Aurora E. Clark

Find an error

Name: Clark, Aurora
Organization: Washington State University , USA
Department: Department of Chemistry
Title: (PhD)

TOPICS

Co-reporter:Tiecheng Zhou, Alex McCue, Yasaman Ghadar, Imre Bakó, and Aurora E. Clark
The Journal of Physical Chemistry B September 28, 2017 Volume 121(Issue 38) pp:9052-9052
Publication Date(Web):September 5, 2017
DOI:10.1021/acs.jpcb.7b07406
Using a unique combination of slab-layering analyses and identification of truly interfacial molecules, this work examines water/vapor and water/n-hexane interfaces, specifically the structural and dynamic perturbations of the interfacial water molecules at different locations within the surface capillary waves. From both the structural and dynamic properties analyzed, it is found that these interfacial water molecules dominate the perturbations within the interfacial region, which can extend deep into the water phase relative to the Gibbs dividing surface. Of more importance is the demonstration of structural and dynamic heterogeneity of the interfacial water molecules at the capillary wave front, as indicated by the dipole orientation and the structural and dynamic behavior of hydrogen bonds and their networks.
Co-reporter:Tiecheng Zhou, Peng Bai, J. Ilja Siepmann, and Aurora E. Clark
The Journal of Physical Chemistry C October 12, 2017 Volume 121(Issue 40) pp:22015-22015
Publication Date(Web):September 19, 2017
DOI:10.1021/acs.jpcc.7b04991
The properties of confined water are relevant to many chemical, geological, and biological phenomena, where they underpin essential changes to molecular scale reactivity–perturbing both the energetic and configurational landscape. Though much prior literature has focused on hydrophilic confinement, the hydrophobic confinement of water is less well understood. Here, we use molecular dynamics simulations to investigate the structures and dynamics of water in hydrophobic all-silica zeolites that have sequentially smaller pore dimensions. Of special interest is the role that pure geometric restriction imparts, relative to the rugged potential energy landscape for water interacting with the atomistic pore surface. These two effects were studied via the hydrogen bond dynamics, specifically the rates and mechanisms of hydrogen bond breakage and formation. Measuring the dynamic features as a function of scaling the water:zeolite interaction energy revealed that geometric restriction is responsible for 67%–86% of the total perturbations to water upon confinement in MFI (depending on the property) while the water:surface interactions are responsible for 14%–33%. The relative magnitude of the interaction of water with the pore surface was confirmed by second order Møller–Plesset perturbation theory. Thus, in a highly confined environment, the weak water-surface interaction should not be neglected—even in hydrophobic adsorbents to which zeolites and other materials like carbon nanotubes belong.
Co-reporter:John W. Freiderich, Adam G. Burn, Leigh R. Martin, Kenneth L. Nash, and Aurora E. Clark
Inorganic Chemistry May 1, 2017 Volume 56(Issue 9) pp:4788-4788
Publication Date(Web):April 14, 2017
DOI:10.1021/acs.inorgchem.6b02369
The equilibrium constants for [NpO2·M]4+ (M = Al3+, In3+, Sc3+, Fe3+) in μ = 10 M nitric acid and [NpO2·Ga]4+ in μ = 10 M hydrochloric acid media have been determined. The trend in the interaction strength follows: Fe3+ > Sc3+ ≥ In3+ > Ga3+ ≫ Al3+. These equilibrium constants are compared to those of previously reported values for NpO2+ complexes with Cr3+ and Rh3+ within the literature. Thermodynamic parameters and bonding modes are discussed, with density functional theory and natural bond orbital analysis indicating that the NpO2+ dioxocation acts as a π-donor with transition-metal cations and a σ-donor with group 13 cations. The small changes in electron-donating ability is modulated by the overlap with the coordinating metal ion’s valence atomic orbitals.
Co-reporter:Morgan P. Kelley, Ping Yang, Sue B. Clark, and Aurora E. Clark
Inorganic Chemistry 2016 Volume 55(Issue 10) pp:4992
Publication Date(Web):April 27, 2016
DOI:10.1021/acs.inorgchem.6b00477
The geometric and electronic structures of the 9-coordinate Cm3+ ion solvated with both water and methanol are systematically investigated in the gas phase at each possible solvent-shell composition and configuration using density functional theory and second-order Møller–Plesset perturbation theory. Ab initio molecular dynamics simulations are employed to assess the effects of second and third solvent shells on the gas-phase structure. The ion–solvent dissociation energy for methanol is greater than that of water, potentially because of increased charge donation to the ion made possible by the electron-rich methyl group. Further, the ion–solvent dissociation energy and the ion–solvent distance are shown to be dependent on the solvent-shell composition. This has implications for solvent exchange, which is generally the rate-limiting step in complexation reactions utilized in the separation of curium from complex metal mixtures that derive from the advanced nuclear fuel cycle.
Co-reporter:Yasaman Ghadar, Shauna L. Christensen, Aurora E. Clark
Fluid Phase Equilibria 2016 Volume 407() pp:126-134
Publication Date(Web):15 January 2016
DOI:10.1016/j.fluid.2015.07.013
Ionic strength of an aqueous solution is often used to alter the efficacy of industrial processes that rely upon liquid:liquid phase boundaries (e.g. solvent extraction), however there is little understanding of how this condition may alter the properties of the phase boundary itself. The current work examines the interfacial structure and processes within the biphasic system (NaNO3)aq:n-hexane from 0 to 10 M NaNO3 at 298 K and 1 atm pressure. The extent of ion-pairing was quantified, alongside the primary solvation environments of both water and ions. Ion concentration was found to be depressed in the interfacial region relative to the bulk, and as such there is less likelihood of the formation of contact ion-pairs or ion-clusters at the interface. This in turn, leads to the interesting observation of more hydrogen bonds per water at the interface than in the bulk. At 10 M NaNO3 enough ions are present at the interface to severely perturb the mesoscopic interfacial properties of width and tension (the latter doubling in size relative to the neat water interface with n-hexane). Chemical theories, like the Kelvin equation and its analogs for liquid:vapor interfaces, would intuit that such a large change in interfacial tension should influence the microsolvation processes of the immiscible solvents (microsolvation being the rare event where water can penetrate n-hexane and n-hexane may penetrate and be fully solvated by water). However, there is no statistical difference between the concentrations of water and n-hexane in their respective co-solvents as a function of interfacial tension in these non-ideal solutions. Based upon these data, the ability of electrolyte concentration to alter transport across the interface does not appear to be related to the perturbation imparted by the electrolyte upon the interfacial properties themselves.
Co-reporter:Aurora E. Clark, Alex Samuels, Katy Wisuri, Sarah Landstrom, and Tessa Saul
Inorganic Chemistry 2015 Volume 54(Issue 13) pp:6216-6225
Publication Date(Web):June 19, 2015
DOI:10.1021/acs.inorgchem.5b00365
The aqueous solvation of U–Pu in the III–VI oxidation states has been examined using density functional theory and hydrated cluster models of the form An(H2O)304+/3+ and AnO2(H2O)302+/+ embedded within a polarizable continuum model to approximate the effect of bulk water. The structural features are compared to available data from extended X-ray absorption fine structure. Then, using a multiple-scattering approach, the X-ray absorption near-edge spectra (XANES) have been simulated and compared to experiment. These structural data are complemented by a detailed thermodynamic analysis using a recently benchmarked protocol. The structural, spectroscopic, and thermodynamic information has been used to assign the primary solvation environments in water, with an emphasis upon understanding how oxidation state and position in the period modifies the hydration number and equilibrium between different solvation shell environments. Tetravalent U is proposed to exist in equilibrium between the 8- and 9-coordinate species. Moving to the right of the period, Np(IV) and Pu(IV) exist solely as the octa-aquo species. Reduction to the trivalent ions leads to thermodynamic favorability for this solvation environment, whose features reproduce the XANES spectra. The actinyl dications (AnO22+) of U and Np have a preferred environment in the equatorial plane consisting of 5 solvating waters; however, changes to the ionic radius and electronic structure at Pu leads to an equilibrium between the 4- and 5-coordinate species for PuO22+. Reduction of the dications to form the monocations generally leads to a preference for the 4-coordinate primary solvation shell, with an equilibrium existing for uranyl, while the neptunyl and plutonyl species exist solely as AnO2(H2O)4+. These data provide accurate thermodynamic information for several rare species and the combined thermodynamic, structural, and spectroscopic approach reveals trends in hydration behavior across actinide oxidation states and within the early actinide period.
Co-reporter:Payal Parmar, Alex Samuels, and Aurora E. Clark
Journal of Chemical Theory and Computation 2015 Volume 11(Issue 1) pp:55-63
Publication Date(Web):December 8, 2014
DOI:10.1021/ct500530q
Contributing factors to the solution-phase correction to the free energy of the molecular clusters U(H2O)n3+/4+ and UO2(H2O)m1+/2+ (n = 8, 9, 30, 41, 77; m = 4, 5, 30, 41, 77) have been examined as a function of cavity type in the integrated-equation-formalism-protocol (IEF) and SMD polarizable continuum models (PCMs). It is observed that the free energy correction, Gcorr, does not smoothly converge to zero as the number of explicitly solvating water molecules approaches the bulk limit, and the convergence behavior varies significantly with cavity and model. The rates of convergence of the gas-phase hydration energy, ΔGhyd, wherein the bare metal ion is inserted into a molecular water cluster and ΔGcorr for the reaction exhibit wide variations as a function of ion charge, cavity, and model. This is the likely source of previously reported discrepancies in predicted free energies of solvation for metal ions when using different PCM cavities and/or models. The cancellation of errors in ΔGhyd and ΔGcorr is optimal for clusters consisting of only a second solvation shell of explicit water molecules (n = m = 30). The UFF cavity within IEF, in particular, exhibits the most consistent cancellation of errors when using a molecular cluster consisting of a second shell of solvating water for all oxidation states of uranium, leading to accurate free energies of solvation ΔGsolv for these species.
Co-reporter:Yasaman Ghadar and Aurora E. Clark  
Physical Chemistry Chemical Physics 2015 vol. 17(Issue 25) pp:16646-16646
Publication Date(Web):08 Jun 2015
DOI:10.1039/C5CP90093J
Correction for ‘Intermolecular network analysis of the liquid and vapor interfaces of pentane and water: microsolvation does not trend with interfacial properties’ by Yasaman Ghadar et al., Phys. Chem. Chem. Phys., 2014, 16, 12475–12487.
Co-reporter:Morgan Kelley, Amber Donley, Sue Clark, and Aurora Clark
The Journal of Physical Chemistry B 2015 Volume 119(Issue 51) pp:15652-15661
Publication Date(Web):December 7, 2015
DOI:10.1021/acs.jpcb.5b07492
Ion pairing can have profound effects upon the ionic strength of electrolyte solutions but is poorly understood in solutions containing more than one solvent. Herein a combined density functional theory and molecular dynamics approach is used to examine the effect of both methanol concentration and interionic distance upon the structure and dynamics within successive solvation shells of Na+ and Cl– in water/methanol binary solutions. The structure and dynamics of the first and second solvation shells were studied along a reaction coordinate associated with ion pair formation using potential of mean force simulations. The lifetimes of the solvent–solvent hydrogen bonds become perturbed when the second solvation shells of the ions begin to interact. In contrast, the structural properties within the first and second solvation shells of the ions were found to be largely independent of both methanol concentration and interionic distance until a contact ion pair is formed. Thus, as the ions are brought together, the effect of the opposing ion manifests itself in the solvation dynamics before any structural changes are observed. As anticipated based upon the decreased dielectric constant of the binary solution, ion pair formation becomes energetically more favorable as the concentration of methanol increases.
Co-reporter:Alex C. Samuels, Cherilynn A. Boele, Kevin T. Bennett, Sue B. Clark, Nathalie A. Wall, and Aurora E. Clark
Inorganic Chemistry 2014 Volume 53(Issue 23) pp:12315-12322
Publication Date(Web):November 12, 2014
DOI:10.1021/ic501408r
Platinum group metals (PGMs), including rhodium, generated by the fission of 235U are present in significant quantities within spent nuclear fuel located on power generation sites in the United States, the amount of which is expected to exceed natural reserves by 2030. Yet, spent fuel raffinates are highly acidic media that may result in complex speciation of the PGM. This work provides an understanding of Rh(III) speciation up to 9 M HCl and HNO3, and utilizes a combination of ultraviolet–visible (UV-vis) and capillary zone electrophoresis data, along with computationally predicted thermochemistry and simulated UV-vis spectra to approximate the relative concentrations of potential species in solution as a function of acid concentration. One Rh(III) species, [Rh(NO3)3], is observed under all conditions in HNO3 and for Rh(III) concentrations smaller than 10–3 M. In contrast, a variety of chloridated Rh(III) species may exist simultaneously in a HCl medium. The species [RhCl2(H2O)4]+ and [RhCl3(H2O)3] are observed in HCl solutions of concentrations ranging from 0 to 1 M; the species [RhCl4(H2O)2]−, [RhCl5(H2O)]2–, and [Rh2Cl9]3– are observed between 2 and 9 M HCl.
Co-reporter:Xiaoning Yang and Aurora E. Clark
Inorganic Chemistry 2014 Volume 53(Issue 17) pp:8930-8940
Publication Date(Web):August 21, 2014
DOI:10.1021/ic5006659
A combined density functional theory and molecular dynamics study has been used to study reactions relevant to the crystallization of a model cluster based upon the metastable phase NH2-MOF-235(Al), which has been previously shown to be an important intermediate in the synthesis of NH2-MIL-101(Al). The clusters studied were of the form Al3O(BDC)6(DMF)n(H2O)m+, where BDC– = NH2-benzenedicarboxylate and DMF = dimethylformamide (n = 1–3; m = {n – 3}). The ionic bonding interaction of the Al3O7+ core with BDC– is much stronger than that with a coordinated solvent and is independent of the bulk solvent medium (water or DMF). The exchange reactions of a coordinated solvent are predicted to be facile, and the dynamic solvent organization indicates that they are kinetically allowed because of the ability of the solvent to migrate into the cleft created by the BDC–Al3O–BDC coordination angle. As BDC– binds to the Al3O7+ core, the solvation free energy (Gsolv) of the cluster becomes less favorable, presumably because of the overall hydrophobicity of the cluster. These data indicate that as the crystal grows there is a balance between the energy gained by BDC– coordination and an increasingly unfavorable Gsolv. Ultimately, unfavorable solvation energies will inhibit the formation of quantifiable metal–organic framework (MOF) crystals unless solution-phase conditions can be used to maintain thermodynamically favorable solute–solvent interactions. Toward this end, the addition of a cosolvent is found to alter solvation of Al3O(BDC)6(DMF)3+ because more hydrophobic solvents (DMF, methanol, acetonitrile, and isopropyl alcohol) preferentially solvate the MOF cluster and exclude water from the immediate solvation shells. The preferential solvation is maintained even at temperatures relevant to the hydrothermal synthesis of MOFs. While all cosolvents exhibit this preferential solvation, trends do exist. Ranking the cosolvents based upon their observed ability to exclude water from the MOF cluster yields acetonitrile < DMF ∼ methanol < isopropyl alcohol. These observations are anticipated to impact the intermediate and final phases observed in MOF synthesis by creating favorable solvation environments for specific MOF topologies. This adds further insight into recent reports wherein DMF has been implicated in the reactive transformation of NH2-MOF-235(Al) to NH2-MOF-101(Al), suggesting that that DMF additionally plays a vital role in stabilizing the metastable NH2-MOF-235(Al) phase early in the synthesis.
Co-reporter:Yasaman Ghadar and Aurora E. Clark  
Physical Chemistry Chemical Physics 2014 vol. 16(Issue 24) pp:12475-12487
Publication Date(Web):01 May 2014
DOI:10.1039/C4CP00602J
Liquid:vapor and liquid:liquid interfaces exhibit complex organizational structure and dynamics at the molecular level. In the case of water and organic solvents, the hydrophobicity of the organic, its conformational flexibility, and compressibility, all influence interfacial properties. This work compares the interfacial tension, width, molecular conformations and orientations at the vapor and aqueous liquid interfaces of two solvents, n-pentane and neopentane, whose varying molecular shapes can lead to significantly different interfacial behavior. Particular emphasis has been dedicated toward understanding how the hydrogen bond network of water responds to the pentane relative to the vapor interface and the sensitivity of the network to the individual pentane isomer and system temperature. Interfacial microsolvation of the immiscible solvents has been examined using graph theoretical methods that quantify the structure and dynamics of microsolvated species (both H2O in C5H12 and C5H12 in H2O). At room temperature, interfacial water at the pentane phase boundary is found to have markedly different organization and dynamics than at the vapor interface (as indicated by the hydrogen bond distributions and hydrogen bond persistence in solution). While the mesoscale interfacial properties (e.g. interfacial tension) are sensitive to the specific pentane isomer, the distribution and persistence of microsolvated species at the interface is nearly identical for both systems, irrespective of temperature (between 273 K and 298 K). This has important implications for understanding how properties defined by the interfacial organization are related to the underlying solvation reactions that drive formation of the phase boundary.
Co-reporter:Chun-Hung Wang ; Peng Bai ; J. Ilja Siepmann
The Journal of Physical Chemistry C 2014 Volume 118(Issue 34) pp:19723-19732
Publication Date(Web):August 7, 2014
DOI:10.1021/jp502867v
Essential topological indices of the hydrogen-bond networks of water, methanol, ethanol, and their binary mixtures adsorbed in microporous silicalite-1 (a hydrophobic zeolite with potential application for biofuel processing) are analyzed and compared to their bulk liquid counterparts. These include the geodesic distribution (the shortest H-bond pathways between molecular vertices), the average length, the geodesic index, the orientation and distance of the adsorbate to the interior of the zeolite, and the sorbate–sorbate and sorbate–sorbent distributions of H-bonds. In combination, they describe how the H-bond networks are altered when going from the bulk to the confined silicalite-1 environment. The speciation of the adsorbed compounds is quantified in terms of their network connectivity, revealing that pure water has a high probability of forming long, contiguous H-bonded chains in silicalite-1 at high loading, while alcohols form small dimeric/trimeric clusters. The extent to which the H-bond network of binary water–alcohol systems is altered relative to either unary system is quantified, demonstrating an enhanced interconnectivity that is reflected in the tendency of individual H2O molecules to become co-adsorbed with alcohol clusters in the zeolite framework. Selectivity for the alcohol over water diminishes with increasing alcohol loading as the H-bonded clusters serve as favorable adsorption sites for H2O.
Co-reporter:Abdullah Ozkanlar, Alex Samuels, Aurora E. Clark
Chemical Physics Letters 2013 560() pp: 10-14
Publication Date(Web):
DOI:10.1016/j.cplett.2013.01.017
Co-reporter:Payal Parmar, Kirk A. Peterson, and Aurora E. Clark
The Journal of Physical Chemistry A 2013 Volume 117(Issue 46) pp:11874-11880
Publication Date(Web):May 16, 2013
DOI:10.1021/jp403078j
High-quality static electric dipole polarizabilities have been determined for the ground states of the hard-sphere cations of U, Np, and Pu in the III and IV oxidation states. The polarizabilities have been calculated using the numerical finite field technique in a four-component relativistic framework. Methods including Fock-space coupled cluster (FSCC) and Kramers-restricted configuration interaction (KRCI) have been performed in order to account for electron correlation effects. Comparisons between polarizabilities calculated using Dirac–Hartree–Fock (DHF), FSCC, and KRCI methods have been made using both triple- and quadruple-ζ basis sets for U4+. In addition to the ground state, this study also reports the polarizability data for the first two excited states of U3+/4+, Np3+/4+, and Pu3+/4+ ions at different levels of theory. The values reported in this work are the most accurate to date calculations for the dipole polarizabilities of the hard-sphere tri- and tetravalent actinide ions and may serve as reference values, aiding in the calculation of various electronic and response properties (for example, intermolecular forces, optical properties, etc.) relevant to the nuclear fuel cycle and material science applications.
Co-reporter:Natalia L. Zakharova;Christina L. Crawford
Journal of The American Society for Mass Spectrometry 2012 Volume 23( Issue 5) pp:792-805
Publication Date(Web):2012 May
DOI:10.1007/s13361-012-0339-5
When utilized in conjunction with modeling, the collision cross section (Ω) from ion mobility spectrometry can be used to deduce the gas phase structures of analyte ions. Gas phase conformations are determined computationally, and their Ω calculated using an approximate method, the results of which are compared with experimental data. Though prior work has focused upon rigid small molecules or large biomolecules, correlation of computational and experimental Ω has not been thoroughly examined for analytes with intermediate conformational flexibility, which constitute a large fraction of the molecules studied in the field. Here, the computational paradigm for calculating Ω has been tested for the tripeptides WGY, YGW, and YWG (Y = tyrosine, W = tryptophan, G = glycine). Experimental data indicate that Ωexp (YWG) > Ωexp (WGY) ≈ Ωexp (YGW). The energy distributions of conformations obtained from tiers of simulated annealing molecular dynamics (SAMD) were analyzed using a wide array of density functionals. These quantum mechanical energy distributions do not agree with the MD data, which leads to structural differences between the SAMD and DFT conformations. The latter structures are obtained by reoptimization of the SAMD geometries, and are the only suite of structures that reproduce the experimental trend in analyte separability. In the absence of fitting Lennard Jones potentials that reproduce experimental results for the Trajectory Method, the Exact Hard Sphere Scattering method produced numerical values that are in best agreement with the experimental cross sections obtained in He drift gas.
Co-reporter:Matthew Hudelson;Barbara Logan Mooney
Journal of Mathematical Chemistry 2012 Volume 50( Issue 9) pp:2342-2350
Publication Date(Web):2012 October
DOI:10.1007/s10910-012-0033-7
Polyhedral representations of the geometric arrangements of atoms and molecules is a pervasive tool in chemistry for understanding chemical bonding and electrostatic interactions. Yet the structural organization within very large systems is often difficult to quantify. In this work, we illustrate that PageRank, when combined with the chemical constraints of a system, can be used to uniquely identify the polyhedral arrangements of atoms and molecules. The PageRank algorithm can be used on any network that can be represented as a graph: a mathematical object where individual points, or vertices, are joined by edges. It is thus well-suited for chemical systems where atoms (considered vertices) are connected to each other via chemical bonding (considered edges) or other forces. This has been implemented in a recently reported series of R-scripts, moleculaRnetworks, and the example provided herein illustrates that the polyhedral arrangement of solvent molecules about a solute results in a unique PR value for the solute and enables rapid identification of the local geometry in the condensed medium. More generally PR can be used as a chemoinformatic tool to search for specific structural patterns within any database of geometric configurations.
Co-reporter:Barbara Logan Mooney, L. Rene Corrales, and Aurora E. Clark
The Journal of Physical Chemistry B 2012 Volume 116(Issue 14) pp:4263-4275
Publication Date(Web):March 14, 2012
DOI:10.1021/jp300193j
A new method for analyzing molecular dynamics simulation data is employed to study the solvent shell structure and exchange processes of mono-, di-, and trivalent metal cations in water. The instantaneous coordination environment is characterized in terms of the coordinating waters’ H-bonding network, orientations, mean residence times, and the polyhedral configuration. The graph-theory-based algorithm provides a rapid frame-by-frame identification of polyhedra and reveals fluctuations in the solvation shell shape—previously unexplored dynamic behavior that in many cases can be associated with the exchange reactions of water between the first and second solvation shells. Extended solvation structure is also analyzed graphically, revealing details of the hydrogen bonding network that have practical implications for connecting molecular dynamics data to ab initio cluster calculations. Although the individual analyses of water orientation, residence time, etc., are commonplace in the literature, their combination with graphical algorithms is new and provides added chemical insight.
Co-reporter:Abdullah Ozkanlar, Jonathan L. Cape, James K. Hurst, and Aurora E. Clark
Inorganic Chemistry 2011 Volume 50(Issue 17) pp:8177-8187
Publication Date(Web):August 10, 2011
DOI:10.1021/ic200646h
Density functional theory (DFT) has been used to investigate the plausibility of water addition to the simple mononuclear ruthenium complexes, [(NH3)3(bpy)Ru═O]2+/3+ and [(NH3)3(bpy)RuOH]3+, in which the OH fragment adds to the 2,2′-bipyridine (bpy) ligand. Activation of bpy toward water addition has frequently been postulated within the literature, although there exists little definitive experimental evidence for this type of “covalent hydration”. In this study, we examine the energetic dependence of the reaction upon metal oxidation state, overall spin state of the complex, as well as selectivity for various positions on the bipyridine ring. The thermodynamic favorability is found to be highly dependent upon all three parameters, with free energies of reaction that span favorable and unfavorable regimes. Aqueous addition to [(NH3)3(bpy)Ru═O]3+ was found to be highly favorable for the S = 1/2 state, while reduction of the formal oxidation state on the metal center makes the reaction highly unfavorable. Examination of both facial and meridional isomers reveals that when bipyridine occupies the position trans to the ruthenyl oxo atom, reactivity toward OH addition decreases and the site preferences are altered. The electronic structure and spectroscopic signatures (EPR parameters and simulated spectra) have been determined to aid in recognition of “covalent hydration” in experimental systems. EPR parameters are found to uniquely characterize the position of the OH addition to the bpy as well as the overall spin state of the system.
Co-reporter:Jadwiga Kuta ; Matthew C. F. Wander ; Zheming Wang ; Siduo Jiang ; Nathalie A. Wall
The Journal of Physical Chemistry C 2011 Volume 115(Issue 43) pp:21120-21127
Publication Date(Web):September 15, 2011
DOI:10.1021/jp204633g
Molecular dynamics simulations were performed to examine trends in trivalent lanthanide [Ln(III)] sorption to ≡SiOH0 and ≡SiO– sites on the 001 surface of α-quartz across the 4f period. Complementary laser-induced fluorescence studies examined Eu(III) sorption to α-quartz at a series of ionic strengths from 1 × 10–4 M to 0.5 M such that properties of the surface-sorbed species could be extrapolated to zero ionic strength, the conditions under which the simulations are performed. Such extrapolation allows for a more direct comparison of the data and enables a molecular understanding of the surface-sorbed species and the role of the ion surface charge density upon the interfacial reactivity. Potential of mean force molecular dynamics as well as simulations of presorbed Ln(III) species agrees with the spectroscopic study of Eu(III) sorption, indicating that strongly bound inner-sphere complexes are formed upon sorption to an ≡SiO– site. The coordination shell of the ion contains 6–7 waters of hydration, and it is predicted that surface silanol OH groups transfer from the quartz to the inner coordination shell of Eu(III). Molecular simulations predict less-strongly bound inner-sphere species in early lanthanides and more strongly bound species in late lanthanides, following trends in the surface charge density of the 4f ions. Hydroxyl ligands that derive from the surface silanol groups are consistently observed to bind in the inner coordination shell of surface-sorbed inner-sphere Ln(III) ions, provided that the ion is able to migrate within 2.0–3.0 Å of the plane formed by the silanol O atoms (∼3.5 Å from an individual ≡SiO– group). Sorption to a fully protonated quartz surface is not predicted to be favorable by any Ln(III), except perhaps Lu. The present work demonstrates a combined theoretical and experimental approach in the prediction of the fate of trivalent radioactive contaminants at temporary and permanent nuclear waste storage sites.
Co-reporter:Jadwiga Kuta and Aurora E. Clark
Inorganic Chemistry 2010 Volume 49(Issue 17) pp:7808-7817
Publication Date(Web):August 2, 2010
DOI:10.1021/ic100623y
The geometric and electronic structures, as well as the thermodynamic properties of trivalent lanthanide hydrates {Ln(H2O)8,93+ and Ln(H2O)8,9(H2O)12,143+, Ln = La−Lu} have been examined using unrestricted density functional theory (UDFT), unrestricted Möller-Plesset perturbation theory (UMP2), and multiconfigurational self-consistent field methods (MCSCF). While Ln-hydrates with 2−5 unpaired f-electrons have some multiconfigurational character, the correlation energy lies within 5−7 kcal/mol across the period and for varying coordination numbers. As such DFT yields structural parameters and thermodynamic data quite close to experimental values. Both UDFT and UMP2 predict free energies of water addition to the Ln(H2O)83+ species to become less favorable across the period; however, it is a non-linear function of the surface charge density of the ion. UDFT further predicts that the symmetry of the metal-water bond lengths is sensitive to the specific f-electron configuration, presumably because of repulsive interactions between filled f-orbitals and water lone-pairs. Within the Ln(H2O)8,9(H2O)12,143+ clusters, interactions between solvation shells overrides this orbital effect, increasing the accuracy of the geometric parameters and calculated vibrational frequencies. Calculated atomic charges indicate that the water ligands each donate 0.1 to 0.2 electrons to the Ln(III) metals, with increasing electron donation across the period. Significant polarization and charge transfer between solvation shells is also observed. The relationship between empirical effective charges and calculated atomic charges is discussed with suggestions for reconciling the trends across the period.
Co-reporter:Matthew C. F. Wander
Inorganic Chemistry 2008 Volume 47(Issue 18) pp:8233-8241
Publication Date(Web):August 21, 2008
DOI:10.1021/ic800750g
Using density functional theory and polarized continuum models, we have determined the most probable coordination number and structure of the first hydration shell of aqueous Pb(II). The geometries and hydration free energies of Pb(H2O)1−92+ were examined and benchmarked against experimental values. The free energies of hydration of Pb(H2O)6−82+ were found to match the experimental value within 10 kcal/mol. Moreover, based upon our thermochemical results for single water addition, primary hydration numbers of 6, 7, and 8 are all thermally accessible at STP. Use of a small-core 60 electron effective core potential (ECP) with the aug-cc-pvdz-PP basis on Pb resulted in structures that are significantly less hemidirected than predicted when using the large-core 78 electron ECP and the lanl2DZ basis on the metal. Our results imply that the hemi- to holo-directed transition in Pb(II)-water complexes is driven by coordination number and not hybridization of the 6s lone-pair orbital or enhanced covalent bonding in the Pb−OH2 bond. In addition to basis set effects, the influence of different solvation models on hydration reactions has further been examined so as to determine the relative accuracy of the calculated hydration thermochemistry.
Co-reporter:Adriana Dinescu and Aurora E. Clark
The Journal of Physical Chemistry A 2008 Volume 112(Issue 44) pp:11198-11206
Publication Date(Web):October 10, 2008
DOI:10.1021/jp8076408
With a single f-electron, Ce(III) is the simplest test case for benchmarking the thermodynamic and structural properties of hydrated Ln(III) against varying density functionals and reaction field models, in addition to determining the importance of multiconfigurational character in their wave functions. Here, the electronic structure of Ce(H2O)x(H2O)y3+ (x = 8, 9; y = 0, 12−14) has been examined using DFT and CASSCF calculations. The latter confirmed that the wave function of octa- and nona-aqua Ce(III) is well-described by a single configuration. Benchmarking was performed for density functionals, reaction field cavity types, and solvation reactions against the experimental free energy of hydration, ΔGhyd(Ce3+). The UA0, UAKS, Pauling, and UFF polarized continuum model cavities displayed different performance, depending on whether one or two hydration shells were examined, and as a function of the size of the metal basis set. These results were essentially independent of the density functional employed. Using these benchmarks, the free energy for water exchange between CN = 8 and CN = 9, for which no experimental data are available, was estimated to be approximately −4 kcal/mol.
Co-reporter:Matthew C. F. Wander and Aurora E. Clark
The Journal of Physical Chemistry C 2008 Volume 112(Issue 50) pp:19986-19994
Publication Date(Web):2017-2-22
DOI:10.1021/jp803642c
The structure, orientation, and dielectric of water at the quartz|water interface has been examined under different hydration levels using classical molecular dynamics. The properties of 1H2O/10 Å2, 2H2O/10 Å2, 4H2O/10 Å2, and bulk water on quartz have been benchmarked against experimental data. Structurally, the simulations match existing sum-frequency spectroscopy data, which indicate the existence and orientation of both frozen and loosely bound water on the quartz surface. Good agreement has also been found with existing experimental dielectric data for the 1H2O/10 Å2 level of hydration, and a clear difference has been found in the values of εs = 48, ε∥ = 48, and ε⊥ = 40 for the first slice of a bulk-water−solid interface and εs= 30, ε∥= 30, and ε⊥= 10 for that of 1H2O/10 Å2 water coverage. Overall there is a fundamental difference in shielding between a single interface and the 1H2O/10 Å2 level of hydration.
Co-reporter:Yasaman Ghadar and Aurora E. Clark
Physical Chemistry Chemical Physics 2015 - vol. 17(Issue 25) pp:NaN16646-16646
Publication Date(Web):2015/06/08
DOI:10.1039/C5CP90093J
Correction for ‘Intermolecular network analysis of the liquid and vapor interfaces of pentane and water: microsolvation does not trend with interfacial properties’ by Yasaman Ghadar et al., Phys. Chem. Chem. Phys., 2014, 16, 12475–12487.
Co-reporter:Yasaman Ghadar and Aurora E. Clark
Physical Chemistry Chemical Physics 2014 - vol. 16(Issue 24) pp:NaN12487-12487
Publication Date(Web):2014/05/01
DOI:10.1039/C4CP00602J
Liquid:vapor and liquid:liquid interfaces exhibit complex organizational structure and dynamics at the molecular level. In the case of water and organic solvents, the hydrophobicity of the organic, its conformational flexibility, and compressibility, all influence interfacial properties. This work compares the interfacial tension, width, molecular conformations and orientations at the vapor and aqueous liquid interfaces of two solvents, n-pentane and neopentane, whose varying molecular shapes can lead to significantly different interfacial behavior. Particular emphasis has been dedicated toward understanding how the hydrogen bond network of water responds to the pentane relative to the vapor interface and the sensitivity of the network to the individual pentane isomer and system temperature. Interfacial microsolvation of the immiscible solvents has been examined using graph theoretical methods that quantify the structure and dynamics of microsolvated species (both H2O in C5H12 and C5H12 in H2O). At room temperature, interfacial water at the pentane phase boundary is found to have markedly different organization and dynamics than at the vapor interface (as indicated by the hydrogen bond distributions and hydrogen bond persistence in solution). While the mesoscale interfacial properties (e.g. interfacial tension) are sensitive to the specific pentane isomer, the distribution and persistence of microsolvated species at the interface is nearly identical for both systems, irrespective of temperature (between 273 K and 298 K). This has important implications for understanding how properties defined by the interfacial organization are related to the underlying solvation reactions that drive formation of the phase boundary.
Pyridine, 2-[1-(phenylmethyl)-1H-1,2,3-triazol-4-yl]-
2-(1-METHYLIMIDAZOL-2-YL)PYRIDINE
1H-1,2,3-Triazole-4-carboxylic acid, 1-(phenylmethyl)-
IRON, COMPD. WITH MAGNESIUM (1:1)
Magnesium, compd. with titanium (1:1)
Propanenitrile, methoxy-
protium
Hydroxyl