides; turbidity develops primarily after the Ir(IV) oxidant is consumed. Water-soluble phosphines are shown to prevent the development of turbidity; in the case of tris-hydroxymethylphosphine (THMP) the product is the corresponding sulfide, THMP═S. THMP diminishes the rates of reduction of Ir(IV), and the rate law with sufficient THMP is first order in [Ir(IV)] and first order in [HS(-)]. The rate-limiting step is inferred to be electron transfer from HS(-) to Ir(IV) with ket = 2.9 × 10(4) M(-1) s(-1) at 25.0 °C and µ = 0.1 M. The kinetic inhibition by THMP is attributed to its interception of a polysulfide chain elongation process.

15.
Inorg Chem ; 55(1): 366-70, 2016 Jan 04.
Artigo em Inglês | MEDLINE | ID: mdl-26678913

RESUMO

There has been steady interest in the aqueous reaction of ClO2• with sulfur(IV) since the 1950s, and a wide variety of rate laws and mechanisms have been proposed. In neutral-to-alkaline media, the reaction is challenging to study because of its great rate. Here it is shown that benzaldehyde can be used as an additive to slow the reaction and make its rates more amenable to study. The rates can be quantitatively modeled by a mechanism that includes reversible binding of sulfur(IV) by benzaldehyde and a rate-limiting mixed second-order reaction of ClO2• with SO3(2-). The latter reaction occurs through parallel electron transfer from SO3(2-) to ClO2• and oxygen-atom transfer from ClO2• to SO3(2-).

16.
J Phys Chem A ; 118(34): 6827-31, 2014 Aug 28.
Artigo em Inglês | MEDLINE | ID: mdl-25099361

RESUMO

The reaction of ClO2(•) with S2O3(2-) in aqueous solution is a component of the "crazy clock" reaction of ClO2(-) with S2O3(2-), and under conditions of excess S2O3(2-) the absorbance at 360 nm due to ClO2(•) decays with sigmoidal kinetics. A chain reaction mechanism is inferred on the basis that very small concentrations of SO3(2-) accelerate the reaction, and methionine inhibits the reaction. Pseudo-first-order kinetics is observed in the presence of relatively large methionine concentrations, leading to the simple rate law -d[ClO2]/dt = (ka[S2O3(2-)] + kb[S2O3(2-)](2))[ClO2], with ka = 452 ± 16 M(-1) s(-1) and kb = (5.7 ± 0.2) × 10(5) M(-2) s(-1) at 25 °C and pH 7.6. Under these conditions, the initial products are ClO2(-) and S4O6(2-). A classical electron-transfer mechanism is assigned to the reaction that occurs under conditions of methionine inhibition.

17.
J Phys Chem B ; 118(4): 1097-101, 2014 Jan 30.
Artigo em Inglês | MEDLINE | ID: mdl-24400869

RESUMO

We report the results of an experimental study of the oxidation of cysteinesulfinic acid (CysSO2H) by [IrCl6](2-) in aqueous media at 25 °C in order to gain insight into the mechanisms of oxidation of alkylsulfinic acids by simple one-electron oxidants. When the reaction is performed with exclusion of O2 between pH 3 and 5, it is complete in several seconds. The products are [IrCl6](3-) and CysSO3H. Kinetic data obtained by stopped-flow UV-vis methods with [CysSO2H] ≫ [Ir(IV)]0 reveal the rate law to be -d[Ir(IV)]/dt = k[Ir(IV)](2)[CysSO2H]/[Ir(III)] with a negligible pH dependence. The value of k is (6.8 ± 0.12) × 10(3) M(-1) s(-1) at µ = 0.1 M (NaClO4). A mechanism is inferred in which the first step is a rapid and reversible electron-transfer equilibrium between Ir(IV) and CysSO2(-) to form Ir(III) and CysSO2(•). The second step is the rate-limiting inner-sphere oxidation of CysSO2(•) by Ir(IV). Production of CysSO3H is proposed to occur through hydrolysis of an Ir(III)-bound sulfonyl chloride that is the immediate product of the inner-sphere second step.


Assuntos
Cisteína/análogos & derivados , Irídio/química , Ácidos Sulfênicos/química , Cisteína/química , Conformação Molecular , Oxirredução
18.
Inorg Chem ; 51(24): 13303-11, 2012 Dec 17.
Artigo em Inglês | MEDLINE | ID: mdl-23186256

RESUMO

The aqueous oxidations of glutathione (GSH) by [IrCl(6)](2-), [Fe(bpy)(2)(CN)(2)](+), and [Fe(bpy)(CN)(4)](-) are described. All three reactions are highly susceptible to catalysis by traces of copper ions, but this catalysis can be fully suppressed with suitable chelating agents. The direct oxidation by [IrCl(6)](2-) yields [IrCl(6)](3-) and GSO(3)(-); some GSSG is also obtained in the presence of O(2). The two Fe(III) oxidants are reduced to their corresponding Fe(II) complexes with nearly quantitative formation of GSSG. The kinetics of these reactions have been studied at 25 °C and µ = 0.1 M between pH 1 and 11. All three reactions have rate laws that are first order in [M(ox)] and [GSH](t) and show a general increase in rate with increasing pH. Detailed studies of the pH dependence enable the rate law to be elaborated with terms for reaction of the individual protonation states of GSH. These pH-resolved rate constants are interpreted with a mechanism having rate-limiting outer-sphere electron-transfer from the various thiolate forms of GSH.


Assuntos
2,2'-Dipiridil/química , Complexos de Coordenação/química , Glutationa/química , Irídio/química , Ferro/química , Nitrilas/química , Piridinas/química , Concentração de Íons de Hidrogênio , Oxirredução , Água/química
19.
Inorg Chem ; 51(9): 4909-11, 2012 May 07.
Artigo em Inglês | MEDLINE | ID: mdl-22515498

RESUMO

Outer-sphere oxidation of phenols is under intense scrutiny because of questions related to the dynamics of proton-coupled electron transfer (PCET). Oxidation by cationic transition-metal complexes in aqueous solution presents special challenges because of the potential participation of the solvent as a proton acceptor and of the buffers as general base catalysts. Here we report that oxidation of phenol by a deficiency of [Os(phen)(3)](3+), as determined by stopped-flow spectrophotometry, yields a unique rate law that is second order in [osmium(III)] and [phenol] and inverse second order in [osmium(II)] and [H(+)]. A mechanism is inferred in which the phenoxyl radical is produced through a rapid PCET preequilibrium, followed by rate-limiting phenoxyl radical coupling. Marcus theory predicts that the rate of electron transfer from phenoxide to osmium(III) is fast enough to account for the rapid PCET preequilibrium, but it does not rule out the intervention of other pathways such as concerted proton-electron transfer or general base catalysis.

20.
Inorg Chem ; 50(24): 12762-73, 2011 Dec 19.
Artigo em Inglês | MEDLINE | ID: mdl-22087767

RESUMO

It has been previously established that the aqueous oxidation of phenol by a deficiency of [IrCl(6)](2-) proceeds through the production of [IrCl(6)](3-) and phenoxyl radicals. Coupling of the phenoxyl radicals leads primarily to 4,4'-biphenol, 2,2'-biphenol, 2,4'-biphenol, and 4-phenoxyphenol. Overoxidation occurs through the further oxidation of these coupling products, leading to a rather complex mixture of final products. The rate laws for oxidation of the four coupling products by [IrCl(6)](2-) have the same form as those for the oxidation of phenol itself: -d[Ir(IV)]/dt = {(k(ArOH) + k(ArO(-))K(a)/[H(+)])/(1 + K(a)/[H(+)])}[ArOH](tot)[Ir(IV)]. Values for k(ArOH) and k(ArO(-)) have been determined for the four substrates at 25 °C and are assigned to H(2)O-PCET and electron-transfer mechanisms, respectively. Kinetic simulations of a combined mechanism that includes the rate of oxidation of phenol as well as the rates of these overoxidation steps show that the degree of overoxidation is rather limited at high pH but quite extensive at low pH. This pH-dependent overoxidation leads to a pH-dependent stoichiometric factor in the rate law for oxidation of phenol and causes some minor deviations in the rate law for oxidation of phenol. Empirically, these minor deviations can be accommodated by the introduction of a third term in the rate law that includes a "pH-dependent rate constant", but this approach masks the mechanistic origins of the effect.

SELEÇÃO DE REFERÊNCIAS
DETALHE DA PESQUISA
...