Physical Chemistry Chemical Physics

Published by Royal Society of Chemistry
Online ISSN: 1463-9084
Print ISSN: 1463-9076
Publications
TiO(2) nanoparticles of different phases play a key role in property alteration of nanocomposite fibers. Polycaprolactone (PCL)/TiO(2) composite fibers were prepared using the electrospinning method. Pure anatase and rutile phases were synthesized using the sol-gel route for nanocomposite synthesis. The Effect of nanoparticle phases on crystallinity of fibers and interaction with polymer molecules have been studied using X-ray diffraction (XRD), Fourier transform infrared (FTIR) spectroscopy, Raman spectroscopy, morphology through SEM, surface properties using BET method and wetting property of fibers commencing from contact angle measurement. Biocompatibility and biodegradation of hybrid materials have been studied in simulated body fluid (SBF) and phosphate buffer (PBS), respectively. The anatase phase with smaller particle dimensions exhibited significant improvement of most of the properties as compared to composites made of the rutile phase. Better interaction between polymer chain and anatase particle PCL-A nanocomposite fibers leads to better mechanical property and biocompatibility vis-à-vis PCL-R and pristine PCL fibers. Biocompatibility of PCL nanocomposite has been testified through proliferation of fibroblast cell and its adhesion; MTT (3-(4,5-dimethythiazol-2-yl)-2,5-diphenyl tetrazolium bromide) assay demonstrates good proliferation rate for cells on PCL-A nanocomposite fibres.
 
Absorption cross sections for the à 2A″ (0,90,0) ← 2A′ (0,01,0) band of HCO were determined at 295 K using pulsed laser photolysis combined with cavity ring-down spectroscopy. Formyl radicals (HCO) were produced from the reaction of atomic chlorine, generated by photolysis of Cl2 at 335 nm, with formaldehyde. The concentration of HCO was calibrated using two independent photochemical methods. The peak cross section of the P(8) line was determined to be (1.98 ± 0.36) × 10−18 cm2, and the intensity of the entire band was normalized to this line. The quoted 2σ uncertainty includes estimated systematic errors. Comparisons to previously reported values of HCO cross sections in this band are discussed.
 
Rate coefficients for the CH(v = 0,1) + D(2) reaction have been determined for all possible channels (T: 200-1200 K), using the quasiclassical trajectory method and a suitable treatment of the zero point energy. Calculations have also been performed on the CH(v = 1) + H(2) reaction and the CH(v = 1) + D(2) → CH(v = 0) + D(2) process. Most of the results can be understood considering the key role played by the deep minimum of the potential energy surface (PES), the barrierless character of the PES, the energy of the reaction channels, and the kinematics. The good agreement found between theory and experiment for the rate coefficients of the capture process of CH(v = 0) + D(2), the total reactivity of CH(v = 1) + D(2), H(2), as well as the good agreement observed for the related CH(v = 0) + H(2) system (capture and abstraction), gives confidence on the theoretical rate coefficients obtained for the capture processes of CH(v = 1) + D(2), H(2), the individual reactive processes of CH(v = 1) + D(2), H(2), the abstraction and abstraction-exchange reactions for CH(v = 0) + D(2), and the inelastic process mentioned above, for which there are no experimental data available, and that can be useful in combustion chemistry and astrochemistry.
 
We have measured differential cross sections (DCSs) for the reaction H + D(2) → HD(v' = 2,j' = 0,3,6,9) + D at center-of-mass collision energies E(coll) of 1.25, 1.61, and 1.97 eV using the photoloc technique. The DCSs show a strong dependence on the product rotational quantum number. For the HD(v' = 2,j' = 0) product, the DCS is bimodal but becomes oscillatory as the collision energy is increased. For the other product states, they are dominated by a single peak, which shifts from back to sideward scattering as j' increases, and they are in general less sensitive to changes in the collision energy. The experimental results are compared to quantum mechanical calculations and show good, but not fully quantitative agreement.
 
A series of double molybdate scheelite-type phosphors LixAg1-xYb0.99(MoO4)2:0.01Er(3+) (x = 0, 0.1, 0.3, 0.5, 0.7, 0.9, 1.0) were synthesized by the solid state reaction method, and their crystal structures and upconversion (UC) luminescence properties were investigated in detail. The phase structure evolution of this series samples was discussed and the selected Li0.5Ag0.5Yb0.99(MoO4)2:0.01Er(3+) was analyzed based on the Rietveld refinement. The UC emission properties and the related UC mechanism were also studied. With an increasing Li/Ag ratio in this host, the UC emission intensities of LixAg1-xYb0.99(MoO4)2:0.01Er(3+) increased obviously, and the enhancement could be attributed to the coupling effect and the nonradiative transition between two energy levels of LixAg1-xYb(MoO4)2 matrices and the activator Er(3+), which have also been analyzed based on the results of the ultraviolet-visible diffuse reflection spectroscopy (UV-vis DRS) and Raman spectroscopy.
 
The local coordination of the Fe(3+)-centers in Li[Co(0.98)Fe(0.02)]O(2) cathode materials for lithium-ion batteries has been investigated by means of XRD and multi-frequency EPR spectroscopy. EPR clearly showed the Fe(3+) being in a high-spin state with S = 5/2. The set of spin-Hamiltonian parameters obtained from multi-frequency EPR experiments with Larmor frequencies ranging between 9.8 and 406 GHz was transformed into structural information by means of an expansion to standard Newton-superposition modeling, termed as Monte-Carlo Newman superposition modeling. Based on this analysis, an isovalent incorporation of the Fe(3+)-ions on the Co(3+)-sites, i.e. Fe(x)(Co), has been shown. With that respect, the positive sign of the axial second-order fine-structure interaction parameter B(0)(2) is indicative of an elongated oxygen octahedron, whereas B(0)(2) < 0 points to a compressed octahedron coordinated about the Fe(3+)-center. Furthermore, the results obtained here suggest that the oxygen octahedron about the Fe(3+)-ion is slightly distorted as compared to the CoO(6) octahedron, which in turn may impose mechanical strain to the cathode material.
 
State-of-the-art differential cross sections (DCSs) have been reported by Wang et al. [Proc. Nat. Acad. Sci. (U.S.), 2008, 105, 6227] for the state-to-state F + H(2)→ FH + H reaction using fully quantum-state-selected crossed molecular beams. We theoretically analyze the angular scattering of this reaction, in order to quantitatively understand the physical content of structure in the DCSs. Three transitions are studied, v(i)=0, j(i)=0, m(i)=0 → v(f)=3, j(f)=0, 1, 2, m(f)=0 at a translational energy of 0.04088 eV, where v, j, m are the vibrational, rotational and helicity quantum numbers respectively for the initial and final states. The input to our analyses consists of accurate quantum scattering (S) matrix elements computed for the Fu-Xu-Zhang potential energy surface, as used by Wang et al. in a computational simulation of their experimental DCSs. We prove that the pronounced peak at forward angles observed in the experimental and simulated DCSs for all three transitions is a glory. At larger angles, it is demonstrated that the 000 → 300 and 000 → 310 DCSs both possess a broad farside rainbow, which is accompanied by diffraction oscillations. We confirm the conjecture of Wang et al. that these diffraction oscillations arise from nearside-farside (NF) interference. We find that the reaction is N dominant for all three transitions. The theoretical techniques used to analyze the angular scattering include uniform semiclassical theories of glory and of rainbow scattering. We also make the first application of a semiclassical formula that is uniform for both glory + rainbow scattering. In addition, structure in the DCSs is analyzed using NF theory and local angular momentum theory, in both cases with three resummations of the partial wave series for the scattering amplitude. We make the first explicit application of the Thiele rational interpolation formula to extract the position and residue of the leading Regge pole from a set of S matrix elements, thereby making contact with complex angular momentum theories of DCSs, which interpret the angular scattering in terms of Regge resonances. Our calculations complement the exit-valley vibrationally-adiabatic analysis of Wang et al.
 
The oxidation and hydration kinetics of a proton conductor oxide, SrCe(0.95)Yb(0.05)O(2.975), were examined via conductivity relaxation upon a sudden change of oxygen activity in a fixed water-activity atmosphere, and vice versa, in the ranges of -4.0 < log a(O(2)) < or = 0.01 and -5.0 < log a(H(2)O) < -2.0 at 800 degrees C. It was found that under an oxygen-activity gradient in a fixed water-vapor-activity atmosphere, the conductivity relaxation with time is monotonic with a single relaxation time (as usual), yielding a chemical diffusivity that is unequivocally that of the component oxygen. In a water-activity gradient in a fixed oxygen activity atmosphere, on the other hand, the conductivity relaxation appears quite unusual, exhibiting an extremum after an initial transient. The conductivity relaxation upon hydration or oxidation, in general, is quantitatively analyzed in terms of two apparent chemical diffusivities for component oxygen and hydrogen, respectively. The inner workings of hydration is discussed, and the as-evaluated chemical diffusivities are reported and compared with the conventional chemical diffusivity of water.
 
Inelastic excitation and charge transfer excitation processes in collisions between neutral Mg atoms and Cs+ ions, both in their ground states, have been studied by means of a crossed molecular beam technique. Decay fluorescent emissions from Mg(3 (1)P1), Mg(3 3D(3,2,1)), Mg(4 (3)S1) and Cs(6 2P(3/2)) states have been detected as well as the phosphorescent emission due to Mg(3 (3)P1) decay. The corresponding absolute cross-sections vs. collision energy functions were determined in the laboratory 0.05-4.20 keV energy range. In order to interpret the experimental results, accurate ab initio calculations using pseudopotentials have been performed for the (Mg-Cs)+ system obtaining a manifold of adiabatic energy curves correlating with the different collision channels and allowing a qualitative interpretation of the emission excitation functions for the different processes studied.
 
An asymmetric cell based on a proton conductor, BaZr(0.1)Ce(0.7)Y(0.1)Yb(0.1)O(3-δ) (BZCYYb), with a well-defined patterned Pt electrode was prepared to study the kinetics and mechanism of the hydrogen oxidation reaction under typical conditions for fuel cell operation and hydrogen separation, including operating temperature and hydrogen partial pressure. Steady-state polarization curves were carefully analyzed to determine the apparent exchange current density, limiting current density, and charge transfer coefficients. The empirical reaction order, as estimated from the dependence of electrode polarization (R(p)) and exchange current density on the partial pressure of hydrogen (P(H(2))), varied from 0.55 to 0.71. The results indicate that hydrogen dissociation contributes the most to the rate-limiting step of the hydrogen oxidation reaction taking place at the Pt-BZCYYb interface. At high current densities, surface diffusion of electroactive species appears to contribute to the rate-limiting step as well.
 
In the present study, we have investigated structures of a CO adlayer on a well-defined Pt(100) electrode surface in 0.1 M HClO4 aqueous solutions saturated with N2, 1% CO/He and 100% CO by using in situ STM. The in situ STM images with molecular resolution demonstrated that highly ordered structures of the CO adlayer, denoted (2 × n) - 2(n - 1)CO with CO coverages of (n - 1)/n, dynamically varied with the electrode potential and the CO partial pressure in solution. As the CO partial pressure increased, more compressed structures of the CO adlayer formed on the electrode surface. In each solution, a phase transition of the CO adlayer on the terrace site was observed to be triggered by increasing the electrode potential, accompanied by a partial desorption of surface CO without charge transfer. A series of in situ STM images revealed transient local structures during the phase transition of the CO adlayer. Specifically, unique structures were found to appear in the vicinity of monoatomic steps in N2- and 1% CO/He-saturated solution, but not in 100% CO-saturated solution.
 
The electron paramagnetic resonance (EPR) properties of the electron-doped manganite La(1-x)Te(x)MnO(3) (0.1 ≤ x ≤ 0.2) are investigated based on the data of EPR spectra, resistivity, and magnetic susceptibility. With decreasing temperature from 400 K, the EPR linewidth ΔH(PP) decreases and passes through a minimum at T(min), then substantially increases with further decreasing temperature. The broadening of the EPR linewidth above T(min) can be understood in terms of the increase in the relaxation rate of spin of e(g) polarons to the lattice with increasing temperature due to the similarity between the temperature dependence of the linewidth ΔH(pp)(T) and the conductivity σ(T). For the samples with x = 0.1 and 0.15, the conductivity activation energy E(σ) is comparable with the activation energy E(a) deduced from the linewidth. Whereas for the x = 0.2 sample, there is a large difference between E(σ) (0.2206 eV) and E(a) (0.0874 eV).
 
Ferroelectric BiFeO3 has attractive properties such as high strain and polarization, but a wide range of applications of bulk BiFeO3 are hindered due to high leakage currents and a high coercive electric field. Here, we report on the thermal behaviour of the electrical conductivity and thermopower of BiFeO3 substituted with 10 and 20 mol% Bi0.5K0.5TiO3. A change from p-type to n-type conductivity in these semi-conducting materials was demonstrated by the change in the sign of the Seebeck coefficient and the change in the slope of the isothermal conductivity versus partial pressure of O. A minimum in the isothermal conductivity was observed at ∼10(-2) bar O2 partial pressure for both solid solutions. The strong dependence of the conductivity on the partial pressure of O2 was rationalized by a point defect model describing qualitatively the conductivity involving oxidation/reduction of Fe(3+), the dominating oxidation state of Fe in stoichiometric BiFeO3. The ferroelectric to paraelectric phase transition of 80 and 90 mol% BiFeO3 was observed at 648 ± 15 and 723 ± 15 °C respectively by differential thermal analysis and confirmed by dielectric spectroscopy and high temperature powder X-ray diffraction.
 
The decomposition of the cubic perovskite-type oxide Ba(x)Sr(1-x)Co(0.8)Fe(0.2)O(3-delta) (BSCF) into hexagonal and cubic perovskite-type phases has been examined by means of Scanning Electron Microscopy (SEM), Transmission Electron Microscopy (TEM), Selected Area Electron Diffraction (SAED) and X-Ray Diffraction (XRD). SEM and TEM measurements reveal that the new hexagonal phase grows predominantly at the grain boundaries of BSCF ceramics and that the cation composition of the newly formed hexagonal phase differs from that of the starting material. An orientational relationship between the hexagonal and the parent cubic phase was also observed. By means of ex situ XRD the phase fraction of the hexagonal phase was determined as a function of annealing time. A kinetic analysis of the data, based on Avrami-type kinetics, indicates that the decomposition is independent of the initial A-site composition, and the obtained reaction order supports the conclusion that the hexagonal phase grows at the grain boundaries in dense ceramic samples.
 
Photoelectrical properties of Tl1-xIn1-xSnxSe2 single crystalline alloys (x = 0, 0.1, 0.2, 0.25) grown using the Bridgman-Stockbarger method were studied. The temperature dependence of electrical and photoconductivity for the Tl1-xIn1-xSnxSe2 single crystals was explored. It has been established that photosensitivity of the Tl1-xIn1-xSnxSe2 single crystals increases with x. The spectral distribution of photocurrent in the wavelength spectral range 400-1000 nm has been investigated at various temperatures. Photoconductivity increases in all the studied crystals with temperature. Therefore, thermal activation of photoconductivity is caused by re-charging of the photoactive centers as the samples are heated. Based on our investigations, a model of center re-charging is proposed that explains the observed phenomena. X-ray photoelectron valence-band spectra for pristine and Ar(+)-ion irradiated surfaces of the Tl1-xIn1-xSnxSe2 single crystals have been measured. These results reveal that the Tl1-xIn1-xSnxSe2 single-crystal surface is sensitive to the Ar(+) ion irradiation that induced structural modification in the top surface layers. Comparison on a common energy scale of the X-ray emission Se Kβ2 bands representing energy distribution of the Se 4p-like states and the X-ray photoelectron valence-band spectra was done.
 
Experimental results show that with an increase of relative humidity, the resistance of La0.875Ca0.125FeO3 decreases at room temperature but increases at higher temperatures (140-360 °C). The humid effect at room temperature is due to the movement of H(+) or H3O(+) inside of the condensed water layer on the surface of La0.875Ca0.125FeO3. Regarding the humid effect at high temperatures, the density functional theory (DFT) calculations show that H2O can be adsorbed onto the La0.875Ca0.125FeO3 surface in the molecular and dissociative adsorption configurations, where the La0.875Ca0.125FeO3 surface gains some electrons from H2O or its dissociative products, consistent with our observation. Experimental results also show that CO2 sensing response at high temperatures decreases with an increase of room-temperature relative humidity. DFT calculations indicate that CO2 adsorbed onto the La0.875Ca0.125FeO3(010) surface, where high concentration oxygen adsorption occurs without water adsorption nearby, releases some electrons into the semiconductor surface, playing the role of a donor. The interaction between CO2 and the local La0.875Ca0.125FeO3(010) surface with pre-adsorption of H2O nearby results in some electron transfer from the La0.875Ca0.125FeO3 surface to CO2, which is responsible for the weakening of CO2 response at high temperatures for La0.875Ca0.125FeO3 with an increase of room-temperature relative humidity.
 
We present a study of the charge-state behavior of the Li-ion battery cathode materials Li(x)MnPO(4) and Li(x)Mn(0.9)Fe(0.1)PO(4) using X-ray absorption spectroscopy (XAS) and resonant inelastic X-ray scattering (RIXS). A set of six identical battery cathodes for each material have been cycled and left in different charge states in the range of x = 0.2…1.0 before disassembly in an Ar glove box. Unexpectedly, we find that the Mn 3d-bands are almost inert to the delithiation process, suggesting that Mn ions participate to a very small extent in the charge compensation process. In Li(x)Mn(0.9)Fe(0.1)PO(4) the Fe 3d-band shows much more response to delithiation than the Mn 3d-band. The O 2p-band hybridizes with the bands of the other ions in Li(x)MnPO(4) and Li(x)Mn(0.9)Fe(0.1)PO(4) and thus, indirectly, carries useful information about the effects of delithiation at all ion sites. We conclude that the redox reactions during lithiation/delithiation of these materials are complex and involve repopulation of charges for all constituent elements.
 
The structure of laser glasses in the system (Y(2)O(3))(0.2){(Al(2)O(3))(x))(B(2)O(3))(0.8-x)} (0.15 ≤ x ≤ 0.40) has been investigated by means of (11)B, (27)Al, and (89)Y solid state NMR as well as electron spin echo envelope modulation (ESEEM) of Yb-doped samples. The latter technique has been applied for the first time to an aluminoborate glass system. (11)B magic-angle spinning (MAS)-NMR spectra reveal that, while the majority of the boron atoms are three-coordinated over the entire composition region, the fraction of three-coordinated boron atoms increases significantly with increasing x. Charge balance considerations as well as (11)B NMR lineshape analyses suggest that the dominant borate species are predominantly singly charged metaborate (BO(2/2)O(-)), doubly charged pyroborate (BO(1/2)(O(-))(2)), and (at x = 0.40) triply charged orthoborate groups. As x increases along this series, the average anionic charge per trigonal borate group increases from 1.38 to 2.91. (27)Al MAS-NMR spectra show that the alumina species are present in the coordination states four, five and six, and the fraction of four-coordinated Al increases markedly with increasing x. All of the Al coordination states are in intimate contact with both the three- and the four-coordinate boron species and vice versa, as indicated by (11)B/(27)Al rotational echo double resonance (REDOR) data. These results are consistent with the formation of a homogeneous, non-segregated glass structure. (89)Y solid state NMR spectra show a significant chemical shift trend, reflecting that the second coordination sphere becomes increasingly "aluminate-like" with increasing x. This conclusion is supported by electron spin echo envelope modulation (ESEEM) data of Yb-doped glasses, which indicate that both borate and aluminate species participate in the medium range structure of the rare-earth ions, consistent with a random spatial distribution of the glass components.
 
Four compositions of Ba(1-x)Sr(x)Co(1-y)Fe(y)O(3-delta) were studied for phase, oxygen uptake-release, and transition metal (TM) oxidation states after solid state processing and with in situ heating from 300 to 1273 K in air. X-Ray diffraction showed that all compositions except one had the cubic perovskite structure at all temperatures; that with x, y = 0.2 was a mixture as prepared, becoming predominantly cubic at high temperature. Thermogravimetry showed a reversible oxygen absorption-desorption of approximately +/-1% from 700 to 1273 K. X-Ray absorption and Mössbauer spectroscopy showed a majority TM(3+) valence, with at most 40% TM(4+). Up to a temperature of 1073 K, the TM(4+) was reduced to TM(3+). Further heating of the composition with x, y = 0.2 to 1233 K resulted in the reduction of Co(3+) to Co(2+). Results from room temperature measurements confirm the thermally activated carrier hopping mechanism with charge fluctuations, while the high temperature delocalized carrier conductivity occurs with a small amount of TM reduction and without phase change for the initially cubic samples.
 
Composite Ni-YSZ fuel electrodes are able to operate only under strongly reducing conditions for the electrolysis of CO(2) in oxygen-ion conducting solid oxide electrolysers. In an atmosphere without a flow of reducing gas (i.e., carbon monoxide), a composite fuel electrode based on redox-reversible La(0.2)Sr(0.8)TiO(3+δ) (LSTO) provides a promising alternative. The Ti(3+) was approximately 0.3% in the oxidized LSTO (La(0.2)Sr(0.8)TiO(3.1)), whereas the Ti(3+) reached approximately 8.0% in the reduced sample (La(0.2)Sr(0.8)TiO(3.06)). The strong adsorption of atmospheric oxygen in the form of superoxide ions led to the absence of Ti(3+) either on the surface of oxidized LSTO or the reduced sample. Reduced LSTO showed typical metallic behaviour from 50 to 700 °C in wet H(2); and the electrical conductivity of LSTO reached approximately 30 S cm(-1) at 700 °C. The dependence of [Ti(3+)] concentration in LSTO on P(O(2)) was correlated to the applied potentials when the electrolysis of CO(2) was performed with the LSTO composite electrode. The electrochemical reduction of La(0.2)Sr(0.8)TiO(3+δ) was the main process but was still present up to 2 V at 700 °C during the electrolysis of CO(2); however, the electrolysis of CO(2) at the fuel electrode became dominant at high applied voltages. The current efficiency was approximately 36% for the electrolysis of CO(2) at 700 °C and a 2 V applied potential.
 
Molecular dynamics and Monte Carlo simulations have been performed for characterizing the structure of the 0.2 and 1 molar aqueous trimethylammonium chloride solutions. Atomic charges were derived through the CHELPG and RESP fits to the molecular electrostatic potentials calculated for the cation in water at the IEF-PCM/B3LYP level using the 6-31G* and 6-311++G** basis sets. Maxima and minima of the calculated radial distribution functions were not sensitive to the four atomic charge sets. Simulated solution structures suggest non-negligible solute-solute interactions and remarkable inhomogeneity at both concentrations. This means that equilibrium ratios, derived for conformers/tautomers by means of ab initio calculations with the IEF-PCM continuum dielectric solvent model, should be corrected for free energy changes following solute association when compared to experimental data obtained for the 0.1-1 molar aqueous solutions.
 
Monte Carlo simulations have been performed for characterizing the 0.22 and 1 molar aqueous dimethylammonium chloride solutions at p = 1 atm and T = 310 K. On the basis of potential of mean force curves for the two systems with increasing concentrations, the N···N separations of about 8.7 and 7.5 Å correspond to a vague and a more pronounced minimum, respectively. Nitrogen separations at the minima are considerably smaller than those what the cations would take if the solutes comprised uniform local solute density on a microscale. The derived N by N coordination numbers predict non-negligible cation association and concomitant inhomogeneity for the studied systems. The calculated N···N distance distribution in the molar solution indicates that about 12% of the N···N separations are shorter than 8.5 Å compared with R(N···N) = 11.84 Å corresponding to the closest distance in a uniform cation local density. Despite a global minimum of -1.79 ± 0.63 kcal mol(-1) at N···Cl separation of 3.24 Å, obtained from the pmf for the 0.22 molar model, the N by Cl coordination number is only 0.14 in the first coordination shell, suggesting frequent disruption of an N-H(+)···Cl(+) hydrogen bond in a relatively dilute solution. The expression for the chemical potential of a solute includes a concentration-dependent activity coefficient, whose varying values are expected to reflect the different degrees of solute association in the 0.2-1 molar range. Theoretical follow-up of the changes in the activity coefficient values is difficult, thus calculation of the K(c) equilibrium constant has been proposed by considering 1 molar solutions as the standard state.
 
A series of group IIIA metal ion electron acceptors doped into Sr(0.25)H(1.5)Ta(2)O(6)·H(2)O (HST) samples have been prepared by an impregnation and calcination method for the first time. The samples are characterized by XRD, TEM, DRS and XPS. The variations in the electronic structure and photoelectric response after metal ion doping are investigated by theoretical calculations and photocurrent experiments, respectively. Results show that the metal ions can be efficiently incorporated into the HST crystal structure, which is reflected in the lattice contraction. Meanwhile, the photoabsorption edges of the metal-doped HST samples are red shifted to a longer wavelength. Taking into account the ionic radii and electronegativities of the dopants, as well as the XRD and XPS results, it is concluded that Ta(5+) ions may be partially substituted by the Al(3+) and Ga(3+) ions in the framework, while In(3+) ions are the favourable substitutes for Sr(2+) sites in the cavity. The first-principles DFT calculations confirm that the variation of the band structure is sensitive to the type of group IIIA metal ion. Introducing the dopant only at the Ta site induces an obvious variation in the band structure and the band gap becomes narrow. Meanwhile, an ''extra step'' appeared in the band gap, which can trap photogenerated electrons from the valance band (VB) and could enhance the charge mobility and the photocurrent. For the photocatalytic degradation of methyl orange in an aqueous solution and in benzene in the gas phase, the doped samples show superior photocatalytic activities compared with both undoped samples and TiO(2). The enhanced photocatalytic activities can be well explained by their electronic structure, photoabsorption performance, photoelectric response, and the concentration of the active species. Due to the fact that Ga ion doping can create an acceptor impurity level and change the electronic band, efficiently narrowing the band gap, the Ga-doped sample shows the highest photocatalytic activity.
 
The fundamental electrical transport properties including ionic conductivity, dielectric constants, loss tangent, and relaxation time constants of Li-excess garnet-type cubic (space group Ia3[combining macron]d) Li5+2xLa3Ta2-xYxO12 (x = 0.25, 0.5 and 0.75) have been studied in the temperature range of -50 to 50 °C using electrochemical AC impedance spectroscopy. A correlation has been established between the excess Li content and the Li(+) ion migration pathways. The loss tangent (tan δ) for all samples exhibits a relaxation peak corresponding to the dielectric loss because of dipolar rotations due to Li(+) migration. Comparing the modulus analysis of Li-excess garnets with fluorite-type oxygen ion conductors, we propose the local migration of Li(+) ions between octahedral sites around the "immobile" Li(+) ions in tetrahedral (24d) sites. In the samples with x = 0.25 and 0.5, Li(+) ions seem to jump from one octahedral (96h) site to another bypassing the tetrahedral (24d) site between them (path A), both in local and long-range order migration processes, with activation energies of ∼0.69 and 0.54 eV, respectively. For the x = 0.75 member, Li(+) ions exhibit mainly long-range order migration, with an activation energy of 0.34 eV, where the Li hopping between two octahedral sites occurs through the edge which is shared between the two LiO6 octahedra and a LiO4 tetrahedron (path B). The present AC impedance analysis is consistent with the ab initio theoretical analysis of Li-excess garnets that showed two conduction paths (A and B) for Li ion conduction with different activation energies.
 
The angle-velocity distribution (HOD) of the OH + D(2) reaction at a relative translational energy of 0.28 eV has been calculated using the quasi-classical trajectory (QCT) method on the two most recent potential energy surfaces available (YZCL2 and WSLFH PESs), widely extending a previous investigation of our group. Comparison with the high resolution experiments of Davis and co-workers (Science, 2000, 290, 958) shows that the structures (peaks) found in the relative translational energy distributions of products could not be satisfactorily reproduced in the calculations, probably due to the classical nature of the QCT method and the importance of quantum effects. The calculations, however, worked quite well for other properties. Overall, both surfaces led to similar results, although the YZCL2 surface is more accurate to describe the H(3)O PES, as derived from comparison with high level ab initio results. The differences observed in the QCT calculations were interpreted considering the somewhat larger anisotropy of the YZCL2 PES when compared with the WSLFH PES.
 
Isobaric crystallization of HDA on heating at 0.51 GPa starting from 77 K. (a), (c) and (e) show the DV versus temperature curves (thick grey lines) obtained on heating at B0.5 (a), B3.5 (c) and B19 (e) K min À1 ; (b), (d) and (f) the X-ray diffractograms (Cu Ka) of the corresponding recovered samples (top), calculated diffractograms (bottom), weight fractions resulting from quantitative phase analysis and tickmarks for the reflection positions. Weight fractions and tickmarks of the different phases are given in the same order from top to bottom. Asterisks denote ice Ih impurity by condensed water vapour. The volume changes for the crystallization of HDA (DV c ) are determined by extrapolation (thin black lines) to the midpoint temperature of the transition and are indicated by thick black bars in (a), (c) and (e). The ordinate bar in (a) indicates a DV value of 1 mm 3 , and the same bar holds for (c) and (e). Data from an isobaric heating experiment without ice sample at 0.51 GPa and a heating rate of B3.5 K min À1 is shown as dashed black curve in (c). (Figure is adapted from Fig. 2 in ref. 57).
(a) Midpoint temperatures ( T c ) of the different crystallization experiments used for the calculation of the density values of rHDA 
Raman spectra ( l 0 = 632.8 nm) of recovered (r)HDA samples as a function of the annealing temperature ( T A ) at annealing pressures ( p A ) of 0.51 and 1.17 GPa. Spectra were recorded at 80 K and 10 mbar. (a) and (c) show spectra of the coupled O–H stretching transition ( n (OH)) of 
Raman spectra (l 0 = 632.8 nm) of recovered rHDA samples recorded at 80 K and 10 mbar. Coupled O-H stretching transitions (n(OH)) and decoupled O-D stretching transitions (n(OD)) are shown in (a) and (c), respectively. Spectra in (a) and (c) were scaled to approximately same peak intensities and shifted vertically for clarity. T A and p A are indicated for each spectrum. Peak positions of the spectra in (a) and (c) are plotted together with additional data in (b) and (d) against the annealing pressure (p A ). The peak positions of the coupled O-H and decoupled O-D stretching transition of (not annealed) HDA are indicated by dotted lines in (b) and (d).
X-ray diffractograms (Cu Ka) of recovered HDA and rHDA samples recorded at B90 K in vacuo. (a) Shows the diffractograms of HDA (1) and rHDA samples annealed at 0.31 GPa/136 K (2), at 0.51 GPa/146 K (3), at 0.81 GPa/152 K (4), at 1.17 GPa/161 K (5) and at 1.61 GPa/170 K (6). Tickmarks indicate the reflection positions of ice Ih which originates from condensation of water vapour onto the sample holder. Peak position (E) and ''full width at half height'' (J) of the strongest diffraction peaks in diffractograms (1-6) are shown in (b) as a function of p A. The peak position and fwhh of HDA are indicated by a solid and a dashed horizontal line, respectively.
We report in situ density values of amorphous ice obtained between 0.3 and 1.9 GPa and 144 to 183 K. Starting from high-density amorphous ice made by pressure-amorphizing hexagonal ice at 77 K, samples were heated at a constant pressure until crystallization to high-pressure ices occurred. Densities of amorphous ice were calculated from those of high-pressure ice mixtures and the volume change on crystallization. In the density versus pressure plot a pronounced change of slope occurs at approximately 0.8 GPa, with a slope of 0.21 g cm(-3) GPa(-1) below 0.8 GPa and a slope of 0.10 g cm(-3) GPa(-1) above 0.8 GPa. Both X-ray diffractograms and Raman spectra of recovered samples show that major structural changes occur up to approximately 0.8 GPa, developing towards those of very high-density amorphous ice reported by (T. Loerting, C. Salzmann, I. Kohl, E. Mayer and A. Hallbrucker, Phys. Chem. Chem. Phys., 2001, 3, 5355) and that further increase of pressure has only a minor effect. In addition, the effect of annealing temperature (T(A)) at a given pressure on the structural changes was studied by Raman spectra of recovered samples in the coupled O-H and decoupled O-D stretching band region: at 0.5 GPa structural changes are observed between approximately 100-116 K, at 1.17 GPa between approximately 121-130 K. Further increase of T(A) or of annealing time has no effect, thus indicating that the samples are fully relaxed. We conclude that mainly irreversible structural changes between 0.3 to approximately 0.8 GPa lead to the pronounced increase in density, whereas above approximately 0.8 GPa the density increase is dominated to a large extent by reversible elastic compression. These results seem consistent with simulation studies by (R. Martonàk, D. Donadio and M. Parrinello, J. Chem. Phys., 2005, 122, 134501) where substantial reconstruction of the topology of the hydrogen bonded network and changes in the ring statistics from e.g. mainly six-membered to mainly nine-membered rings were observed on pressure increase up to 0.9 GPa and further pressure increase had little effect.
 
The effectiveness of dynamic nuclear polarization (DNP) as a tool to enhance the sensitivity of liquid state NMR critically depends on the choice of the optimal polarizer molecule. In this study the performance of (15)N labelled Frémy's salt as a polarizing agent in Overhauser DNP is investigated in detail at X-band (0.35 T, 9.7 GHz EPR, 15 MHz (1)H NMR) and compared to that of TEMPONE-D,(15)N employed in previous studies. Both radicals provide similar maximum enhancements of the solvent water protons under similar conditions but a different saturation behaviour. The factors determining the enhancement and effective saturation were measured independently by EPR, ELDOR and NMRD and are shown to fulfil the Overhauser equation. In particular, following the theory of EPR saturation we provide analytical solutions for the dependence of the enhancement on the microwave field strength in terms of saturation transfer between two coupled hyperfine lines undergoing spin exchange. The negative charge of the radical in Frémy's salt solutions can explain the peculiar properties of this polarizing agent and indicates different suitable application areas for the two types of nitroxide radicals.
 
This work has initiated an investigation on the electrochemical behaviors and the structure changes of the composite electrode 0.3Li(2)MnO(3)·0.7LiMn(1/3)Ni(1/3)Co(1/3)O(2) when charged with different cut-off voltages. It is found that the charge cut-off voltages could not only affect the capacity property and coulombic efficiency, but also alter the electrode kinetics of the composite. As a consequence, the electrochemical activation of the composite electrode is highly dependent on the charge cut-off voltages: when the charge cut-off voltage is higher than 4.5 V, the inert component Li(2)MnO(3) in the composite electrode is completely activated. At the meanwhile, there occurred an irreversible oxygen loss during the initial charge process, which yielded a hollow sphere in the electrode. Regardless of charge voltages, Mn ions in the composite electrode were presented in an oxidation state of +4, while Co(2+) ions were detected at the surface of the electrode when cycled at low voltages. Ni ions in the composite could react with organic or inorganic species and then cover the surface of the cycled electrode.
 
Fig.S1 A representative XRD pattern of precursor MnCO 3 prepared by co-precipitation. The diffraction peaks could be well indexed to a pure rhombohedral phase of MnCO 3 (JCPDS No. 44-1472). 
Fig. S4 XRD patterns for the 0.3Li 2 MnO 3 @BULLET0.7LiNi 0.5 Mn 0.5 O 2 products prepared by a sol-gel method (nanoparticles) and a solid state reaction (bulk), respectively. 
Fig. S5 FESEM images for the electrodes before cycling (a) and after 200 discharge/charge cycles at 55 ºC at the current densities of (b) 50 (c) 100 (d) 200, and 
Fig. S5 FESEM images for the electrodes before cycling (a) and after 200 discharge/charge cycles at 55 ºC at the current densities of (b) 50 (c) 100 (d) 200, and (e) 500 mA g-1 , respectively. The smaller particles are acetylene black.
Hollow 0.3Li(2)MnO(3)·0.7LiNi(0.5)Mn(0.5)O(2) microspheres are synthesized on a large scale through a simple in situ template-sacrificial route. Starting from porous MnO(2) microspheres, the hollow microspheres assembled with 0.3Li(2)MnO(3)·0.7LiNi(0.5)Mn(0.5)O(2) nanocrystals are formed by a nanoscale Kirkendall effect. The nanocrystal-assembled hollow 0.3Li(2)MnO(3)·0.7LiNi(0.5)Mn(0.5)O(2) microspheres exhibit a highly reversible capacity as high as 295 mAh g(-1) over 100 cycles and excellent rate capability (125 mAh g(-1) at 1000 mA g(-1)). Benefitting from a unique hollow and nanocrystalline architecture, the as-formed hollow microspheres show much enhanced high-temperature (55 °C) electrochemical performances, compared with the products obtained by conventional sol-gel/solid-state reaction methods. This work demonstrates that a fabrication strategy based on the present in situ template-sacrificial approach offers a new method for the design of high-performance cathode materials with hollow interiors for Li-ion battery applications.
 
Structure of Cu(hfipbb)(H 2 hfipbb) 0.5 in [010] direction (left) and [10-1] direction (right). The right structure schematically shows the locations of the one-dimensional channels available for molecular adsorption and diffusion.
Single component and mixture adsorption isotherms of CH 4 and CO 2 at 298 K in Cu(hfipbb)(H 2 hfipbb) 0.5 . Mixture isotherms are for an equimolar bulk phase mixture. For CO 2 , results are shown with two different methods for assigning partial charges to the MOF framework atoms.  
Minimum potential energy of CH 4 as a function of position along the pore axis in Cu(hfipbb)(H 2 hfipbb) 0.5 , calculated using the methods described in the text.  
Self diffusivities of CO 2 at 298 K in CO 2 /CH 4 mixtures as a function of CO 2 loading at various compositions with single component diffusivity of CO 2 in Cu(hfipbb)(H 2 hfipbb) 0.5 .  
The identification of membrane materials with high selectivity for CO(2)/CH(4) mixtures could revolutionize this industrially important separation. We predict using computational methods that a metal organic framework (MOF), Cu(hfipbb)(H(2)hfipbb)(0.5), has unprecedented selectivity for membrane-based separation of CO(2)/CH(4) mixtures. Our calculations combine molecular dynamics, transition state theory, and plane wave DFT calculations to assess the importance of framework flexibility in the MOF during molecular diffusion. This combination of methods should also make it possible to identify other MOFs with attractive properties for kinetic separations.
 
Layered Li2MnO3·3LiNi0.5-xMn0.5-xCo2xO2 (x = 0, 0.05, 0.1, 0.165) microspheres with Mn-rich core were successfully synthesized by a simple two-step precipitation calcination method and intensively evaluated as cathode materials for lithium ion batteries. The X-ray powder diffractometry (XRD) results indicate that the growth of Li2MnO3-like regions is impeded due to the presence of cobalt (Co) in the material. The field-emission scanning electron microscopy (FESEM) data reveal the core-shell-like structure with a Mn-rich core in the as-prepared particles. The charge-discharge testing reveals that the capacity is markedly improved by adding Co. The activation of the cathode after Co doping becomes easier and can be accomplished completely when charged to 4.6 V at the C/40 rate in the initial cycle. Superior electrochemical performances are obtained for samples with x = 0.05 and 0.1. The corresponding initial discharge capacities are separately 281 and 285 mA h g(-1) at C/40 between 2 and 4.6 V at room temperature. After 250 cycles at C/2, the respective capacity retentions are 71.2% and 70.4%, which are better compared to the normal Li-excess Li2MnO3·3LiNi0.4Mn0.4Co0.2O2 sample with a uniform distribution of Mn element in the particles. The initial discharge capacities of both samples are approximately 250 mA h g(-1) at a rate of C/2 between 2 and 4.6 V at 55 °C after activation. Furthermore, the samples are investigated by electrochemical impedance spectroscopy (EIS) at room and elevated temperature, revealing that the key factor affecting electrochemical performance is the charge transfer resistance in the particles.
 
Catalysts of 1 wt% copper oxide supported on cerium oxide or cerium-terbium mixed oxides are comparatively examined with respect to their redox and catalytic properties for CO oxidation. Characterization of the catalysts had shown that they contain highly dispersed CuO-type entities on the corresponding nanostructured fluorite supports with copper dispersion increasing with increasing amounts of terbium in the support. In contrast, the CO oxidation catalytic activity decreases with increasing amounts of terbium in the support. On the basis of operando-DRIFTS experiments, one of the factors that could explain such behaviour is related to the greater difficulty in generating reduced copper sites active for the reaction in the presence of terbium, which in turn is evidenced to constitute an induction stage. Analysis of the redox properties is complemented by XPS which confirms the greater resistance to copper reduction in the presence of terbium.
 
Here we report the observation of electron delocalization in nano-dimension xLiFePO(4):(1 - x)FePO(4) (x = 0.5) using high temperature, static, (31)P solid state NMR. The (31)P paramagnetic shift in this material shows extreme sensitivity to the oxidation state of the Fe center. At room temperature two distinct (31)P resonances arising from FePO(4) and LiFePO(4) are observed at 5800 ppm and 3800 ppm, respectively. At temperatures near 400 °C these resonances coalesce into a single narrowed peak centered around 3200 ppm caused by the averaging of the electronic environments at the phosphate centers, resulting from the delocalization of the electrons among the iron centers. (7)Li MAS NMR spectra of nanometre sized xLiFePO(4):(1 - x)FePO(4) (x = 0.5) particles at ambient temperature reveal evidence of Li residing at the phase interface between the LiFePO(4) and FePO(4) domains. Moreover, a new broad resonance is resolved at 65 ppm, and is attributed to Li adjacent to the anti-site Fe defect. This information is considered in light of the (7)Li MAS spectrum of LiMnPO(4), which despite being iso-structural with LiFePO(4) yields a remarkably different (7)Li MAS spectrum due to the different electronic states of the paramagnetic centers. For LiMnPO(4) the higher (7)Li MAS paramagnetic shift (65 ppm) and narrowed isotropic resonance (FWHM ≈ 500 Hz) is attributed to an additional unpaired electron in the t(2g) orbital as compared to LiFePO(4) which has δ(iso) = -11 ppm and a FWHM = 9500 Hz. Only the delithiated phase FePO(4) is iso-electronic and iso-structural with LiMnPO(4). This similarity is readily observed in the (7)Li MAS spectrum of xLiFePO(4):(1 - x)FePO(4) (x = 0.5) where Li sitting near Fe in the 3+ oxidation state takes on spectral features reminiscent of LiMnPO(4). Overall, these spectral features allow for better understanding of the chemical and electrochemical (de)lithiation mechanisms of LiFePO(4) and the Li-environments generated upon cycling.
 
Fig. S1. Simulated XRD patterns of LiNi 0.5 Mn 0.5 O 2 with different degrees of cation mixing. 
Fig. S2. In situ XRD patterns for LiNi 0.5 Mn 0.5 O 2 (003) with a large degree of cation mixing. The peak shifts during anodic polarization from the initial voltage to 5.0 V (a) and the shift during cathodic polarization to 3.0 V (b). [Ref. 15] 
Fig. S3. Voltage dependence of LiNi 0.5 Mn 0.5 O 2 (003) lattice parameters with a large degree of cation mixing. [Ref. 15] 
Surface and bulk structural changes of LiNi(0.5)Mn(0.5)O(2) were investigated during electrochemical reaction using synchrotron X-ray scattering and a restricted reaction plane consisting of two-dimensional epitaxial-film electrodes. The changes in bulk structure confirmed lithium diffusion through the (110) surface, which was perpendicular to the two-dimensional (2D) edges of the layered structure. No (de)intercalation reaction was observed through the (003) surface at voltages of 3.0-5.0 V. However, intercalation did proceed through the (003) plane below 3.0 V, indicating unusual three-dimensional (3D) lithium diffusion in the over-lithiated 2D structure. During the electrochemical process, the surface of the electrode showed different structure changes from those of the bulk structure. The reaction mechanism of the intercalation electrodes for lithium batteries is discussed on the basis of surface and bulk structural changes.
 
In olivines, (A,B)(2)SiO(4), the A and B cations are distributed over two non-equivalent sites of octahedral coordination, M1 and M2. In the case of temperature dependent cation distributions, the kinetics of cation redistribution between these two sublattices can be studied by means of temperature-jump experiments. In situ experiments of this type are reported for a cobalt-containing olivine single crystal, (Co(0.6)Mg(0.4))(2)SiO(4). The relaxation experiments were performed by means of optical spectroscopy under in situ conditions in the temperature range between 480 and 690 degrees C yielding an activation energy of about 2.0 eV. The results are discussed in the framework of microscopic models of cation sublattice exchange. Implications for quench experiments are addressed.
 
The study of 28 porous carbons shows that the specific capacitance in the electrolyte (C(2)H(5))(4)NBF(4)/acetonitrile is relatively constant between 0.7 and 15 nm (0.094 ± 0.011 F m(-2)). The increase in pores below 1 nm and the lower values between 1 and 2 nm reported earlier are not observed in the present work.
 
The present study reports on the synthesis and the electrochemical behavior of Na(0.71)CoO(2), a promising candidate as cathode for Na-based batteries. The material was obtained in two different morphologies by a double-step route, which is cheap and easy to scale up: the hydrothermal synthesis to produce Co(3)O(4) with tailored and nanometric morphology, followed by the solid-state reaction with NaOH, or alternatively with Na(2)CO(3), to promote Na intercalation. Both products are highly crystalline and have the P2-Na(0.71)CoO(2) crystal phase, but differ in the respective morphologies. The material obtained from Na(2)CO(3) have a narrow particle length (edge to edge) distribution and 2D platelet morphology, while those from NaOH exhibit large microcrystals, irregular in shape, with broad particle length distribution and undefined exposed surfaces. Electrochemical analysis shows the good performances of these materials as a positive electrode for Na-ion half cells. In particular, Na(0.71)CoO(2) thin microplatelets exhibit the best behavior with stable discharge specific capacities of 120 and 80 mAh g(-1) at 5 and 40 mA g(-1), respectively, in the range 2.0-3.9 V vs. Na(+)/Na. These outstanding properties make this material a promising candidate to construct viable and high-performance Na-based batteries.
 
The new compound MnF2-x(OH)x (x ∼ 0.8) was synthesized by a hydrothermal route from a 1 : 1 molar ratio of lithium fluoride and manganese acetate in an excess of water. The crystal structure was determined using the combination of single crystal X-ray and neutron powder diffraction measurements. The magnetic properties of the title compound were characterized by magnetic susceptibility and low-temperature neutron powder diffraction measurements. MnF2-x(OH)x (x ∼ 0.8) crystallizes with orthorhombic symmetry, space group Pnn2 (no. 34), a = 4.7127(18), b = 5.203(2), c = 3.2439(13) Å, V = 79.54(5) Å(3) and Z = 2. The crystal structure is a distorted rutile-type with [Mn(F,O)4] infinite edge-sharing chains along the c-direction. The protons are located in the channels and form O-HF bent hydrogen bonds. The magnetic susceptibility measurements indicate an antiferromagnetic ordering at ∼70 K and the neutron powder diffraction measurements at 3 K show that the ferromagnetic chains with spins parallel to the c-axis are antiferromagnetically coupled to each other, similarly to the magnetic structure of tetragonal rutile-type MnF2 with isoelectronic Mn(2+). MnF2-x(OH)x (x ∼ 0.8) is expected to be of great interest as a positive electrode for Li cells if the protons could be exchanged for lithium.
 
Buffer-gas pressure broadening for the (3 00 1)III ← (0 0 0) band of CO2 in the 1600 nm region was investigated with continuous wave cavity ring-down spectroscopy within the temperature range 263–326 K. The measured absorption profiles were analyzed with Voigt functions. Pressure broadening coefficient, γ(gas), and the temperature dependent parameter (broadening exponent), n, were determined for a variety of buffer gases: N2, O2, He, Ne, Ar, Kr and Xe. γ(air) values estimated subsequently are 0.096(2) for R(0), 0.085(5) for P(8), 0.075(2) for P(16), 0.070(4) for P(26), and 0.069(2) for P(38) in units of cm−1 atm−1, where numbers in parentheses are one standard deviation in units of the last digits quoted. n(air) values are 0.77(4) for R(0), and 0.73(11) for P(8).
 
The quantum superposition principle, a key distinction between quantum physics and classical mechanics, is often perceived as a philosophical challenge to our concepts of reality, locality or space-time since it contrasts with our intuitive expectations with experimental observations on isolated quantum systems. While we are used to associating the notion of localization with massive bodies, quantum physics teaches us that every individual object is associated with a wave function that may eventually delocalize by far more than the body's own extension. Numerous experiments have verified this concept at the microscopic scale but intuition wavers when it comes to delocalization experiments with complex objects. While quantum science is the uncontested ideal of a physical theory, one may ask if the superposition principle can persist on all complexity scales. This motivates matter-wave diffraction and interference studies with large compounds in a three-grating interferometer configuration which also necessitates the preparation of high-mass nanoparticle beams at low velocities. Here we demonstrate how synthetic chemistry allows us to prepare libraries of fluorous porphyrins which can be tailored to exhibit high mass, good thermal stability and relatively low polarizability, which allows us to form slow thermal beams of these high-mass compounds, which can be detected using electron ionization mass spectrometry. We present successful superposition experiments with selected species from these molecular libraries in a quantum interferometer, which utilizes the diffraction of matter-waves at an optical phase grating. We observe high-contrast quantum fringe patterns of molecules exceeding a mass of 10 000 amu and having 810 atoms in a single particle.
 
Computational methods have been used in the past to generate large libraries of hypothetical zeolite structures, but to date analysis of these structures has typically been limited to relatively simple physical properties such as density. We use efficient methods to analyze the adsorption and diffusion properties of simple adsorbate molecules in a library of >250,000 hypothetical silica zeolites that was generated previously by Deem and co-workers (J. Phys. Chem. C, 2009, 113, 21353). The properties of this library of materials are compared to the complete set of ∼190 zeolites that have been identified experimentally. Our calculations provide information on the largest cavities available in each material for adsorption, and the size of the largest molecules that can diffuse through each material. For a subset of ∼8000 materials, we computed the Henry's constant and diffusion activation energy for adsorbed CH(4) and H(2). We show that these calculations provide a useful screening tool for considering large collections of nanocrystalline materials and choosing materials with particular promise for more detailed modeling.
 
(a) Boron nitride supercell. (b)–(e) Optimized geometries for the sequential adsorption of H atoms on the hexagonal BN sheet. Cross sections of the electronic charge density in the BN(0001) plane are also plotted (in units of e A ˚ À3 ).  
(a) Electronic charge density cross section of the most stable H–H pair, i.e. the orthodimer, adsorbed on BN(0001) (in units of e A ˚ À3 ), plotted in a plane perpendicular to the plane of the BN sheet. Highest occupied molecular orbital (HOMO) of (b) a single H atom and (c) the energetically favorable H triplet adsorbed on the BN sheet.  
Binding energy as a function of the pair interatomic separation for the adsorption of (a) the second, (b) third, and (c) fourth H atoms on BN(0001). Indices correspond to the site positions given in Fig. 1(a). For (a), (b) and (c), hydrogen atoms are initially adsorbed at B 0 , B 0 –N 1 , and B 0 –N 1 –B 6 top sites, respectively. The different colored traces refer to the separations between the last adsorbed hydrogen atom, denoted by H, at the stable/metastable site positions investigated and the hydrogen atoms initially adsorbed on the BN sheet.  
Using first-principles calculations, we have studied successive adsorption of hydrogen atoms on a sp2-bonded boron nitride graphitic sheet. Our calculations show that clustering proceeds through the creation of contiguous H-H orthodimer structures stabilizing the H adsorbate cluster on the BN(0001) surface, leading eventually to the formation of hydrogen-contiguous boat-shaped quartets.
 
Epitaxial graphene on Ru(0001) was studied by means of large-scale density functional theory (DFT) calculations. The results agree well with scanning tunneling microscopy experiments. In contrast to the current understanding, we show that the measured corrugation originates mainly from a geometric buckling of the graphene sheet, induced by alternating weak and strong chemical interactions with the Ru support. In the strong contact regions, charge transfer is evidenced and the opening of a considerable band gap in the graphene is found.
 
(a) Ru(0001) surface after deposition of 0.5 ML Pt at 300 K (217217 nm 2 , I t =0.56 nA, U t =+1V); (b) the same surface after annealing to 1350 K (image size 217217 nm 2 , I t =0.56 nA, U t =+1V); (c) atomically resolved image of the terrace showing 50% Pt atoms (dark due to chemical contrast) (1313 nm 2 , I t =28nA, U t =+52 mV); (d) edge of an island (see frame marked in (b)), showing an identical composition of island and underlying terrace (1313 nm 2 , I t =28nA, U t =+52 mV).
Atomically resolved STM images of Pt x Ru 1-x / Ru(0001) surface alloys, Pt content as indicated above the images. Tunnel parameters (I t and U t ): (a) 71 nA, 8 mV; 
The formation of PtRu surface alloys by deposition of submonolayer Pt films on a Ru(0001) substrate and subsequent annealing to about 1350 K and the distribution of the Pt atoms in the surface layer were investigated by scanning tunneling microscopy. Quantitative statistical analysis reveals (i) negligible losses of Pt into subsurface regions up to coverages close below 1 monolayer, (ii) a homogeneous distribution of the Pt atoms over the surface, and (iii) the absence of a distinct long-range or short-range order in the surface layer. In addition, the density of specific adsorption ensembles is analyzed as a function of Pt surface content. Possible conclusions on the process for surface alloy formation are discussed. The results are compared with the properties of PtRu bulk alloys and the findings in previous adsorption studies on similar surface alloys (H. Rauscher, T. Hager, T. Diemant, H. Hoster, F. Bautier de Mongeot and R. J. Behm, Surf. Sci., 2007, 601, 4608; T. Diemant, H Rauscher and R. J. Behm, J. Phys. Chem. C, in press).
 
The interaction of silicene, the silicon counterpart of graphene, with (0001) ZnS surfaces is investigated theoretically, using first-principles simulations. The charge transfer occurring at the silicene/(0001) ZnS interface leads to the opening of an indirect energy band gap of about 0.7 eV in silicene. Remarkably, the nature (indirect or direct) and magnitude of the energy band gap of silicene can be controlled by an external electric field: the energy gap is predicted to become direct for electric fields larger than about 0.5 V Å(-1), and the direct energy gap decreases approximately linearly with the applied electric field. The predicted electric field tunable energy band gap of the silicene/(0001) ZnS interface is very promising for its potential use in nanoelectronic devices.
 
We review here recent progress on epitaxial graphene grown on a SiC substrate. Epitaxial graphene can be easily grown by heating the SiC single crystal in a high vacuum or in an inert gas atmosphere. The SiC surfaces used for graphene growth contain Si- and C-terminated faces. On the Si-face, homogeneous and clean graphene can be grown with a controlled number of layers, and the carrier mobility reaches as high as several m(2) V s(-1), although this is reduced by the presence of the substrate steps. On the C-face, although the number of layers is not homogeneous, twisted bilayer graphene can be grown, which is expected to be the technique of choice to modify the electronic structure of graphene. From the application point of view, graphene on SiC will be the platform used to fabricate high-speed electronic devices and dense graphene nanoribbon arrays, which will be used to introduce a bandgap.
 
The pathways for the reaction of ethanol on model catalysts consisting of Co and CoO films and particles supported on single crystal ZnO(0001) surfaces were studied using X-ray Photoelectron Spectroscopy (XPS) and Temperature Programmed Desorption (TPD). On supported metallic Co films and particles ethanol was found to primarily undergo decarbonylation forming CO, H(2), and adsorbed methyl groups. In contrast, supported CoO particles were found to be largely unreactive toward ethanol. High selectivity to the dehydrogenation product, acetaldehyde, was only observed when the supported Co was partially oxidized and contained both Co(0) and Co(2+). Since acetaldehyde is thought to be a critical intermediate during steam reforming of ethanol (SRE) to produce H(2) and CO(2), the results of this study suggest that partially oxidized Co species provide the active sites for this reaction. This result is consistent with studies of high surface area Co/ZnO catalysts which also suggest that both Co(0) and Co(2+) species are present under typical SRE reaction conditions.
 
The interaction of atomic oxygen and nitrogen on the (0001) surface of corundum (alpha-alumina) is investigated from first-principles by means of periodic density functional calculations within the generalized gradient approximation. A large Al(2)O(3) slab model (18 layers relaxing 10) ended with the most stable aluminium layer is used throughout the study. Geometries, adsorption energies and vibrational frequencies are calculated for several stationary points for two spin states at different sites over an 1 x 1 unit cell. Two stable adsorption minima over Al or in a bridge between Al and O surface atoms are found for oxygen and nitrogen, without activation energies. The oxygen adsorption (e.g., E(ad) = 2.30 eV) seems to be much more important than for nitrogen (e.g., E(ad) = 1.23 eV). Transition states for oxygen surface diffusion are characterized and present not very high-energy barriers. The computed geometries and adsorption energies are consistent with similar adsorption theoretical studies and related experimental data for O, N or alpha-alumina. The present results along with our previous results for beta-cristobalite do not support the assumption of an equal E(ad) for O and N over similar oxides, which is commonly used in some kinetic models to derive catalytic atomic recombination coefficients for atomic oxygen and nitrogen. The magnitude of O and N adsorption energies imply that Eley-Rideal and Langmuir-Hinshelwood reactions with these species will be exothermic, contrary to what happens for beta-cristobalite.
 
The adsorption and desorption of CO on the hydroxylated, O-terminated polar ZnO(0001) surface has been studied using He-atom scattering. The experimental results reveal the formation of a physisorbed ordered CO overlayer. In addition to recording angular distributions of elastically scattered He atoms, also the dynamical properties of the CO overlayer have been investigated using inelastic He-atom scattering. With the aid of electronic structure calculations a loss peak with an energy transfer of 7.2 meV is assigned to the frustrated translation of the CO molecule normal to the surface.
 
The adsorption of fluoroform molecules on a hexagonal ice (0001) surface was studied using static density functional theory (DFT) calculations and Car-Parrinello molecular dynamics (CP-MD) simulations. Extending our previous work on isolated molecules we focus in the present study on the interplay between molecule-molecule and molecule-substrate interactions. Coverages of up to a full monolayer were modeled by introducing two, three and four fluoroform molecules per unit cell of the ice (0001) substrate. Lowest-energy structures of fluoroform aggregates on the ice surface were determined in a systematic search by performing geometry optimizations from a large set of initial configurations chosen by chemical intuition and from snapshots taken from CP-MD simulations. In the vibrational analysis of the optimized geometries both conventional red- and unusual blue-shifting hydrogen bonds were found. The finite temperature stability of the lowest-energy configurations was probed by CP-MD simulations and conformational changes were analyzed in terms of transformations between the global and local minima structures.
 
Top-cited authors
Martin Head-Gordon
  • University of California, Berkeley
Jeng-Da Chai
  • National Taiwan University
Jiaguo Yu
  • China University of Geosciences
Donald Truhlar
  • University of Minnesota Twin Cities
Stefan Grimme
  • University of Bonn