Your browser doesn't support javascript.
loading
Show: 20 | 50 | 100
Results 1 - 20 de 45
Filter
Add more filters











Publication year range
1.
J Phys Chem B ; 126(15): 2876-2890, 2022 04 21.
Article in English | MEDLINE | ID: mdl-35389658

ABSTRACT

What happens at the ionic-liquid (IL)/water interface when the Eu3+ cation is complexed and extracted by bis(dimethyltriazinyl) pyridine "BTP" ligands has been investigated by molecular dynamics and potential of mean force simulations on the interface crossing by key species: neutral BTP, its protonated BTPH+ form, Eu3+, and the Eu(BTP)33+ complex. At both the [BMI][Tf2N]/water and [OMI][Tf2N]/water interfaces, neither BTP nor Eu(BTP)33+ are found to adsorb. The distribution of Eu(BTP)23+ and Eu(BTP)3+ precursors of Eu(BTP)33+, and of their nitrate adducts, implies the occurrence of a stepwise complexation process in the interfacial domain, however. The analysis of the ionic content of the bulk phases and of their interface before and after extraction highlights the role of charge buffering by interfacial IL cations and anions, by different amounts depending on the IL. Comparison of ILs with octanol as the oil phase reveals striking differences regarding the extraction efficiency, the affinity of Eu(BTP)33+ for the interface, the effects of added nitric acid and of counterions (NO3- vs Tf2N-), charge neutralization mechanisms, and the extent of "oil" heterogeneity. Extraction into octanol is suggested to proceed via adsorption at the surface of water pools, nanoemulsions, or droplets, with marked counterion effects.


Subject(s)
Ionic Liquids , Cations , Ligands , Liquid-Liquid Extraction , Octanols , Water
2.
J Phys Chem B ; 122(44): 10143-10157, 2018 Nov 08.
Article in English | MEDLINE | ID: mdl-30351128

ABSTRACT

The formation of aqueous biphasic systems (ABSs) based on aqueous ionic liquid (IL)/salt mixtures has been investigated via molecular dynamics simulations (with IL butyl-methyl-imidazolium triflate; salts NaCl, CsCl, SrCl2, and EuCl3). The analysis of ion distributions, solvation, and mutual interactions during the dynamics reveals the heterogeneity of all solutions due to ion segregation into mutually exclusive IL and salt domains, even in monophasic solutions ("ionic sociology"). Ion segregation and ABS formation are found to increase with (i) the salt content and (ii) the IL content, (iii) in the order Na+ < Sr2+ < Eu3+, and (iv) when the IL ion "polarity" is diminished, following experimental trends. The structuration of the solution is rationalized as a synergistic water transfer from the best donating ion pair (first hydration shell of hydrophobic moieties of IL ions) to the best accepting pair (M n+ and Cl- ions, beyond their first shell). In ABSs, the IL- and salt-containing phases are linked by a well-defined "interface" that decreases in width when MCl n becomes more hydrophilic and/or more concentrated. In the IL-rich phase of ABSs, the hydration of IL ions and their mutual interactions are shown to be similar to those displayed at aqueous interfaces.

3.
Inorg Chem ; 56(7): 4148-4156, 2017 Apr 03.
Article in English | MEDLINE | ID: mdl-28318233

ABSTRACT

The mechanism by which Zr-substituted and other transition metal-substituted polyoxometalates (POMs) form covalently linked dimers has been analyzed by means of static density functional theory (DFT) calculations with a continuous solvent model as well as Car-Parrinello molecular dynamics (CPMD) simulations with explicit solvent molecules. The study includes different stages of the process: the formation of the active species by alkalination of the solution and formation of intercluster linkages. CPMD simulations show that the Zr-triaqua precursor, [W5O18Zr(H2O)3]2-, under basic conditions, reacts with hydroxide anions to form Zr-aqua-hydroxo active species, [W5O18Zr(OH)(H2O)]3-. We computed the DFT potential energy profile for dimerization of [W5O18TM(OH)]n- [TM = ZrIV(H2O), ZrIV, TiIV, and WVI] anions. The resulting overall energy barrier is low for ZrIV, moderate for TiIV, and high for WVI. The computed thermodynamic balance favors the dibridged (µOH)2 linkages for ZrIV, the monobridged (µOH) linkage for TiIV, and the monomeric forms for WVI, in agreement with experimentally observed trends. The lowest energy barrier and largest coordination number of Zr-substituted POMs are both a consequence of the flexible coordination environment and larger radius of Zr.

4.
Chemistry ; 22(47): 17037-17046, 2016 Nov 14.
Article in English | MEDLINE | ID: mdl-27727476

ABSTRACT

Purely organic radical ions dimerize in solution at low temperature, forming long, multicenter bonds, despite the metastability of the isolated dimers. Here, we present the first computational study of these π-dimers in solution, with explicit consideration of solvent molecules and finite temperature effects. By means of force-field and ab initio molecular dynamics and free energy simulations, the structure and stability of π-[TCNE]22- (TCNE=tetracyanoethylene) dimers in dichloromethane have been evaluated. Although the dimers dissociate at room temperature, they are stable at 175 K and their structure is similar to the one in the solid state, with a cofacial arrangement of the radicals at an interplanar separation of approximately 3.0 Å. The π-[TCNE]22- dimers form dissociated ion pairs with the NBu4+ counterions, and their first solvation shell comprises approximately 20 CH2 Cl2 molecules. Among them, the eight molecules distributed along the equatorial plane of the dimer play a key role in stabilizing the dimer through bridging C-H⋅⋅⋅N contacts. The calculated free energy of dimerization of TCNE.- in solution at 175 K is -5.5 kcal mol-1 . These results provide the first quantitative model describing the pairing of radical ions in solution, and demonstrate the key role of solvation forces on the dimerization process.

5.
Dalton Trans ; 44(6): 2623-38, 2015 Feb 14.
Article in English | MEDLINE | ID: mdl-25412447

ABSTRACT

We report DFT static and dynamic studies on uranyl complexes [UO(2)(NO(3))x(H(2)O)(y)L(z)](2-x) involved in the uranyl extraction from water to an "oil" phase (hexane) by an amide ligand L (N,N-dimethylacetamide). Static DFT results "in solution" (continuum SMD models for water and hexane) predict that the stepwise formation of [UO(2)(NO(3))(2)L(2)] from the UO(2)(H(2)O)(5)(2+) species is energetically favourable, and allow us to compare cis/trans isomers of penta- and hexa-coordinated complexes and key intermediates in the two solvents. DFT-MD simulations of [UO(2)(NO(3))(2)L(2)], [UO(2)(NO(3))(2)(H(2)O)L(2)], and [UO(2)(NO(3))(H(2)O)L(2)](+) species in explicit solvent environments (water, hexane, or the water/hexane interface) represented at the MM or full-DFT level reveal a versatile solvent dependent binding mode of nitrates, also evidenced by metadynamics simulations. In water and at the interface, the latter exchange from bi- to monodentate, via in plane rotational motions in some cases. Remarkably, structures of complexes at the interface are more "water-like" than gas phase- or hexane-like. Thus, the order of U-O(NO(3))/U-O(L) bond distances observed in the gas phase (U-O(nit) < U-OL) is inverted at the interface and in water. Overall, the results are consistent with the experimental observation of uranyl extraction from nitric acid solutions by amide analogues (bearing "fatty" substituents), and allow us to propose possible extraction mechanisms, involving complexation of L "right at the interface". They also point to the importance of the solvent environment and the dynamics on the structure and stability of the complexes.


Subject(s)
Acetamides/chemistry , Molecular Dynamics Simulation , Organometallic Compounds/chemistry , Quantum Theory , Uranium/chemistry , Ligands , Molecular Conformation , Organometallic Compounds/chemical synthesis
6.
J Phys Chem A ; 118(38): 8842-51, 2014 Sep 25.
Article in English | MEDLINE | ID: mdl-25133873

ABSTRACT

Quantum chemical calculations are used to derive distributed polarizability models sufficiently accurate and compact to be used in classical molecular dynamics simulations of imidazolium-based room temperature ionic liquids. Two distributed polarizability models are fitted to reproduce the induction energy of three imidazolium cations (1,3-dimethyl-, 1-ethyl-3-methyl-, and 1-butyl-3-methylimidazolium) and four anions (tetrafluoroborate, hexafluorophosphate, nitrate, and thiocyanate) polarized by a point charge located successively on a grid of surrounding points. The first model includes charge-flow polarizabilities between first-neighbor atoms and isotropic dipolar polarizability on all atoms (except H), while the second model includes anisotropic dipolar polarizabilities on all atoms (except H). For the imidazolium cations, particular attention is given to the transferability of the distributed polarizability sets. The molecular polarizability and its anisotropy rebuilt by the distributed models are found to be in good agreement with the exact ab initio values for the three cations and 23 additional conformers of 1-ethyl-3-methyl-, 1-butyl-3-methyl-, 1-pentyl-3-methyl-, and 1-hexyl-3-methylimidazolium cations.

7.
Inorg Chem ; 53(2): 778-86, 2014 Jan 21.
Article in English | MEDLINE | ID: mdl-24364825

ABSTRACT

Static density functional theory (DFT) calculations with a continuous solvent model as well as classical and Car-Parrinello molecular dynamics (MD) simulations with explicit solvent molecules were performed to study the nature of Zr-monosubstituted monomeric and dimeric polyoxometalates (POMs) in water at different pHs. We have analyzed Zr-aqua, -hydroxo, and -aqua-hydroxo species derived from Linqvist- and Keggin-type anions. Both DFT and Car-Parrinello MD methods suggest that the Zr center tends to have coordination number greater than 6 and can bind up to 3 water molecules. Car-Parrinello MD simulations also show that the Zr atom fluctuates within the oxide POM framework, providing a flexible coordination environment. There is a small thermodynamic preference for the Zr-aqua species over the protonated Zr-hydroxo species; however the prevalence of one or the other species might depend on the pH. Classical MD simulations show that H3O(+) interacts mainly with hydroxo ligand, while OH(-) anions prefer the protons of the H2O ligands. In general, an increase of the acidity favors the formation of Zr-aqua species, explaining why dimer dissociation is promoted at low pH. At basic conditions Zr-hydroxo species are generated, providing the reactive groups to form Zr···Zr linkages.

8.
Dalton Trans ; 43(29): 11129-37, 2014 Aug 07.
Article in English | MEDLINE | ID: mdl-24310370

ABSTRACT

Static ab initio and density-functional computations, as well as Car-Parrinello molecular dynamics simulations in aqueous solution are reported for [UO2(OH)(κ(2)-O2)(H2O)2](-) and [UO2(OH)2(κ(1)-O2H)(H2O)](-). Whereas the κ(1)-hydroperoxo isomer is found to be more stable than the κ(2)-peroxo form in the gas phase, the order of stability is reversed in explicit bulk solution. Based on free energies from thermodynamic integration (BLYP functional), the peroxo form is favoured by ca. 32 kJ mol(-1) in water. This stabilisation is discussed in terms of the hydration shells about the individual ligands and dipole moments of the complexes in water, and highlights the importance of explicit solute-solvent interactions and bulk solvation for the speciation of uranyl(vi) compounds.

9.
Inorg Chem ; 51(24): 13396-407, 2012 Dec 17.
Article in English | MEDLINE | ID: mdl-23181554

ABSTRACT

Car-Parrinello molecular dynamics (CMPD) simulations and static computations are reported at the BLYP level of density functional theory (DFT) for mixed [LaCl(x)(H(2)O)(y)(MeCN)(z)](3-x) complexes in aqueous and nonaqueous solution (acetonitrile). Both methodologies predict coordination numbers (i.e., x + y + z) that are successively lower than nine as the Cl content increases from x = 0 to 3. While the static DFT method with implicit solvation through a polarizable continuum model overestimates the binding strength of chloride and erroneously predicts [LaCl(2)(H(2)O)(5)](+) as global free-energy minimum, constrained CPMD simulations with explicit solvent and thermodynamic integration reproduce the weak binding of chloride in water reasonably well. Special attention is called to the dipole moments of coordinated water molecules as function of coligands and solvent, evaluated through maximally localized Wannier function centers along the CPMD trajectories. Cooperative polarization of these water ligands by the metal cation and the surrounding solvent is remarkably sensitive to fluctuations of the La-O distances and, to a lesser extent, on the La-water tilt angles. The mean dipole moment of water ligands is rather insensitive to the other coligands, oscillating around 3.2 D, 3.5 D, and 3.3 D in MeCN, water, and [dmim]Cl solution, respectively, the latter being an archetypical ionic liquid.

10.
Chemphyschem ; 13(7): 1677-86, 2012 May 14.
Article in English | MEDLINE | ID: mdl-22556119

ABSTRACT

We report a molecular dynamics study on the EuBr(n)(3-n) complexes (n=0 to 6) formed upon complexation of Br(-) by Eu(3+) in the [BMI][PF(6)], [BMI][Tf(2)N] and [MeBu(3)N][Tf(2)N] ionic liquids (ILs), to compare the effect of the IL anion (PF(6)(-) versus Tf(2)N(-)), the IL cation (BMI(+) versus MeBu(3)N(+)) and the "IL humidity" on their solvation and stability. In "dry" solutions all complexes remain stable and the first coordination shell of Eu(3+) is purely anionic (Br(-) and IL anions), surrounded by IL cations (BMI(+) or MeBu(3)N(+) ions). Long range "onion type" solvation features (up to 20 Å from Eu(3+)), with alternating cation-rich and anion-rich solvent shells, are observed around the different complexes. The comparison of gas phase-optimized structures of EuBr(n)(3-n) complexes (that are unstable for n=5 and 6) with those observed in solution points to the importance of solvation forces on the nature of the complex, with a higher stabilization by imidazolium- than by ammonium-based dry ILs. Adding water to the IL has different effects, depending on the IL. In the highly hygroscopic [BMI][PF(6)] IL, Br(-) ligands are displaced by water, to finally form Eu(H(2)O)(9)(3+). In the less "humid" [BMI][Tf(2)N], the EuBr(n)(3-n) complexes do not dissociate and coordinate at most 1-2 H(2)O molecules. We also calculated the free-energy profiles (Potential of Mean Force calculations) for the stepwise complexation of Br(-), and found significant solvent effects. EuBr(6)(3-) is predicted to form in both [BMI][PF(6)] and [BMI][Tf(2)N], but not in [MeBu(3)N][Tf(2)N], mainly due to weaker interactions with the cationic solvation shell. First steps are found to be more exergonic in the PF(6)(-)- than in the Tf(2)N(-)-based IL. Molecular dynamics (MD) comparisons between ILs and classical solvents (acetonitrile and water) are also reported, affording good agreement with the experimental observations of Br(-) complexation by trivalent lanthanides in these classical solvents.

11.
J Phys Chem B ; 116(10): 3205-19, 2012 Mar 15.
Article in English | MEDLINE | ID: mdl-22329618

ABSTRACT

The complexation of perrhenate (ReO(4)(-)) anions by the uranyl (UO(2)(2+)) cation has been investigated by joint molecular dynamics simulations and spectroscopic (UV-vis, TRLFS, and EXAFS) studies in aqueous solution, acetonitrile, and three ionic liquids (ILs), namely, [Bmi][Tf(2)N], [Me(3)BuN][Tf(2)N], and [Bu(3)MeN][Tf(2)N] that are based on the same Tf(2)N(-) anion (bis(trifluoromethylsulfonyl)imide) and either Bmi(+) (1-butyl,3-methylimidazolium), Me(3)BuN(+), or Bu(3)MeN(+) cations. They show that ReO(4)(-) behaves as a weak ligand in aqueous solution and as a strong ligand in acetonitrile and in the ILs. According to MD simulations in aqueous solution, the UO(2)(ReO(4))(2) complex quickly dissociates to form UO(2)(H(2)O)(5)(2+), while in acetonitrile, a stable UO(2)(ReO(4))(5)(3-) species forms from dissociated ions. In the ILs, the UO(2)(ReO(4))(n)(2-n) complexes (n = 1 to 5) remained stable along the dynamics, and to assess their relative stabilities, we computed the free energy profiles for stepwise ReO(4)(-) complexation to uranyl. In the two studied ILs, complexation is favored, leading to the UO(2)(ReO(4))(5)(3-) species in [Bmi][Tf(2)N] and to UO(2)(ReO(4))(4)(2-) in [Bu(3)MeN][Tf(2)N]. Furthermore, in both acetonitrile and [Bmi][Tf(2)N] solutions, MD and PMF simulations support the formation of dimeric uranyl complexes [UO(2)(ReO(4))(4)](2)(4-) with two bridging ReO(4)(-) ligands. The simulation results are qualitatively consistent with spectroscopic observations in the different solvents, without firmly concluding, however, on the precise composition and structure of the complexes in the solutions.

12.
Inorg Chem ; 51(3): 1943-52, 2012 Feb 06.
Article in English | MEDLINE | ID: mdl-22233256

ABSTRACT

Optimizations at the BLYP and B3LYP levels are reported for the mixed uranyl chloro/water/acetonitrile complexes [UO(2)Cl(n)(H(2)O)(x)(MeCN)(5-n-x)](2-n) (n = 1-3) and [UO(2)Cl(n)(H(2)O)(x)(MeCN)(4-n-x)](2-n) (n = 2-4), in both the gas phase and a polarizable continuum modeling acetonitrile. Car-Parrinello molecular dynamics (CPMD) simulations have been performed for [UO(2)Cl(2)(H(2)O)(MeCN)(2)] in the gas phase and in a periodic box of liquid acetonitrile. According to population analyses and dipole moments evaluated from maximally localized Wannier function centers, uranium is less Lewis acidic in the neutral UO(2)Cl(2) than in the UO(2)(2+) moiety. In the gas phase the latter binds acetonitrile ligands more strongly than water, whereas in acetonitrile solution, the trend is reversed due to cooperative polarization effects. In the polarizable continuum the chloro complexes have a slight energetic preference for water over acetonitrile ligands, but several mixed complexes are so close in free energy ΔG that they should exist in equilibrium, in accord with previous interpretations of EXAFS data in solution. The binding strengths of the fifth neutral ligands decrease with increasing chloride content, to the extent that the trichlorides should be formulated as four-coordinate [UO(2)Cl(3)L](-) (L = H(2)O, MeCN). Limitations to their accuracy notwithstanding, density functional calculations can offer insights into the speciation of a complex uranyl system in solution, a key feature in the context of nuclear waste partitioning by complexant molecules.

13.
Chemphyschem ; 12(17): 3095-105, 2011 Dec 09.
Article in English | MEDLINE | ID: mdl-21960471

ABSTRACT

First-principles and purely classical molecular dynamics (MD) simulations for complexes of the uranyl ion (UO(2)(2+)) are reviewed. Validation of Car-Parrinello MD simulations for small uranyl complexes in aqueous solution is discussed. Special attention is called to the mechanism of ligand-exchange reactions at the uranyl centre and to effects of solvation and hydration on coordination and structural properties. Large-scale classical MD simulations are surveyed in the context of liquid-liquid extraction, with uranyl complexes ranging from simple hydrates to calixarenes, and nonaqueous phases from simple organic solvents and supercritical CO(2) to ionic liquids.

14.
J Phys Chem B ; 115(10): 2338-48, 2011 Mar 17.
Article in English | MEDLINE | ID: mdl-21338145

ABSTRACT

We report a molecular dynamics study on pertechnetic acid (PTA) extraction from water to an oil phase containing either pure TBP (tri-n-butyl phosphate) or a TBP/n-hexane mixture, with the main aim to understand the state of the acid (associated TcO(4)H vs dissociated TcO(4)(-)H(+)) and its "complexation" by TBP. Experimentally, Tc(VII) is extracted from acidic water to TBP:alkane solutions in 1:3 or 1:4 Tc:TBP ratio, suggesting that three or four TBPs coordinate to TcO(4)H or TcO(4)(-). According to simulations in TBP solution, however, neither TcO(4)H nor TcO(4)(-) species displays specific coordination to TBP. We thus investigated several hypothetical states of the proton of the dissociated pertechnetic acid in organic phases and at their aqueous interfaces, comparing "pH neutral" to nitric acid containing systems. Proton hydrates are shown to coordinate 3-4 TBPs, in the form of H(3)O(+)(TBP)(3) and H(5)O(2)(+)(TBP)(4) hydrogen-bonded adducts, whereas TBPH(+) binds 1 TBP. The MD and PMF results complemented by those of QM investigations suggest that Tc(VII) is extracted as TcO(4)(-)(H(3)O)(+)(TBP)(3) or TcO(4)(-)(H(5)O(2))(+)(TBP)(4) contact ion pairs instead of the neutral form TcO(4)H of the acid. They explain why nitric acid promotes the Tc(VII) extraction. Comparison between nitric acid (mainly extracted via its neutral form NO(3)H) and pertechnetic acid is discussed.

15.
Inorg Chem ; 50(1): 299-308, 2011 Jan 03.
Article in English | MEDLINE | ID: mdl-21126026

ABSTRACT

Optimizations at the BLYP and B3LYP levels are reported for mixed uranyl-water/acetonitrile complexes [UO(2)(H(2)O)(5-n)(MeCN)(n)](2+) (n = 0-5), in both the gas phase and a polarizable continuum modeling acetonitrile. Car-Parrinello molecular dynamics (CPMD) simulations have been performed for these complexes in the gas phase, and for selected species (n = 0, 1, 3, 5) in a periodic box of liquid acetonitrile. According to structural and energetic data, uranyl has a higher affinity for acetonitrile than for water in the gas phase, in keeping with the higher dipole moment and polarizability of acetonitrile. In acetonitrile solution, however, water is the better ligand because of specific solvation effects. Analysis of the dipole moment of the coordinated water molecule in [UO(2)(H(2)O)(MeCN)(4)](2+) reveals that the interaction with the second-shell solvent molecules (through fairly strong and persistent O-H···N hydrogen bonds) causes a significant increase of this dipole moment (by more than 1 D). This cooperative polarization of water reinforces the uranyl-water bond as well as the water solvation via strengthened (UO(2))OH(2)···NCMe hydrogen bonds. Such cooperativity is essentially absent in the acetonitrile ligands that make much weaker (UO(2))NCMe···NCMe hydrogen bonds. Beyond the uranyl case, this study points to the importance of cooperative polarization effects to enhance the M(n+) ion affinity for water in condensed phases involving M(n+)-OH(2)···A fragments, where A is a H-bond proton acceptor and M(n+) is a hard cation.

16.
Phys Chem Chem Phys ; 11(44): 10299-310, 2009 Nov 28.
Article in English | MEDLINE | ID: mdl-19890513

ABSTRACT

Norbadione A (NBA) is a pigment present in edible mushrooms which is presumed to selectively complex Cs(+) cations. Due to a very uncommon complexation mechanism, we used a combination of several experimental techniques, including (1)H-NMR, (133)Cs-NMR, isothermal calorimetric, potentiometric titrations and molecular dynamics MD simulations to determine the nature of the complexed species, as well as their stability constants for the NBA-M(+) systems (M(+) = Cs(+), K(+), Na(+)) in methanol:water 80:20 solutions at 25.0 degrees C. We show that almost no complexation occurs below pH 7.5, as long as a proton, involved in a strong hydrogen bond, bridges both carboxylic and enolic groups of each pulvinic moiety of NBA. Thus, neutralization of that proton is necessary to both set free potential coordination sites and to trigger a conformational change, two conditions needed to bind successively a first, then a second metallic cation. The stability constants determined in this study are in good agreement with each other, leading to the stability order Cs(+) > K(+) > Na(+) for both mono- and bimetallic complexes, which is the reversed order to the one generally observed for low molecular weight carboxylic ligands in water. According to MD simulations in solution, complexation involves a mixture of Z/E isomers and conformers of NBA with a broad diversity of binding modes. Some pH and environment dependent aggregation phenomena are considered to also contribute to the binding process, and to possibly explain the accumulation of radionuclides in mushrooms.


Subject(s)
4-Butyrolactone/analogs & derivatives , Cesium/chemistry , Hydrogen Bonding , Phenylacetates/chemistry , Potassium/chemistry , Sodium/chemistry , 4-Butyrolactone/chemistry , Calorimetry , Hydrogen-Ion Concentration , Magnetic Resonance Spectroscopy , Spectrophotometry, Ultraviolet
17.
Inorg Chem ; 48(21): 9977-9, 2009 Nov 02.
Article in English | MEDLINE | ID: mdl-19780567

ABSTRACT

The inclusion of NH(4)(+) as counterions in Car-Parrinello molecular dynamics (CPMD) simulations of anionic uranyl(VI) complexes is proposed as a viable approach to modeling "real" aqueous solutions. For [UO(2)F(4)(H(2)O)](2-) in water, it is shown that the inclusion of two NH(4)(+) ions strengthens the bond between uranyl and the water ligand by ca. 2 kcal/mol, improving the accordance with experiment. According to CPMD simulations for [UO(2)X(5)][NH(4)](3) (X = F, OH) in water, the fifth fluoride is bound much stronger than the fifth OH(-). Implications for a recently proposed model for oxygen exchange in uranyl hydroxide are discussed.

18.
Phys Chem Chem Phys ; 10(46): 6940-53, 2008 Dec 14.
Article in English | MEDLINE | ID: mdl-19030589

ABSTRACT

Aqueous solutions of the polyoxometallate alpha-PW(12)O(40)(3-) Keggin anion "PW(3-)" have been simulated by molecular dynamics, comparing two anion concentrations (0.06 and ca. 0.15 mol l(-1)) and Cs(+), NBu(4)(+), UO(2)(2+), Eu(3+), H(3)O(+) and H(5)O(2)(+) as neutralizing M(n+) counterions. They reveal marked counterion effects of the degree of salt dilution, cation-anion and anion-anion relationships. The hydrophobic NBu(4)(+) cations tend to surround PW(3-)'s via loose contacts, leading to "phase separation" between water and a humid salty domain, overall neutral, where all ions are concentrated. The other studied cations are more hydrophilic and generally separated from the PW(3-) anions. The most important finding concerns the aggregation of PW's, mostly as dimers with short contacts (PP < 12 A), but also as higher (PW(3-))(n) oligomers (n = 3 to 5) in concentrated solutions where the proportion of the aggregates ranges from ca. 9 to 46%, depending on the counterion. While Eu(3+) and UO(2)(2+) are fully hydrated and interact at short distances with PWs as solvent-separated ion pairs, Cs(+) can form contact ion pairs, as well as solvent-separated ions. Among the mono-charged counterions, H(5)O(2)(+) gives highest aggregation (ca. 47%, involving 32% of dimers, 11% of trimers and 3% of tetramers), pointing to the influence of the proton state (H(5)O(2)(+)vs. H(3)O(+)) on PW's aggregation and condensation. The dynamic properties are also dependent on M(n+): the PW's diffusion coefficients are lowest with NBu(4)(+), and highest for Cs(+), thus reflecting the degree of ion condensation in water. The role of water on the solution state of the PW salts is further demonstrated by simulating the most concentrated systems in methanol solution. Because MeOH solvates less well the M(n+) cation than does H(2)O and cannot afford bridging relays between PW's, one finds a higher proportion of PW(3-)M(n+) contacts, and no (PW(3-))(n) oligomers with short contacts in methanol.

19.
J Phys Chem B ; 112(38): 12014-23, 2008 Sep 25.
Article in English | MEDLINE | ID: mdl-18754633

ABSTRACT

The stepwise addition of 1 to 4 Cl(-) anions to the uranyl cation has been studied via potential of mean force (PMF) calculations in the [BMI][Tf 2N] ionic liquid based on the 1-butyl-3-methylimidazolium cation (BMI(+)) and the bis(trifluoromethylsulfonyl)imide anion (Tf2N(-)). According to these calculations, the four Cl(-) complexation reactions are favored and UO2Cl4(2-) is the most stable chloride complex in [BMI][Tf2N]. The solvation of the different chloro-complexes is found to evolve from purely anionic (ca. 5 Tf2N(-) ions around UO2(2+)) to purely cationic (ca. 8.5 BMI (+) cations around UO2Cl4(2-)), with onion-type alternation of solvent shells. We next compare the solvation of the UO2Cl4(2-) complex to its reduced analogue UO2Cl4(3-) in the [BMI][Tf2N] and [MeBu3N][Tf2N] liquids that possess the same anion, but differ by their cation (imidazolium BMI(+) versus ammonium MeBu3N(+)). The overall solvation structure of both complexes is found to be similar in both liquids with a first solvation shell formed exclusively of solvent cations (about 9 BMI(+) cations or 7 MeBu3N(+) cations). However, a given complex is better solvated by the [BMI][Tf2N] liquid, due to hydrogen bonding interactions between Cl(-) ligands and imidazolium-ring C-H protons. According to free energy calculations, the gain in solvation energy upon reduction of UO2Cl4(2-) to UO2Cl4(3-) is found to be larger in [BMI][Tf2N] than in [MeBu3N][Tf2N], which is fully consistent with recent experimental results (Inorg. Chem. 2006, 45, 10419).

20.
J Phys Chem A ; 112(11): 2428-36, 2008 Mar 20.
Article in English | MEDLINE | ID: mdl-18269268

ABSTRACT

Mixed uranyl aquo chloro complexes of the type [UO2(H2O)xCly]2-y (y = 1, 2, 3, 4; x + y = 4, 5) have been optimized at the BLYP, BP86, and B3LYP levels of density functional theory in vacuo and in a polarizable continuum modeling bulk water (PCM) and have been studied at the BLYP level with Car-Parrinello molecular dynamics (MD) simulations in the gas phase and in explicit aqueous solution. Free binding energies were evaluated from static PCM data and from pointwise thermodynamic integration involving constrained MD simulations in water. The computations reveal significant solvent effects on geometric and energetic parameters. Based on the comparison of PCM-optimized or MD-averaged uranyl-ligand bond distances with EXAFS-derived values, the transition between five- and four-coordination about uranyl is indicated to occur at a Cl content of y = 2 or 3.

SELECTION OF CITATIONS
SEARCH DETAIL