100 research outputs found

    Dynamics of Bulk vs. Nanoscale WS_2: Local Strain and Charging Effects

    Full text link
    We measured the infrared vibrational properties of bulk and nanoparticle WS2_2 in order to investigate the structure-property relations in these novel materials. In addition to the symmetry-breaking effects of local strain, nanoparticle curvature modifies the local charging environment of the bulk material. Performing a charge analysis on the \emph{xy}-polarized E1u_{1u} vibrational mode, we find an approximate 1.5:1 intralayer charge difference between the layered 2H material and inorganic fullerene-like (IF) nanoparticles. This effective charge difference may impact the solid-state lubrication properties of nanoscale metal dichalcogenides.Comment: 6 pages, 5 figure

    Far-infrared optical properties of the pyrochlore spin ice compound Dy2Ti2O4

    Full text link
    Near normal incident far-infrared reflectivity spectra of [111] dysprosium titanate (Dy2Ti2O4) single crystal have been measured at different temperatures. Seven phonon modes (eight at low temperature) are identified at frequency below 1000 cm-1. Optical conductivity spectra are obtained by fitting all the reflectivity spectra with the factorized form of the dielectric function. Both the Born effective charges and the static optical primitivity are found to increase with decreasing temperature. Moreover, phonon linewidth narrowering and phonon modes shift with decreasing temperature are also observed, which may result from enhanced charge localization. The redshift of several low frequency modes is attributed to the spin-phonon coupling. All observed optical properties can be explained within the framework of nearest neighbor ferromagnetic(FM) spin ice model

    Optical Properties of MFe_4P_12 filled skutterudites

    Full text link
    Infrared reflectance spectroscopy measurements were made on four members of the MFe_4P_12 family of filled skutterudites, with M=La, Th, Ce and U. In progressing from M=La to U the system undergoes a metal-insulator transition. It is shown that, although the filling atom induces such dramatic changes in the transport properties of the system, it has only a small effect on lattice dynamics. We discuss this property of the compounds in the context of their possible thermoelectric applications.Comment: Manuscript in ReVTeX format, 7 figures in PostScirpt forma

    Far-infrared vibrational properties of high-pressure-high-temperature C60 polymers and the C60 dimer

    Get PDF
    We report high-resolution far-infrared transmission measurements of the 2 + 2 cycloaddition C-60 dimer and two-dimensional rhombohedral and one-dimensional orthorhombic high-pressure high-temperature C60 polymers. In the spectral region investigated(20-650 cm(-1)), we see no low-energy interball modes, but symmetry breaking of the linked C-60 balls is evident in the complex spectrum of intramolecular modes. Experimental features suggest large splittings or frequency shifts of some IhC60-derived modes that are activated by symmetry reduction, implying that the balls are strongly distorted in these structures. We have calculated the vibrations of all three systems by first-principles quantum molecular dynamics and use them to assign the predominant IhC60 symmetries of observed modes. Pur calculations show unprecedentedly large downshifts of T-1u(2)-derived modes and extremely large splittings of other modes, both of which are consistent with the experimental spectra. For the rhombohedral and orthorhombic polymers, the T-1u(2)-derived mode that is polarized along the bonding direction is calculated to downshift below any T-1u(1)-derived modes. We also identify a previously unassigned feature near 610 cm(-1) in all three systems as a widely split or shifted mode derived from various silent IhC60 vibrations, confirming a strong perturbation model for these linked fullerene structures

    Six-Coordinate Nitrito and Nitrato Complexes of Manganese Porphyrin

    Full text link
    Reaction of small increments of NO2 gas with sublimed amorphous layers of Mn(II)(TPP) (TPP = meso-tetra-phenylporphyrinato dianion) in a vacuum cryostat leads to formation of the 5-coordinate monodentate nitrato complex Mn(III)(TPP)(η(1)-ONO2) (II). This transformation proceeds through the two distinct steps with initial formation of the five coordinate O-nitrito complex Mn(III)(TPP)(η(1)-ONO) (I) as demonstrated by the electronic absorption spectra and by FTIR spectra using differently labeled nitrogen dioxide. A plausible mechanism for the second stage of reaction is offered based on the spectral changes observed upon subsequent interaction of (15)NO2 and NO2 with the layered Mn(TPP). Low-temperature interaction of I and II with the vapors of various ligands L (L = O-, S-, and N-donors) leads to formation of the 6-coordinate O-nitrito Mn(III)(TPP)(L)(η(1)-ONO) and monodentate nitrato Mn(III)(TPP)(L)(η(1)-ONO2) complexes, respectively. Formation of the 6-coordinate O-nitrito complex is accompanied by the shifts of the ν(N═O) band to lower frequency and of the ν(N-O) band to higher frequency. The frequency difference between these bands Δν = ν(N═O) - ν(N-O) is a function of L and is smaller for the stronger bases. Reaction of excess NH3 with I leads to formation of Mn(TPP)(NH3)(η(1)-ONO) and of the cation [Mn(TPP)(NH3)2](+) plus ionic nitrite. The nitrito complexes are relatively unstable, but several of the nitrato species can be observed in the solid state at room temperature. For example, the tetrahydrofuran complex Mn(TPP)(THF)(η(1)-ONO2) is stable in the presence of THF vapors (∼5 mm), but it loses this ligand upon high vacuum pumping at RT. When L = dimethylsulfide (DMS), the nitrato complex is stable only to ∼-30 °C. Reactions of II with the N-donor ligands NH3, pyridine, or 1-methylimidazole are more complex. With these ligands, the nitrato complexes Mn(III)(TPP)(L)(η(1)-ONO2) and the cationic complexes [Mn(TPP)(L)2](+) coexist in the layer at room temperature, the latter formed as a result of NO3(-) displacement when L is in excess

    TORSIONAL FREQUENCIES IN THE FAR INFRARED THE FORM OF THE POTENTIAL CURVE FOR HINDERED INTERNAL ROTATION OF A METHYL GROUP^{\ast}

    No full text
    ^{\ast} This work was supported by grant 16833 from the National Science Foundation.Author Institution: Mellon InstituteTransitions between excited torsional levels (01,12,23,)(0\rightarrow 1, 1\rightarrow 2, 2\rightarrow 3, \ldots) has been measured in the infrared for several compounds. The observed frequencies provide the best test of the potential function for hindered internal rotation which has yet been made. The constants of the usual expression. V(α)=V3(1cos3α)/2+V6(1cos6α)/2V (\alpha) = V_{3} (1-\cos 3\alpha)/2+V_{6} (1-\cos 6\alpha)/2 are: CompoundV3V6CH3CH2Cl1310±10cm110±0.5cm1CH3CH=CH2706±1021±1CH3CF=CH2806±40OCH3CHCH895±89±0.05\begin{array}{ccc}Compound & V_{3} & V_{6}\\ CH_{3}CH_{2}Cl& 1310\pm10 cm^{-1} &10\pm0.5 cm^{-1}\\ CH_{3}CH = CH_{2} & 706 \pm 10 & -21 \pm 1\\ CH_{3}CF = CH_{2} & 806\pm 4 & 0\\ O\\ CH_{3} -CH-CH &895\pm 8 & -9\pm 0.05\\\end{array} Thus V6<3%V_{6} < 3\% of V3V_{3}, and may be either positive or negative. Frequencies are also given for propylene sulfide, but lack of structural information prevents further interpretation. Pitzer and Hollenberg's earlier results for CH3CCl3CH_{3}CCl_{3} could not be confirmed. The spectra for all six compounds were examined from 100 to 435cm1435 cm^{-1}, allowing some bending frequencies to be observed
    corecore